Archives

Menopause and Hormone Replacement

ABSTRACT

Menopause, defined as twelve months after a woman’s final menstrual period, is a natural event that marks the end of spontaneous ovulation and thus reproductive capabilities. In the Western world, the average age of menopause is 51 years.

During the time preceding and following the menopause, many women experience symptoms including hot flashes, vaginal irritation, trouble sleeping, fatigue, and weight gain. Each woman experiences perimenopause uniquely; although menopause symptoms may represent minor inconveniences for some women, other women find these symptoms more disruptive. This period of a women’s life also coincides with the time she is more likely to develop diseases associated with advancing age such as osteoporosis, cardiovascular disease, and cancer.

The clinical use of estrogens to treat menopausal symptoms was first evaluated in the late 1920s. By 1928, the first commercially available injectable estrogen was developed; and by 1942, the first oral formulation of estrogen was marketed. Over the years, data from clinical studies have refined the indications for hormone therapy. For example, estrogen remains the most effective therapy for hot flashes. However, it has also recently been established that estrogen is not appropriate to prevent chronic disease.

Thus, the challenges to clinicians and patients who consider prescribing and using hormone therapy are: whether to treat; with which agent (formulation, dose, delivery method); and for how long.

THE MENOPAUSAL TRANSITION

Early in the menopausal transition (which starts in the late 40s and lasts about 4 years), ovarian estradiol production is erratic and associated with irregular menstrual cycle
length. FSH levels rise in response to a decrease in levels of inhibin, a protein produced by the granulosa cells (1-7). An FSH level > 10 mIU/ml (measured between cycle day 2-5) indicates ovarian aging. As the final menstrual period approaches, estradiol secretion diminishes and finally ceases. Estrone, derived primarily from peripheral aromatization of androstenedione, becomes the predominant circulating estrogen. The postmenopausal ovary does continue to produce androstenedione and testosterone at premenopausal levels (8). The menopausal transition has been more specifically redefined from the early changes in menstrual cycle length and shortening of the cycle to the full postmenopause(9). However no test at the present time will make the diagnosis and the symptom complex remains the best clinical tool.

A variety of symptoms may accompany the menopausal transition (Table 1). While age at menopause ranges from 49-52 years, cigarette smokers can undergo menopause 1-2 years earlier compared to nonsmokers (10).

THE MENOPAUSAL SYNDROME

Estrogen production during natural menopause does not stop abruptly. For five to seven years before the onset of the last menstrual period, ovarian function begins to diminish. Menopause can also be induced

TABLE 1: selected menopause symptoms
Abnormal uterine bleedingVasomotor symptoms*

Vulvovaginal dryness, irritation, atrophy*

Urinary incontinence

Trouble sleeping*

Sexual dysfunction

Dyspareunia

Depression

Anxiety

Labile mood

Fatigue

Headache

Myalgias

Arthralgias

Weight gain

Poor memory

Dry skin

Dry eyes

Thinning scalp hair

Hirsutism

surgically (i.e. bilateral oophorectomy) or medically (e.g. chemotherapy or pelvic irradiation). Because ovarian estrogen levels fall abruptly with induced menopause, these women generally experience more severe menopausal symptoms (11, 12). Menopause occurring at or before age 40 is called premature menopause; up to 90% of cases of spontaneous premature menopause (primary ovarian insufficiency) are idiopathic.

After the last menstrual period, the ovary ceases to secrete estradiol. However, smaller amounts of a weaker estrogen, estrone, are still synthesized from androstenedione in the cortex of the adrenal gland and in the interstitial ovarian cells (in minor amounts)(8). Small amounts of this estrone can be transformed into estradiol.

Body mass is directly correlated with the rate of peripheral production of estrone and estradiol in postmenopausal women. Estrogen synthesis takes place largely in adipose tissue. Therefore, an obese woman may produce twice as much estrone and estradiol as a thin woman. This may help explain the increased prevalence of hypo-estrogenemic symptoms and the higher risk of osteoporosis observed in thin women.

PIVOTAL STUDIES

The results of the large Women’s Health Initiative (WHI) study have been both influential and controversial. In 2002, the estrogen-progestin arm of the WHI was stopped prematurely because of increases in the risk of breast cancer and coronary heart disease (13). In 2004, the estrogen-only arm was also prematurely discontinued, reporting that estrogen therapy had no effect on CHD risk and increased the risk of stroke and deep vein thrombosis in this population(14). Post-hoc analyses suggest no increase in CHD in women starting treatment within 10 years of menopause(15). Other doses and types of estrogens and progestins were not studied in WHI; smaller studies are now underway to further investigate whether age at therapy initiation and different types/doses of estrogens and progestins will result in different health outcomes. Table 2 provides a brief outline of pivotal studies (including ongoing ones) involving perimenopausal women.

The findings of the Women’s Health Initiative study, a prospective, randomized trial of more than 16,000 healthy, post-menopausal women, published in July 2002, have thrown the use of HT into question in both the medical and lay communities. The estrogen plus progestin arm of the study was halted because there was a small, increased risk of invasive breast cancer among women receiving the combined therapy, as well as an increased risk of heart attacks, stroke and clotting. These risks were not offset by the benefits: a decrease in colon cancer and hip fractures (13).

However, the average age of the women in the WHI was 63.2 years, and does not reflect normal clinical practice where replacement is used mainly for symptoms, including hot flashes, in women 10 to 30 years younger. Women in the WHI also had an average BMI of 28, one-third had hypertension, and one-half had a history of smoking (66). Thus, at the present time, the relative risk to benefit of using HT in younger, healthier women is largely unknown, and physicians cannot make all clinical decisions based on the WHI study, as it appears to apply specifically to the population studied. Hormone replacement is still an important therapeutic modality for women with symptoms and quality of life issues which deserves further study, and should be considered by physicians for their patients on an individual basis. At the present time there is global consensus that women with early or premature menopause should be treated until the normal age of menopause (age 50) and their treatment during these years should not be considered in the calculation of years of postmenopausal therapy(16).

TABLE 2: Important studies involving perimenopausal women

Study

Type

Location

Dates

N

Ages; mean

Hormone formulation

WHI (E+P)

RCT

US

1993-2002

16608

50-79; 63

CE 0.625 mgMPA 2.5 mg
Intact uterus. Stopped early after 5.2 yrs (planned for 8); increased CHD events & invasive breast cancer.
WHI (E)

RCT

US

1993-2004

10739

50-79; 63

CE 0.625 mg
Status-post hysterectomy. Stopped early after 6.8 yrs; increased risk CVA; lack of CHD benefits.
PEPI

RCT

US

1989-1994

875

45-64; 67

CE 0.625 mg± MPA 10 mg (days 1-12)± MPA 2.5 mg QD

± P4 200 mg (days 1-12)

Healthy women; 3 years follow-up. HT improved lipoprotein profiles. Unopposed estrogen associated with high rate endometrial hyperplasia.
HERS

RCT

US

1993-1998

2763

67

CE 0.625 mgMPA 2.5 mg
Subjects had known CHD. HT 36 months.
NHS

obs

34-59

SWAN

obs

US

1996-current

3302

42-52

As per patient preference (including no HT)
Multiracial, multiethnic (Caucasian, African American, Hispanic, Chinese, Japanese); includes premenopausal; yearly visits—currently tracking 12th-13th visits. Following bone density, cardiovascular health, mood, symptoms.
MWS

obs

UK

1996-current

1084110

50-64; 56

As per patient preference
WISDOM

RCT

UK, Australia, New Zealand

1999-2002

5692

50-69; 63

CE 0.625 mg± MPA 2.5 or 5 mg
Stopped early after median 12 months follow-up (planned 10 yrs) because of WHI results.
ELITE

RCT

US

2004-2013

643

6 yrs vs. 10 yrs postmeno.

E2 1 mg PO
2.5 yrs planned; endpoint atherosclerosis by carotid ultrasound.
KEEPS

RCT

US

2005-2012

727

42-58,52

CE 0.45 mgor E2 50 mcg transdermal P4 200 mg (days 1-12)
Harvard Mood

obs

US

1995-2006

460

36-45

No RX
(DOPS)Schierback et al

RCT

Denmark

1990-2008

1006

49.7±2.8

2 mg synthetic 17-β-estradiol for 12 days, 2 mg 17-β-estradiol plus 1 mg norethisterone acetate for 10 days, and 1 mg 17-β-estradiol for six days or 2mg 17β estradiol for hysterectomized

 


Abbreviations:

WHI, Women’s Health Initiative (13, 14).

PEPI, Postmenopausal Estrogen/Progestin Interventions (17).

HERS, Heart and Estrogen/progestin Replacement (18).

NHS, Nurses Health Study(19-22).

SWAN, Study of Women’s Health Across the Nation (23-25).

MWS, Million Women Study (26).

WISDOM, Women’s International Study of long Duration Estrogen after Menopause (27).

ELITE, Early vs. Late Intervention Trial with Estradiol. (28)

KEEPS, Kronos Early Estrogen Prevention Study.

Harvard Study of Moods and Cycles(29)

DOPS Danish Osteoporosis Prevention Study(30)

RCT, randomized controlled trial.

obs, observational study.

CE, conjugated estrogens.

MPA, medroxyprogesterone acetate.

E2, estradiol.

P4, micronized progesterone.

The results of the large Women’s Health Initiative (WHI) study have been both influential and controversial. In 2002, the estrogen-progestin arm of the WHI was stopped prematurely because of increases in the risk of breast cancer and coronary heart disease (13). In 2004, the estrogen-only arm was also prematurely discontinued, reporting that estrogen therapy had no effect on CHD risk and increased the risk of stroke and deep vein thrombosis in this population(14). Post-hoc analyses suggest no increase in CHD in women starting treatment within 10 years of menopause(15). Other doses and types of estrogens and progestins were not studied in WHI; smaller studies are now underway to further investigate whether age at therapy initiation and different types/doses of estrogens and progestins will result in different health outcomes. Table 2 provides a brief outline of pivotal studies (including ongoing ones) involving perimenopausal women.

The findings of the Women’s Health Initiative study, a prospective, randomized trial of more than 16,000 healthy, post-menopausal women, published in July 2002, have thrown the use of HT into question in both the medical and lay communities. The estrogen plus progestin arm of the study was halted because there was a small, increased risk of invasive breast cancer among women receiving the combined therapy, as well as an increased risk of heart attacks, stroke and clotting. These risks were not offset by the benefits: a decrease in colon cancer and hip fractures (13).

However, the average age of the women in the WHI was 63.2 years, and does not reflect normal clinical practice where replacement is used mainly for symptoms, including hot flashes, in women 10 to 30 years younger. Women in the WHI also had an average BMI of 28, one-third had hypertension, and one-half had a history of smoking (66). Thus, at the present time, the relative risk to benefit of using HT in younger, healthier women is largely unknown, and physicians cannot make all clinical decisions based on the WHI study, as it appears to apply specifically to the population studied. Hormone replacement is still an important therapeutic modality for women with symptoms and quality of life issues which deserves further study, and should be considered by physicians for their patients on an individual basis. At the present time there is global consensus that women with early or premature menopause should be treated until the normal age of menopause (age 50) and their treatment during these years should not be considered in the calculation of years of postmenopausal therapy(16).

This article will review the current state of knowledge concerning menopause and menopausal hormone therapy (HT). Unless otherwise noted the term HT will be used in this chapter to refer to use of estrogen and a progestogen in women with a uterus and to use of estrogen alone in women who do not have a uterus. The following questions will be addressed: What is the effect of HT on hot flashes, genitourinary tract atrophy, and other symptoms of the menopausal syndrome? Does HT reduce a woman's risk of osteoporosis, cardiovascular disease, or cancer of the breast, endometrium, or colon? Can HT slow the decline of cognitive function and prevent Alzheimer's disease? What recommendations should women be given in light of the Women’s Health Initiative findings?

VASOMOTOR SYMPTOMS

A hot flash is a transient feeling of warmth especially over the face and neck, which lasts for several minutes. For some women, hot flashes are associated with drenching sweats, increased heart rate, and post-flash chills. When they occur at night, they can cause sleep disturbances, fatigue, and depression. Vasomotor symptoms typically begin during the menopausal transition, reach maximal frequency and intensity during the two years after menopause, and gradually subside—ultimately lasting 1-5 years. Perhaps 75% of perimenopausal women experience hot flashes (5.) Up to 10% of women experience hot flashes for 10 years or longer. Recent data document vasomotor symptoms for a mean duration of 7.4 years with women starting in the pre or perimenopause having a longer duration ( 9.4 to 11.8 years) than women starting in the postmenopause (3-4 years). Prevalence of hot flashes differs by culture and ethnicity and can range from 0-80%. African Americans appear to have the longest duration (10.1 years) followed by Hispanics (8.9 years) followed by non-hispanic whites (6.5 years) Chinese (5.4years) and Japanese the shortest, (4.8 Years) (11, 25, 31-33). Hot flashes occur with greater frequency in women who undergo surgical menopause than in those who experience natural menopause. They are also more common at night, which often results in sleep disturbances, fatigue and depression. Additional factors including warm environments, consumption of alcohol or caffeine, and stress can exacerbate hot flash occurrence.

Several studies have documented the effectiveness of estrogen therapy (in various formulations, doses, and delivery methods) for hot flashes (11, 34). For example, the Women’s Health, Osteoporosis, Progestin, Estrogen (HOPE) study randomized postmenopausal women to 3 doses of oral conjugated equine estrogens (CEE) 0.625 mg/d, 0.45 mg/d, 0.3 mg/d; with or without medroxyprogesterone acetate (MPA) 2.5 mg/d or 1.5 mg/d. Among the 241 subjects, there was a was a significant reduction in the number and severity of hot flashes as compared to baseline and placebo after three weeks of hormone therapy(35). Average daily hot flashes fell from nine per day at baseline to two per day after one year of therapy. This reduction was seen in all doses of CEE and CEE/MPA, including the lowest doses (35). In one comparative 12-week trial of 204 postmenopausal women (average age, 52 years), oral CEE 0.625 mg and transdermal estradiol 50 mg/day provided similar symptom relief (36)

SLEEP DISTURBANCES

HT for relief of hot flashes is effective in up to 90% of menopausal women. However, many women still experience sleep disturbances. Poor sleep (including difficulty falling asleep, disrupted sleep, insufficient quantity, and poor quality) affects approximately 45% of perimenopausal women in the U.S. and is associated with reduced productivity, irritability, depression, and cardiovascular disease (37, 38). The causes are multiple and include: hot flashes, nocturia, anxiety, depression, and primary sleep disorders (i.e. apnea, periodic limb movements/restless legs syndrome).

Time of night (i.e. first vs. second half) appears to influence the etiology of poor sleep. Laboratory sleep studies have shown that hot flashes tend to cause arousals in the first half of the night and are associated with subjective poor sleep. However, apneas and periodic limb movements tend to cause arousals in the second half of the night—a time when rapid eye movement (REM) sleep predominates and suppresses thermoregulatory effector responses like hot flashes. Thus in the second half of the night, primary sleep disorders cause arousals (hence loss of REM sleep) and may subsequently precipitate hot flashes (39, 40).

Determinants of subjective versus objective sleep quality may also be different. Whereas subjective sleep quality tends to be lower in women who experience hot flashes and report anxiety, objective sleep efficiency (ratio of time-asleep to time-spent-in-bed) tends to be lower in women who have apnea or periodic limb movements (40).

Although it is often assumed that complaints of poor sleep are due to hot flashes, treatment of hot flashes may not improve sleep quality if there is an underlying primary sleep disorder or psychiatric condition syndrome. Thus, providers should assess patients for apnea, restless legs, anxiety, and depression, and consider appropriate treatment.

A double-blind, crossover study in hypogonadal postmenopausal women compared the effects of CEE (0.625 mg/d) and placebo on sleep patterns(41). Results showed no relation between hot flashes or night sweats and sleep disturbances. For the women on estrogen, however, sleep quality improved, along with length of REM sleep and sleep latency. The causes of sleep disturbances in the postmenopause are complex and further research is needed.

DEPRESSION

Although prior epidemiologic studies have concluded that postmenopausal women are not at increased risk for depression, studies in the past few years including the Harvard study of Moods and Cycles have shown depressive symptoms are observed more frequently in this during the menopausal transition.(42). Increased psychological distress was seen in the SWAN study of women in the transition (29, 43, 44). This vulnerability occurs even if women do not have a previous history of depression (44, 45), although women with a previous history appear to be at greater risk (46), and depression is more apt to occur in the later stages of perimenopause (47-50). In particular, a subgroup of women who show abnormal mood responses to periods of estrogen withdrawal such as postpartum and premenstrually with a diagnosis of PMDD (premenstrual dysphoric disorder) may be especially vulnerable (51, 52). Women suffering from hot flashes (53) and sleep disorder(54) are also vulnerable, but depressive episodes can occur regardless of the presence of hot flashes(55).Women with early onset of perimenopause also have a significantly increased risk of first onset depression(42, 56) as well as those with a longer length of perimenopause(47, 48). Thus hormonal fluctuations may be a psychological destabilizer and there is some evidence that sex hormones may prevent, attenuate or even treat depressive episodes in the perimenopause (50, 57). This is an area of evolving research, but two independent studies demonstrate successful treatment of depression with transdermal estradiol (49, 58). This treatment; however, does not appear to be successful following menopause (59). Whether asymptomatic postmenopausal women benefit is unproven but the KEEPS study showed improvement in depression, anxiety and sexual function(60) Current recommendations include the use of hormone therapy as well as selective serotonin reuptake inhibitors when treating refractory depression (61). However, the progestin in these patients should be chosen carefully as depressive symptoms may occur with these medications, and some patients cannot tolerate any progestin, even progesterone(62).

With regard to mood, the data supports the theory that symptoms of depression may be alleviated with the use of ET/HT. In a double-blind placebo controlled trial of perimenopausal women, Soares et al found that depressive symptoms were significantly relieved in women receiving estradiol compared to placebo (58). In addition, estrogen has been shown to improve mood in post menopausal women without clinical depression (63, 64). Further research is still needed in this area, but the preliminary data suggest that estrogen may be a possible treatment method for some depression symptoms.

GENITOURINARY TRACT ATROPHY

Large numbers of estrogen receptors are found in the vagina, vulva, urethra, and trigone of the bladder. Thus, atrophy of the genitourinary tract can occur as estrogen levels diminish.

Vulvovaginal atrophy causes significant complaints and is common in the menopause. Symptoms include dryness, dyspareunia, discharge, itching and occasionally bleeding. The symptoms increase with age and may lead to vulvovaginal fissures and stenosis. These symptoms are described as moderate to severe in the majority of women who report them which in one survey was 30%(65).

After menopause, the vaginal walls thin and lose their elasticity. They also produce fewer secretions and lose much of their lubricating ability in response to sexual stimuli. The vulva becomes flattened and thin as a result of the loss of collagen, adipose tissue and the ability to retain water(66). The urethra also becomes thinner and less efficient, with detrusor pressure at the urethral opening decreasing, both during and after voiding. Estrogen deficiency also leads to an increase in fibrosis of the bladder neck, reduced collagen in surrounding tissues, and a decrease in the number and diameter of the muscle fibers in the pelvic floor. There is a decrease in the superficial layer of the vaginal epithelium, a decrease in vaginal secretions and pH (normal is under 4.5) and an increase in vaginal infections due to loss of the normal acidic environment and overgrowth of opportunistic fecal bacteria at the expense of normal lactobacilli. Loss of subcutaneous fat leads shrinkage of the labia and retraction of the urethra(67) Estrogen treatment, both systemic and local can greatly relieve these problems(67-69).

These changes increase a woman's risk of vaginal and urinary tract infection. Atrophic genitourinary tissues are also at increased risk of injury by trauma. Estrogen replacement therapy can significantly lessen these problems. The advantage of using local vaginal therapy is that minimal if any absorption occurs after the first two weeks of therapy, and it can be used without the side effects of systemic therapy. Multiple studies have shown that the absorption of vaginal estrogen therapy is strictly dose dependent and is maximal in the first two weeks of treatment when the vagina is thin and atrophic. With the return of the normal superficial layer of the vagina, serum levels of estradiol remain in the postmenopausal range when used in minimal doses. Studies have followed women for up to 3 months including patients with breast cancer (69-71). Endometrial safety is maintained if the local therapy is minimal (0.5 grams of cream or 10-25ug of the vaginal pill twice a week or the equivalent) but evaluation is warranted for any bleeding(72).

SEXUAL DYSFUNCTION

All of the changes to the genitourinary tract can result in dyspareunia, leading to a decreased interest in sexual intercourse. Fatigue and depression brought on by the vasomotor symptoms and sleep disturbances of menopause can exacerbate this lack of interest in coitus.

Decreased levels of endogenous testosterone, both in women who have undergone surgical menopause, as well as in those who experience natural menopause, may cause decreased libido(73). Women who complain of lack of sex drive may be candidates for androgen replacement, as well as estrogen. In general, androgen levels do not decrease abruptly at menopause but decrease gradually as women age so that decreased libido may be a problem of older postmenopausal women.

OSTEOPOROSIS

The loss of ovarian hormone production after menopause puts women at increased risk for osteoporosis. Without estrogen, osteoclast activity and bone resorption are increased, and bone mass decreases. This reduced skeletal mass and microarchitectural deterioration increase the risk of fracture. At age 50, a Caucasian woman has a 16% lifetime risk of hip fracture, a 15% risk of a Colles’ fracture, and a 32% chance of an atraumatic vertebral fracture (74).

Peak bone mass—typically attained by the third decade of life—is determined by genetic and environmental (nutrition, lifestyle, physical activity) factors (75). There is often only slight bone loss between age 30 and the perimenopausal transition (76). The period of accelerated bone loss appears to last approximately 5 years starting 2 years before the final menstrual period lasting until 2-4 years following the final menstrual period. For example, a prospective study of 75 Caucasian women followed for 9.5 years found that subjects lost 10.5% of bone at the lumbar spine over the critical 5-year period while estrogen levels were declining (77).

Although estrogen therapy can prevent bone loss and reduce the risk of fracture in perimenopausal women, it is no longer recommended as first-line therapy for osteoporosis because of the risks associated with hormone therapy and because alternative therapies exist. Studies have shown estrogen to not only prevent bone loss (by decreasing osteoclastic activity), but also to reduce fracture rates by as much as 65%(12). Women in the Women’s Health Initiative receiving estrogen plus progestin suffered 5 fewer hip fractures per 10,000 compared to women on placebo (66). Estrogen therapy may be considered in women for whom the alternative agents (e.g. bisphosphonates, raloxifene, teriparatide) are intolerable or contraindicated; in whom estrogen therapy is also indicated for other reasons (e.g., vasomotor symptoms); or in whom the benefits of estrogen therapy outweigh the risks (78). In the WHI, women who received estrogen plus progestin had hazard ratios of 0.66 for hip fractures, 0.66 for vertebral fractures, and 0.77 for fragility fractures at any site, as compared to women on placebo (13). Prophylactic benefit increases when estrogen replacement is begun as soon after menopause as possible. Because bone loss continues as soon as estrogen replacement is stopped, treatment will be needed so as to maintain the positive effects on bone metabolism. However, the longer a woman has been taking estrogen, the more bone she will have when treatment is stopped and bone loss resumes.

While estrogen prevents bone loss in most postmenopausal women, some continue to lose bone mass despite the therapy, presumably because of genetic or environmental factors. Bone density studies should be conducted during the perimenopausal period and then repeated as needed to assess the status of bone loss(79).

All postmenopausal women should have an appropriate calcium intake (up to 1200 mg) and vitamin D (400 IU) supplementation. Concerns about calcium supplements and cardiovascular disease from both observational and randomized studies have changed recommendations, although the issue is controversial(80). Calcium is best obtained from food and women should aim to meet requirements primarily through nutrition and take supplements only if needed to reach RDA .Mean dietary intake of midlife and older women is 700mg/day (81) so supplement in the range of 500mg is appropriate when dietary intake of calcium is low. Recent studies have shown that more than 50% of women over age 50 are vitamin D insufficient and these replacement doses are probably inadequate(82). 1000-2000 IU of Vit D3 are probably a better estimate of replacement. Only 10% of calcium is absorbed when Vitamin D is low. Supplements can help compensate for poor dietary intake of calcium and inefficient vitamin D synthesis. Because calcium carbonate requires acid for absorption, women taking acid-suppressing drugs or with atrophic gastritis should take calcium citrate, which does not require gastric acid for absorption.

CARDIOVASCULAR DISEASE

Cardiovascular disease accounted for 30.7% of deaths in American women in 1999. It surpasses cancer, cerebrovascular disease, lung disorders, infectious disease, diabetes, suicide, and renal disease as the leading cause of death in women today(83). A woman has about 10 times the lifetime risk of dying of ischemic heart disease than of breast cancer, reproductive cancer, or osteoporotic fracture.

An acceleration of heart disease occurs after age 50, and approximately one third of the women who die of cardiovascular disease every year are under 65 years old (more than 100,000). This suggests that menopause (whether surgical, premature, or natural) may be a risk factor for heart disease(84). Because premenopausal women have lower incidences of cardiovascular disease than men and lose this advantage after menopause, it is logical to conclude that estrogen has a cardioprotective effect. It is thought that estrogen deficiency is at least partially responsible for the increased risk of developing heart disease after menopause.

Considerable controversy and confusion has recently erupted over the role of estrogen replacement therapy in preventing cardiovascular disease. A number of trials reported an increased risk of ischemic events when hormone therapy was started in older women with a history of heart disease(21, 85, 86). In response, the American Heart Association recommended that hormone replacement therapy not be used for primary prevention of cardiovascular disease. The results of the estrogen-progestin arm of the WHI showed similar results. Women on estrogen-progestin therapy suffered 7 more CHD events per 10,000 women than women on placebo. They also suffered 8 more strokes per 10,000 women than those taking placebo(13).

Emerging evidence suggests hormone therapy is most effective in protecting women whose hearts are not yet compromised from future cardiovascular disease as seen in a recent study by Hodis et al.(87). Researchers randomized 222 postmenopausal women with no history of cardiovascular disease, stroke, or cancer who had high levels of LDL (≥ 130 mg/dL) to receive either 1 mg unopposed 17-ß estradiol or placebo. After two years, women on estrogen had significantly less thickening of the inner carotid artery wall. Recent data published from the WHI study show that the risk of coronary heart disease is largely dependent on age of the women initiating therapy and the number of years since menopause. The lower risk in the 50 to 59 year age group and in those experiencing menopause within the last 10 years (15, 88) and those on therapy more than 6 years (15). Data on estrogen treatment alone in WHI showed a decrease in coronary calcium, particularly in younger women although the effect was observed in all ages (89). In contrast to these findings, other publications from the same study suggested that the gap between menopause and initiation of therapy has no effect on cardiovascular disease, contradicting their previous report which showed some protection if started early (90, 91). However these observations are from a combination of the randomized and observational studies with most women who were recently menopausal were previously taking hormone therapy. One study showed some protection after 6 years of use(92). Overall, most studies have shown convergence between the observational and the randomized control publications suggesting that younger women starting hormone therapy at menopause are not at increased risk for heart attacks (93). The KEEPS study examined the effects of hormone treatment on surrogate markers of cardiovascular disease in recently menopausal women including carotid intima-media thickness (IMT) and coronary calcium. Carotid IMT increased in a similar fashion in both treated and placebo groups and there was a non-significant trend for less coronary calcium in the hormone arms(94) The DOPS study followed women on hormone therapy for 16 years and although osteoporosis was the primary endpoint, mortality and hospitalizations for both congestive heart failure and MI was reduced in the treated arms. Younger women appeared to show more benefit(30) Probably most convincing are the results of the Elite trial showing that younger recently women treated with hormone therapy showed an attenuation of IMT thickness while women treated who were 10 years past menopause showed no such benefit(95). When women stopped therapy in WHI, the increased risk seen in the treated arm was no longer apparent after a mean of 2.4 years (96). Endothelial dysfunction, not atherosclerosis, appears to be significantly increased in women with hot flashes, perhaps explaining their increased cardiovascular risk profile (97). Since symptomatic women were not studied in WHI, the role of HT in relief of symptoms and in turn of their effect on coronary risk is unclear. An

However, at the present time, HT should not be recommended for the prevention of heart disease.

ALTERED LIPOPROTEIN PROFILES

The increased risk of cardiovascular disease after menopause might be explained by the atherogenic changes in plasma lipoprotein levels associated with estrogen deficiency. At menopause, plasma levels of low-density lipoprotein (LDL) increase by 10% to 15%. According to data from the SWAN study (Study of Women’s Health Across the Nation), women experience a very specific increase in lipids at menopause. This includes total cholesterol, low-density lipoprotein cholesterol, and apoliprotein B. These changes were similar across all ethnic groups (98).

This increase can be prevented with estrogen replacement therapy. Plasma levels of high-density lipoprotein (HDL) increase by 10% to 15% with estrogen therapy and may be an important factor in the cardioprotective effect of estrogen.

The use of progestins, however, in conjunction with estrogen seems to attenuate these beneficial effects on plasma lipoprotein levels to some extent. Data from the Nurses' Health Study showed that women who took estrogen and progestin in combination had the same apparent protection from coronary events as did the women who took estrogen alone(21).

However, as noted previously, the randomized HERS trial showed that HT (with 0.625 mg/d of conjugated equine estrogen and 2.5 mg/d of medroxyprogesterone acetate) increased the risk of coronary events in women with a mean age of 65 who had established cardiovascular disease(18). This effect was noted during the first year of HT use. Following the second year, a progressive protective trend was found with HT, although there was no overall beneficial effect in the study as a whole. Another study examined the effect of HT/ERT as well as ERT in women with angiographically verified coronary disease(99). Again, no benefit was seen. However, these women had proven heart disease and were, on average, 65 years of age. This is considerably older than the age when HT is usually started. These data suggest that HT raised the possibility that started prior to the development of cardiovascular disease might be protective.

Progestins may have variable effects on lipoproteins based on their androgenicity. More androgenic progestins tend to lower HDL levels to a greater degree than do the less androgenic progestins (100). The two types of progestins most commonly used for hormone replacement therapy are those derived from 19-norestosterone and 17-hydroxyprogesterone. The former are the more androgenic, while the latter have a little androgenicity. Medroxyprogesterone is the most commonly prescribed progestin in the United States and is derived from 17-hydroxyprogesterone.

More recently, micronized progesterone has become available. The Postmenopausal Estrogen/Progestin Interventions Trial (PEPI) showed that micronized progesterone, used with conjugated equine estrogen, had less attenuation of the favorable lipid profile induced by estrogen than medroxyprogesterone acetate (101).

VASODILATION

As important as estrogen's effects on lipid metabolism may be its vasodilatory properties. It appears to potentiate the effects of endothelium-derived relaxing factor (EDRF) in the coronary arteries. It also may affect vasodilation through an endothelium-independent pathway in the peripheral vasculature.

One study looking at postmenopausal women with angina and normal coronary arteries (syndrome X) saw diminished vasodilation before initiation of estrogen therapy and normalized hyperemic response after two months of treatment. Vasodilation was measured by testing hyperemic response to forearm blood flow occlusion. Chest pain either improved markedly, or resolved, in 19 of the 20 subjects. This improvement in angina symptoms suggests that the impaired vasodilatory response to an EDRF/nitric oxide stimulus may be systemic(102).

An additional study reported a beneficial effect for sublingual estradiol in reducing symptoms of exercise-induced myocardial ischemia in postmenopausal women with coronary artery disease(103). These results suggest both a reduction in peripheral vascular resistance and a direct vasodilatory effect in the coronary arteries.

OTHER EFFECTS

Additional studies have found an association between HT and a marked reduction in the pulsatility index of the internal carotid and middle cerebral arteries(104). This finding may help explain the reduction in stroke risk and the improvement in cognitive function seen with estrogen plus progesterone. According to recent data from the Nurses' Health Study, this effect is seen at low doses only (0.3 mg conjugated equine estrogen).

A recent study also looked at the effects of estrogen in women who had recently suffered ischemic stroke or transient ischemic attacks and found no reduced mortality or recurrence prevention with 1.0 mg estradiol compared to placebo(105). These findings discourage the use of HT for secondary stroke prevention.

Other factors associated with estrogen use which could lower the risk for cardiovascular disease include decreases in levels of the proatherosclerotic factor, lipoprotein (a), the procoagulant factor, fibrinogen, and increases in levels of factor 11 (prothrombin). One study showed that with discontinuation of hormone therapy there was a rise in use of antihypertensive medication(106).

BREAST CANCER

One in 8 women will be diagnosed with breast cancer in her lifetime, and risk increases with age(107). In 2001, approximately 40,200 women died of breast cancer, although survival rates have been increasing. The five-year survival rate for women with localized breast cancer has risen from 72% in the 1940s to 97% today. This high survival rate, however, decreases to 77% if the cancer has spread regionally, and to 21% if it has spread distantly(107).

Estrogen, a trophic growth hormone, may promote the growth of preexisting breast cancer. It is still unknown whether it may also induce the growth of new cancers. Use of estrogen alone for at least five years, may be associated with a slightly increased risk of breast cancer according to the Nurses' Health Study. However, a report from the Women’s Health Initiative study showed an small increase in breast cancer in women on estrogen plus progestin , women on estrogen only showed no increased incidences of breast cancer compared to women on placebo (13, 14). Recent publications showed a significant decrease in the incidence of breast cancer in this group(108), a surprising finding which may be related to the type of estrogen used in the WHI study (conjugated equine estrogen). The study is ongoing but clarification of this discrepancy has not been forthcoming. The relative risk of the Estrogen plus Progestin (E+P) arm of the study has varied from 1.24 to 1.28 and follow up publication from WHI showed a non significant risk of 1.20 (0.94-1.53)(109). It has been suggested that the effect of E+P is to promote the growth of occult tumors which are present on the initiation of therapy.(110)The risk is very small although the data interpretation has implied otherwise. The absolute number of excess cases is stated as 8/10000 per year and is related to cumulative exposure. Women who had never received hormones in the past in WHI did not have a significant risk over the 5.6 years of the trial and the risk was not significant in younger women.(111)There was no increase in risk for at least 7 years(109)

Many studies have not shown an increased risk of breast cancer with estrogen use. A large meta-analysis of 51 epidemiologic studies (involving more than 160,000 women from 21 countries) showed that HT increases the risk of breast cancer and that risk increases with longer use(112). That is, for every 1,000 women who began using HT at age 50 and continued using it for 5, 10, or 15 years, an additional 2, 6, or 12 cases of breast cancer would be expected to occur. However, another review showed that at doses of 0.625 mg/d conjugated estrogens, there was no increased risk of breast cancer.

Data from the Iowa Women's Health Study showed no increased risk of breast cancer in women who had used HT versus those who had not taken hormones(113). Additionally, when researchers went back and analyzed data from women who had developed breast cancer, they found that HT, in a very small number of women, was associated with cancer with a favorable prognosis(114). This finding is supported by other studies which have shown that women who use HT are less likely to have metastatic disease, and have a longer life expectancy than women who have not used HT(19). The findings of these studies suggest that rather than acting as a carcinogen, estrogen may act as a mitogen. However, one possible explanation for these findings is that women on HT are more likely to be seeing a doctor regularly and to undergo regular breast examinations and mammograms.

Data from the Nurses' Health Study showed a survival advantage for women taking estrogen at the time their breast cancer was diagnosed. The increased survival rate was associated with a lower frequency of late-stage disease and undoubtedly reflects earlier diagnosis in estrogen users(19). However, other evidence suggests that estrogen users develop better differentiated tumors and that surveillance or detection bias is not the only explanation for better survival(115, 116).

A number of recent studies have aroused concern over the effect of menopausal HT on breast tissue density. In women not on HT, breast density has been found to be an independent risk factor for breast cancer(117). Hormone therapy has been found to increase breast density, with the greatest increase in women on conjugated estrogen and progesterone(118).

Although an association between breast density and breast cancer has not been seen in women on HT, there has been some concern that mammograms may be less effective in women on HT with greater breast density. However, Rutter et al. showed that two weeks after discontinuing HT, women's breast density returned to normal(119). Therefore, until this issue is better understood, it may be advisable for women to discontinue HT for two weeks before a mammogram exam, especially in the case of prior problematic mammograms.

Evidence suggests, however, that estrogen plus progestin may have an impact on breast cancer. In July 2002, the estrogen plus progestin arm of the Women’s Health Initiative study was stopped due to a small increase in the incidence of breast cancer among women taking this combination. This risk amounted to approximately 8 more women per 10,000 being diagnosed with breast cancer compared to those on placebo(13). It is important to note, however, that the average age of women in this study was 63.2 years and does not reflect women on HT in normal clinical practice. In addition, 50% of the women in WHI were either current or former smokers, they had an average BMI of 28 (well-above normal), and 1/3 suffered from hypertension.

In the MWS (Million Women Study), the large British study, women on HT followed for 2.6 years were found to have an increased risk of breast cancer (RR 1.66) (26). Various hormone preparations were tested in this trial and similar risks were reported for all types, suggesting that risks are not confined to the standard CEE/MPA dose used in WHI. It is important to note though, that women taking estrogen only had a significantly lower increase in risk compared with women taking both an estrogen and progestogen. It is an important to recognize that this was an observational study only and hence has a larger potential area for error.

Although there is some evidence that combination therapy may increase risk of breast cancer above that of estrogen alone, neither a protective, nor a detrimental effect has been demonstrated convincingly, particularly for younger, healthier women. One study interviewed nearly 4000 women with and without breast cancer and found a significant correlation between use of continuous combined replacement therapy and breast cancer(120). However, the risks were higher in thin women than in heavier women which may confound the results. Also, it is possible that the use of cyclic therapy could provide the additional risk, and HT was generally given at higher doses that are rarely used today.

While there has been little consistency among the findings of the various studies on the effects of menopausal HT on breast cancer, one issue that is consistent in the literature is the observation that mortality from breast cancer is decreased among ET/HT users. A summary of the literature from 1990-2001 shows the RR of mortality consistently to be <1.0 with HT use [75-80]. One hypothesis to explain this observation is that HT may promote the development of slow-growing tumors or discourage the development of more aggressive tumors. Hulley et al, reported that tumors in women taking /HT were smaller, had a better histologic differentiation, an a lower cell-proliferation rate compared to nonusers(121). It has also been posited that better screening of these women leads to lower mortality rates.

The argument that menopausal HT should not be given to women who have a personal history of breast cancer may seem reasonable based on evidence that breast cancer is a hormone responsive tumor. However, while women with a first-degree relative (mother, sister, or daughter) who has or had premenopausal breast cancer are at increased risk by virtue of their family history alone, their risk of breast cancer is not thought to be increased further by HT use. Eighty percent of women who develop breast cancer do not have a family history. Sellers et al., examined HT use and breast cancer risk in women with a family history of breast cancer and found no statistically significant increase in risk in past or current users, regardless of duration of use (113). This is supported by the findings of Rebbeck et al., who studied women who were carriers of the BRCA1 gene mutation (122). Bilateral prophylactic oophorectomy was associated with a 47% reduction in breast cancer risk in this population. HT use did not negate the observed reduction in cancer risk. Interestingly, studies of breast cancer survivors showed that women using HT had a lower risk of recurrence compared to survivors not using HT (123, 124).

Breast cancer incidence is thought to increase after hormone use and since WHI there has been much interest on the role of the progestin in combination with estrogen in contrast to the use of estrogen alone(13, 14, 109). In general most studies that have shown a small increase have shown more of an effect with the combination (26) and nurse health and collaborative study). This has led to speculation as to the role of progestin, and to the minimization of progestin use despite the well-recognized and significant risk of endometrial cancer with the use of unopposed estrogen. Some recent studies suggest the progesterone and dydrogesterone may be safer than other progestins but no randomized studies examine this question(125). In general, some effect is seen with treatment duration and some studies show an effect although small. WHI reported an increase in breast cancer risk in the combined therapy arm in subjects who had used hormones prior to enrollment but only after 5 years (109). A later paper from the WHI study however suggested that the risk was higher in women who initiated therapy soon after menopause (within 3 to 5 years) (90). However, in this study, a much larger group of women who were recently menopausal had been on HT and the effect was more pronounced in the less rigorous observational arm. In general the effect takes several years to appear and is small. When hormones are discontinued the effect starts to decline within one year (96). All of this confusing and contradictory data suggests that the combined HT may be acting as a promoter in susceptible women with undiagnosed subclinical cancer and the promoter effect may disappear with discontinuation of therapy. This may also explain the overall drop in breast cancer seen with the Seer (Surveillance, Epidemiology and End Result) cancer registries database report. This report showed a drop in breast cancer rates after 2002 when women stopped hormone therapy after the WHI publications(126). This effect has not been seen universally and the trend was actually seen prior to the reports. In fact there has been a drop in many different cancer rates, possibly due to earlier detection and earlier treatment(127). Another item of interest is that the use of the less common lobular cancer of the breast (approximately 16 % vs. 70% for more common ductal cancers) is increased with hormone use (128). However this effect was not seen in WHI. Both combined hormone use and estrogen alone lead to denser breasts and more abnormal mammograms (111, 129). This effect is rapidly reversible and stopping hormones 10 to 30 days before a mammography may decrease abnormalities requiring follow up(130). One group of women who benefit from hormone therapy is the women with BRAC 1 and 2 mutations who undergo oophorectomy as prophylaxis. Use of HT does not appear to place them at risk for the genetically determined breast cancer and will improve quality of life(131). It will also prevent the effects of estrogen deprivation at a young age. The effects of stopping hormones are contradictory depending on the study. The Nurse’s Health Study reports that the risk is no longer present after 5 years while follow up in The WHI study shows a persistence of effect after 11 years of follow up(132).

Breast cancer prognosis does not appear to be influenced by the high hormone levels during pregnancy, nor has oral contraceptive use been shown to increase breast cancer risk. These observations may allay some of the fear regarding the use of exogenous hormones after menopause

OVARIAN CANCER

Data on ovarian cancer has not shown a consistent risk with use of hormone therapy. There is possible weak association with long term (at least 10 years) of therapy but data are inconclusive for recommendations (133). Its use does not adversely affect the risk of cancer in BRCA mutations (134). While WHI researchers reported an increased risk of ovarian cancer (HR 1.58), it did not reach statistical significance (135). Other studies too, including HERS and a meta-analysis of 15 case-controlled studies found no significant association(135-137).

ENDOMETRIAL CANCER

In 2001, 38,300 cases of endometrial cancer were diagnosed, and 6,600 women died of the disease. The mean age at diagnosis is 61 years, with most cases occurring in women 50 to 59 years old.

Estrogen alone causes endometrial hyperplasia and a two to three-fold increase in the risk of endometrial cancer. However, the addition of progestogen reduces this risk to lower levels than those seen in women not on HT(138, 139). Thus, the addition of a progestational agent to postmenopausal estrogen therapy is now standard for women with an intact uterus. While there have been some reports that the risk of endometrial cancer may be slightly increased even with the combined therapy, most studies have not confirmed this. Women in the WHI study on combined therapy showed no difference in endometrial cancer rates compared to women on placebo (13). Recent research has focused on the use of lower doses of estrogen and a progestogen in HT to reduce the risk of endometrial cancer(140).

The dose of progestogen given depends on several factors, including the number of days given each month, the amount of estrogen given, the individual needs of the patient, and her ability to tolerate the medication. Side effects of progestogen can include anxiety, irritability, depressed mood, acne, bloating, fluid retention, headaches, breast tenderness, and bleeding problems. The inability to tolerate these effects is the main reason for poor compliance or discontinuation of HT.

COLON CANCER

Despite being one of the major causes of cancer-related mortality in women, colon cancer is often overlooked by patients in their risk assessment of HT. Case-controlled and cohort studies have both found a 50% decrease in relative risk of colon cancer in women who are current or long-term HT users compared to women not on HT. In addition, reports from the WHI study showed that the combined estrogen plus progestin therapy was associated with a decrease in the incidence of colon cancer compared to women on placebo (6 fewer cases per 10,000 women on HT(13). This was not found with estrogen alone(14). Although the exact mechanism of estrogen and progestin’s protective effect on the colon is unclear, it has been suggested that estrogen acts to decrease bile acids, which are thought to be carcinogenic. At present; however, although the evidence that HT may be beneficial in reducing the risk of colon cancer should be considered, there is insufficient evidence to warrant recommending long-term HT solely for this purpose.

NEUROLOGIC FUNCTION

Cognition

The existence of estrogen receptors in the hippocampus, a part of the brain essential to learning and memory, has been known for some time. Several mechanisms may account for the effects of estrogen on the brain. Firstly, estrogen increases levels of choline O-acetyl-transferase, the enzyme needed to synthesize acetylcholine, a neurotransmitter thought to be critical for memory(141). Studies on healthy middle-aged and elderly postmenopausal women have supported the theory that estrogen may help to maintain aspects of cognitive function(142),(143). Data also suggest that estrogen therapy may enhance short- and long-term memory(144),(145). Additional effects of estrogen on neural function include: protecting neurons from oxidative stress and glutamate toxicity (146),(147), increasing glucose transport and cerebral blood flow , and stimulating the branching of neurites (148). A recent review of clinical trials of hormone therapy suggest that there is a clear difference between the effects of estrogen therapy and estrogen plus progestin (149). There is modest support for the beneficial effect of estrogen alone on verbal memory in women under 65, and possibly surgically menopausal, while a harmful effect is seen with estrogen plus progestin in women over 65. Conjugated estrogen with medroxyprogesterone acetate may also have some detrimental effect on younger women. Estrogen alone appears to be neutral in women over 65. Thus the age of initiation of therapy and the use of progestins are important when evaluating possible effects on verbal memory(149). At present there is no combination which appears to be neutral to verbal memory and there is suggestion of some harm even with micronized progesterone (150). Hot flashes appear to relate to memory dysfunction, and some of the cognitive improvement on hormone therapy may relate to the treatment of the hot flashes(151).

Alzheimer's Disease

For every five years after the age of 65, the prevalence of Alzheimer's disease doubles in the population. Nearly 50% of women over the age of 75 may suffer from the condition(152). As the population ages over the next 20 years, these numbers are expected to increase.

According to epidemiologic evidence, there is reason to believe that estrogen deficiency may contribute to Alzheimer's disease. Low body weight is associated with low levels of circulating estrogens in postmenopausal women. Women who suffer from Alzheimer's disease tend to have lower body weights than women without the disorder(153). Incidences of Alzheimer's disease are low or its expression is delayed in postmenopausal women with high levels of endogenous estrogenic steroids or those receiving long-term HT.

One explanation for estrogen's apparent protective effect may involve neurotransmission. Estrogen acts as a trophic factor for cholinergic neurons in vitro. Cholinergic depletion is the most prominent neurotransmitter deficit in Alzheimer's disease.

With regard to the association between risk of Alzheimer’s Disease and HT use, however, there is little consistency in the literature. However, while HT does show promise in preventing or delaying the onset of the disease, a recent study showed no benefit of either 0.625 mg/d or 1.25 mg/d of estrogen on Alzheimer's progression(154). Most likely, estrogen may merely delay the deterioration seen in Alzheimer's patients. Paganini-Hill and Henderson(155) reported a 35% decrease in risk for ET users compared to placebo, and Zandi et al.,(156) reported a 41% reduced risk for ever users of HT. However, the results from the WHIMS, the Women’s Initiative Memory Study, a substudy of WHI, reported that while HT did not significantly increase the risk of mild cognitive impairment (HR 1.07), it did increase the risk of probable dementia (HR 2.05) (157). The effect of HT on different subtypes of dementia could not be determined because the number of cases was too small. It must be noted, however, that because the WHIMS participants were all 65 or older, these results may not apply to women who initiate HT at a younger age.

The results of the Cache County Study(156) serve to further confuse the issue. In this prospective study of incident dementia in older women (mean age 74.5 years), the risk of AD was increased in current HT users with 10 or fewer years of therapy (HR 2.41 for fewer than 3 years of therapy, 2.12 for 3-10 years). For current users with more than 10 years of therapy the HR was 0.55, indicating a decrease in risk, but this value did not reach statistical significance. Interestingly, in past users, reductions were present in all age groups and showed a duration effect (HR 0.58 for fewer than 3 years, 0.32 for 3-10 years, and 0.17 for more than 10 years).

OTHER POSSIBLE RISKS

Thromboembolic disease

The Nurses' Health Study showed a twofold increase in the risk of pulmonary embolism among postmenopausal women who were current estrogen users. The recent findings of the WHI study confirmed these findings for women on combined estrogen plus progestin therapy. Women on this treatment suffered 8 more pulmonary emboli per 10,000 than women on placebo(13). Although estrogen use has been associated with an increase in the relative risk of venous thromboembolism (VTE), the absolute risk remains low, as VTE occurs infrequently in this setting. Women on combined estrogen-progestin therapy in the WHI study suffered 18 cases of more venous thromboembolism than women on placebo. However, when considered against a 50% reduction in cardiovascular disease risk, the increased risk of VTE does not contraindicate estrogen replacement. It does, however, show that patients should be screened for a history of idiopathic thrombosis as this has been a consistent finding (22).

Gallbladder disease

Some epidemiologic studies have found an increased risk of gallstones among women who use HT. Estrogen has been shown to increase cholesterol saturation of bile, alter bile acid composition, and decrease bile flow. Each of these effects can enhance gallstone formation. Data from the Nurses' Health Study (54,845 postmenopausal women monitored for eight years) showed that current HT users were more likely to have undergone cholecystectomy than nonusers (relative risk, 2.1). This risk tends to increase with long-term therapy and with high doses of estrogen(158).

Weight gain

Because many women gain weight as they age, a common fear is that HT will exacerbate this problem. However, this is unconfirmed by prospective studies. The PEPI trial showed that women on HT gained less weight than women not taking hormones(101). Attention to diet (with reduced fat intake) and regular aerobic exercise for weight maintenance should be recommended to all postmenopausal women. Data from WHI also showed an attenuation of increases in weight seen with age in the combined hormone treated arm(159). This suggests there may be some beneficial effect to HT on the normal increases that are seen in postmenopausal women and that the effect may protect against the increase in central obesity seen in hypoestrogenic menopausal women. A decrease in the incidence of diabetes, and lower insulin levels suggestive of better insulin sensitivity may be related to this attenuated weight gain.(160).

OTHER POSSIBLE BENEFITS

About 35% of patients over the age of 75 are affected by macular degeneration, the leading cause of severe vision loss in the elderly. One study showed that women who experienced menopause earlier in life had a 90% increased risk of developing symptoms of age-related macular degeneration later in life as compared to women who underwent menopause at an older age(161). Some studies have shown a small reduction in the incidence of this eye disorder among users of HT(162, 163).

Skin

It is thought that skin may be an important target organ for reproductive hormones. In postmenopausal women, dermal collagen decreases, and skin becomes thinner. Applying estrogen cream to the skin after menopause improves the external appearance of facial skin. In addition, systemic HT increases dermal collagen and limits age-related skin extensibility. To date, of the eleven clinical trials that examined the effect of HT on collagen levels, only one failed to demonstrate efficacy (164). Furthermore, results from a recent study indicates that estrogen also increases skin thickness (165).

HT has also been shown to accelerate cutaneous wound healing, both microscopically and macroscopically, in postmenopausal women (166). This study also showed delayed repair of acute incisional wounds in ovariectomized young female rodents; the delay was reversed by the topical application of estrogen.

Tooth loss

The risk of tooth loss increases after menopause. Osteoporosis, as well as estrogen deficiency, could both be contributing to this effect. Data from the Nurses' Health Study indicate that the risk of tooth loss may be decreased in women with a history of estrogen therapy (167).

TREATMENT

Non hormonal and combination treatments

Several treatments have recently become available and have FDA approval for relief of vasomotor symptoms. This includes a selective serotonin reuptake inhibitor, low dose paroxetine(168, 169) and basedoxifene/conjugated estrogens which also affords protection of bone. The latter consists of a combination of CEE and a SERM and is indicated for women with a uterus. A progestin is not necessary as this combination offers endometrial safety(170-172), Another SERM ospemifene has been approved for the treatment of postmenopausal vulvovaginal atrophy (173-175). Another treatment consists of a Swedish pollen extract femal, which has been shown to be effective in a small study for a composite of menopausal symptoms including vasomotor symptoms, fatigue and quality of life(176).

HORMONE THERAPY PRINCIPLES

Over the years, doses of estrogen in hormone therapy have been decreasing: until the mid-1970s, daily doses of CE as high as 1.25 or 2.5 mg were commonly used. Today, a CE dose of 0.625 mg/day is considered the “standard” dose for estrogen therapy while many women have relief of symptoms with even lower doses.

The goal of hormone therapy is to reduce menopausal symptoms (e.g., vasomotor symptoms, sleep disturbance, vulvovaginal symptoms, decreased libido) using the lowest effective dose for the shortest amount of time. Use of the lowest clinically effective dose of HT for relief of menopause-related symptoms and for prevention of osteoporosis is now recommended. Low-dose estrogen therapy (ET) is currently defined as a dose of oral CEE of ≤0.45 mg/d, oral estradiol ≤0.5mg/d, transdermal estradiol ≤0.0.375 mg/d, or the equivalent. The benefit-risk ratio of hormone therapy for each woman is influenced by the severity of her menopausal symptoms and their impact on quality of life, her current age, age at menopause, time since menopause, cause of menopause, and baseline disease risks. Some patients may require “standard” doses; however, and doses can be reduced if desired after 6 months to a year.

Generally appropriate indications include also treatment or prevention of osteoporosis in women who are not candidates for (or cannot tolerate) other osteoporosis therapies including bisphosphonates or teriparatide.

Absolute contraindications for systemic HT include hormone-related cancer, active liver disease, history of hormone-induced venous thromboembolism, history of pulmonary embolism not caused by trauma, vaginal bleeding of unknown etiology, and pregnancy. Relative contraindications include chronic liver disease, severe hypertriglyceridemia, endometriosis, history of endometrial cancer, history of breast cancer, coronary artery disease.

Guidelines for hormone use are reviewed in the statement of the North American Menopause Society(177) and recently by the Endocrine Society(178).

Considerable confusion has developed as a result of the numerous transdermal preparations which have appeared on the market. The effective dose depends on the delivery rate and the surface area applied so that there is much variation in terms of estradiol delivered to the blood stream. The following charts attempt to present equivalent doses. Lower doses take longer (4-7 weeks) for effective relief, and it is important to individualize therapy. Most preparations take a full 12 weeks for maximum effect although standard therapy provides relief sooner (2-3 weeks). There is also much debate as to the safety of oral vs. transdermal estrogen and the issue of dose vs delivery has not been resolved by double blind randomized trials. One study suggests that venous thromboembolism may be lower with transdermal products, but the doses compared were not equivalent(179). Another study shows a decreased risk of stroke in women on transdermal preparations with higher doses of both oral and transdermal estrogen showing significant effect (180). One study suggests progesterone may be associated with a lower risk of breast cancer than progestins but this again awaits further study (181).

Bioidentical Hormones.

The unfortunate publicity concerning compounded hormones mislabeled as ‘bioidentical” has suggested that custom made preparations based on saliva or blood levels were safer or better tolerated has lead to a cottage industry which has no scientific basis. These preparations offer no advantage over regulated and tested preparations approved by the FDA, and their risk is equivalent to commercial compounds. Claims that they are safer are misleading particularly since they have not been studied and one of the estrogens used, estriol, has no safety or efficacy data. Prescribers who claim they are more “natural “ do not inform patients that they are synthesized from plant chemicals extracted from yams or soy similar to some commercial preparations.

In general initiation of treatment of the symptomatic newly menopausal women will provide benefit which greatly outweighs risk and provides protection from bone loss. Older women who continue to be symptomatic may. continue treatment preferably with lower doses.

TREATMENT GUIDELINES

Although the decision to treat menopausal women rests on individualized risk vs. benefit for the patient some helpful clinical guidelines are useful for the clinician. In general, hormone treatment is being used for symptoms. These include vasomotor symptoms and vulvovaginal atrophy. There are, however, a variety of symptoms which make up the menopausal syndrome and are not strictly classified as vasomotor or vulvovaginal symptoms and are distressing to the patient and may also be a consideration for treatment(182). These include: mood disorders appearing at the perimenopause and menopause, migraines, severe insomnia, anxiety, difficulty concentrating, memory issues, severe fatigue and somatic symptoms especially joint pains and rarely muscle pain or a generalized crawling feeing on the skin. Some patients can endure two hot flashes a day while others who are in stressful or public jobs cannot. Patients are usually uncomfortable and distressed by more than two hot flashes per day. In particular the patient who wakes at night two or more times and suffers from sleep deprivation is usually in need of treatment. Patients suffering from five to seven hot flashes a day are experiencing moderate to severe symptoms and should be offered treatment. The physician should help the patient make a quality of life decision and advise these patients on the low risks associated with treatment particularly for a few years. Some patients may be experiencing bone loss and hormone therapy is ideal for this type of patient. Some of the estrogens on the market are also approved for prevention of osteoporosis and data shows they are very effective and prevent fractures. A patient on hormone therapy does not need a second drug for prevention of bone loss. If bone loss is occurring on hormone therapy a secondary cause should be searched for such as vitamin D deficiency, over treatment with thyroid hormone or hyperparathyroidism. Patients with mood issue may have problems with progestins and micronized progesterone or a vaginal delivery system may be better tolerated. Estrogens should be started first and a progestin added after a few weeks. Patients with migraines also have special tolerability issues and fluctuations of hormone levels which may be triggering the headaches may persist or be aggravated initially by hormone treatment. A transdermal patch may be the best option and a progestin should be started after a trial of treatment with estrogen. The issue of duration of hormone treatment will arise. Two to five years is usual. The small risk of breast cancer is also important to review with the patient. This risk surfaces after 5 years of use and did not surface at all with estrogen alone therapy after 7 seven years. This interval does not apply to patients with premature menopause who have been shown to be at risk for osteoporosis and premature heart disease if they are not replaced. All patients need a yearly mammogram and the increase in density can be avoided by stopping hormones for two weeks prior to the mammogram if she can tolerate it. Some patients stay on hormone therapy long term because of mood or other issues or they are in the unfortunate 10 percent who continue to suffer form vasomotor symptoms or cannot tolerate other drugs for osteoporosis. Patients with severe mood issues may require antidepressants. Recent data has shown the efficacy of low doses for vasomotor symptoms and many are available. However the patient with severe symptoms may prefer a standard dose which may be lowered after 6 months when symptoms are well controlled. Lastly, vaginal estrogens are an excellent option for patients with symptoms of vaginal atrophy and do not have the risks associated with systemic use. In particular, recurrent urinary tract infections and or vulvovaginitis are a hallmark of genitourinary estrogen deficiency which can be easily relieved or prevented with the use of vaginal estrogen. Treatment with hormone therapy is very individualized and quality of like may be greatly improved its use. When therapy is discontinued, a return of symptoms is common(183) although generally in a milder form. Unfortunately there is little data to guide the physician but many clinicians slowly taper doses over several months.

When assessing risk vs. benefit for long-term risks, the following conclusions from the WHI should be taken into consideration (13, 14, 88):

 

Over 1 year, per 10,000 women, estrogen/progestogen treated women had the following observed differences compared to controls:Estrogen and Progestin Estrogen Alone

  • 7 more CHD Events 5 fewer CHD events
  • 8 more strokes 12 more strokes
  • 8 more invasive breast cancers 7 fewer breast cancers
  • 18 more VTEs 7 more VTE
  • 8 more PEs 3 more PEs
  • 6 fewer colorectal ca 1 more colonrectal ca
  • 5 fewer hip fractures 6 fewer hip fractures
  • Long term follow up of subjects in WHI after discontinuation of treatment(132):
  • Intervention: CEE+MPA RR
  • Breast1.24, Stroke 1.37, Pulm embolism1.98
  • Colorectal CA 0.62, hip fracture 0.67, diabetes 0.81
  • Post Intervention :CEE + MPA.
  • Breast 1.28 all others attenuated
  • Intervention: CEE RR
  • Stroke 1.35, Hip fracture 0.67, diabetes 0.87
  • Post intervention: CEE RR
  • Breast CA 0.70, under 60 favorable mortality, less MI
  • Overall mortality not affected

 

 

 

1. Assess patient’s risk and symptoms.
Risks Symptoms
Osteoporosis
  • Amenorrhea or missed menstrual periods
  • Hot flashes or night sweats
Cardiovascular disease
  • Urogenital symptoms
  • Decreased sex drive, libido
Surgical menopause
  • Insomnia
  • Dyspareunia
Premature menopause
  • Osteoporotic-related height loss, disability, pain
  • Depression, mood change
Family history of Alzheimer’s disease
  • Headache
  • Irritability, emotional lability

 

 

 2. If risk or symptoms are present, screen for HT appropriateness.

CONTRAINDICATIONS

Absolute Relative
Hormone-related cancer or active liver disease Chronic liver disease
History of hormone-induced thromboembolism Severe hypertriglyceridemia
History of pulmonary embolism not caused by trauma Endometriosis
Vaginal bleeding of unknown etiology History of endometrial cancer
Pregnancy History of breast cancer
Proven coronary heart disease or recent event

 

 

3. If not appropriate, consider alternative therapies.

 

 

4. If appropriate for HT, consider the following:

ESTROGEN ALONE IF UTERUS IS ABSENT OR ESTROGEN-PROGESTOGEN IF UTERUS PRESENT

ESTROGEN/ANDROGEN (E/A) THERAPY

Risks present for: Symptoms present: Risks present for: Symptoms present:
Osteoporosis
  • Hot flashes or night sweats
Osteoporosis, not responsive to HT Symptoms as for ERT and/or:

  • Low energy
  • Decreased sex drive, libido
  • Muscle wasting
Cardiovascular disease
  • Urogenital symptoms
  • Osteoporotic-related height loss, disability, pain
Surgical menopause
  • Insomnia
Premature menopause
  • Headache
  • Irritability, emotional lability
  • Depression, mood change
  • Dyspareunia

 

 

5. After starting ET therapy, re-evaluate at 3 to 6 months.
  • If symptoms are controlled, continue HT.
  • If symptoms are not controlled or undesirable side effects are present:
Symptoms:
  • Headaches
  • Breast Pain
  • Urogenital symptoms
  • Irritability, emotional lability
  • Persistent hot flashes or night sweats
  • Decreased sex drive, libido
  • Fatigue
  • Insomnia
  • Depression, mood change
  • Irritability, emotional lability
  • Headache

 

Treatment:
  • Re-evaluate HT (dose and/or type)
  • Consider lower dose
  • Re-evaluate (3-6 months)
  • If symptoms controlled, continue treatment.
  • If symptoms not controlled or undesirable side effects persist, consider E/A or alternative therapy or consultation.
  • Issues of treatment duration will vary with individualized consideration of osteoporosis, dementia, or emerging issues. Breast cancer risk is present and is very small, less than 1/1000 women or less than 0.1%. Lower doses can be considered after symptoms are controlled. Time frame to consider lower doses should be individualized.

 

 6. Hormone Products for Treatment of Menopausal Symptoms
Estrogen Preparations
Doses (standard) Low dose Therapy
  • Conjugated (equine and synthetic) estrogens: 0.625, 0.9, 1.25 mg,
  • Micronized estradiol: 1, 2 mg
  • Ethinyl estradiol: 5 μg
  • Estradiol valerate: 2 mg
  • Piperzaine estrone sulfate: 0.625,1.25, 2.5 mg
  • Esterified estrogen: 0.625, 1.25, 2mg
  • estradiol acetate 0.9, 1.8 mg
  • Conjugated (equine and synthetic) estrogens: 0.3, 0.45 mg
  • Micronized estradiol: 0.5mg
  • Ethinyl estradiol: N/A
  • Estradiol valerate: N/A
  • Piperzaine estrone sulfate: N/A
  • Estradiol acetate 0.45 mg

 

Transdermal
  • E2 patch to deliver 0.05 mg/d
  • E2 patch to deliver 0.0375 ,0025,0.014 mg /day
  • E2 gel
  • to deliver 0.035, 0.025, ------------------
  • 0.027,0.0125,0.009,0.003/day
  • E spray
  • To deliver 0.021 mg/day

 

 

Vaginal Preparations
  • Conjugated Estrogen Cream
  • Estradiol Vaginal Cream
  • Estradiol Vaginal Tablets 25 or 10μg
  • Estrogen Vaginal Ring 2 mg
Treatment regimen:
  • Continuous
  • Cyclic

 

 

Progestin and Progesterone doses and types; The doses are for standard estrogen regimens. Doses can be halved with half doses and must be increased with higher estrogen doses. In general doses of estrogen therapy producing 35 to 60 pg/ml serum levels require standard doses of progestin but little literature is available.
CYCLIC
Medroxyprogesterone acetate: 5-10 mg for days 1-14 of each month
Norethindrone acetate: 2.5 mg for days 1-14 of each month
Micronized progesterone: 100 mg a.m. and 200 mg p.m. for days 1-14 of each month or 200mg in p.m., 2 hours after a meal
COMBINED (Use half with Half doses)
Medroxyprogesterone acetate: 2.5mg or 5mg daily
Micronized progesterone: 100 mg daily in the p.m., 2 hours after a meal
Other
Vaginal progesterone 4% : 6 doses every other day monthly. 45 mg per applicator (not FDA approved for menopausal use)
Levonorgestrel containing IUD 20ug/day release -5 year use (not FDA approved for menopausal use)

 

Combination EPT products
Cyclic
Regimen Route Available dose combinations
Conjugated Estrogen+medroxyprogesterone acetate :E alone for 14 days then E+P days 15-28 Oral, once a day 0.625 E+2.5 mg P0.625 E+5.0 mg P(2 tablets: E and E+P)
COMBINED
Regimen Route Available dose combinations
Conjugated equine estrogens (CEE) + medroxyprogesterone (MPA) Oral, once per day 0.625 mg CEE + 2.5 mg MPA, 0.625 mg CEE + 5.0 mg MPA, 0.45 or 0.3mg CEE +1.5mg MPA
Ethinyl estradiol (EE) + norethindrone acetate (NA) Oral, once per day 5 g EE + 1 mg NA, 2.5 ug EE + 1mg NA
17B-estradiol+norethindrone acetate 1MG e+0.5 MG na,0.5 mg E + 0.1MG NA,
17β-estradiol (E) +drospirenone (P) Oral, once a day 1 mg e=0.5 mg p
Oral intermittent combined
Micronized estradiol (E) + norgestimate (N) Oral, once/day 1 mg E + 0.09 mg NE (3 days on E/3 on E+P)

 

Transdermal Combinations
17β estradiol(E) + norethindrone acetate (P) One patch twice a week 0.05 mg E +0.25 mg P, 0.05 mg E +0.14 mg P,
17β estradiol (E)+ levonorgestrel (P) One patch once a week 0.45 mg E + 0.015 mg P
Micronized estradiol (ME) + norethindrone acetate (NA) Transdermal patch, replaced every 3-4 days 0.62 mg E + 2.7 mg NA
0.62 mg E + 4.8 mg NA

 

7. Therapeutic Hints:
  • Half doses of estrogen preparation can be used to decrease bleeding and breast tenderness. Recent data show good maintenance of bone with this approach.
  • Progestins may add to breast tenderness.
  • Moodiness and bloating may be due to progestins. Consider changing progestin, or using a lower dose of HT.
  • Consider using a patch in patients with high triglycerides to avoid “first pass” affect through the liver.
  • To improve HDL cholesterol use oral preparations. Increases are induced via a first pass mechanism.
  • Lower doses of HT/ERT have been found to be bone protective in doses equivalent to CEE 0.3mg.
  • Lower doses of HT can control vasomotor symptoms effectively when combined
    with progestin.
  • Lower doses or HT/ERT show favorable lipid profiles and changes are intermediate between standard dose and placebo.
  • Expect light bleeding or spotting in the first 3 months of therapy particularly with combined regimens. Abnormal bleeedoing (not the withdrawal bleeding after the progestin in a cyclic regimen should be evaluated with a pelvi/vaginal sonogram. If the endometrium is greater than 5mm an endometrial biopsy should be done to rule out hyperplasia.

 

REFERENCES

1. McKinlay SM, Brambilla DJ, Posner JG. The normal menopause transition. Maturitas. 1992;14(2):103-15.

2. Sevringhaus EL, Evans JS. Clinical observations on the use of an ovarian hormone: amniotin. Am J Med Sci. 1929;178:638-45.

3. Davis SR, Dinatale I, Rivera-Woll L, Davison S. Postmenopausal hormone therapy: from monkey glands to transdermal patches. J Endocrinol. 2005;185(2):207-22.

4. Ettinger B. Overview of estrogen replacement therapy: a historical perspective. Proc Soc Exp Biol Med. 1998;217(1):2-5.

5. Schmidt-Gollwitzer K. Estrogen/hormone replacement therapy present and past. Gynecol Endocrinol. 2001;15(suppl 4):11-6.

6. Nelson HD. Postmenopausal estrogen for treatment of hot flashes: clinical applications. JAMA. 2004;291(13):1621-5.

7. ASRM. The menopausal transition. Fertil Steril. 2008;90(5 Suppl):S61-5.

8. Longcope C, Franz C, Morello C, Baker R, Johnston CC, Jr. Steroid and gonadotropin levels in women during the peri-menopausal years. Maturitas. 1986;8(3):189-96.

9. Harlow SD, Gass M, Hall JE, Lobo R, Maki P, Rebar RW, et al. Executive summary of the Stages of Reproductive Aging Workshop + 10: addressing the unfinished agenda of staging reproductive aging. Menopause. 2012;19(4):387-95.

10. Gold EB, Bromberger J, Crawford S, Samuels S, Greendale GA, Harlow SD, et al. Factors associated with age at natural menopause in a multiethnic sample of midlife women. Am J Epidemiol. 2001;153(9):865-74.

11. Bachmann GA. Vasomotor flushes in menopausal women. Am J Obstet Gynecol. 1999;180(3 Pt 2):S312-6.

12. Haney AF, Wild RA. Options for hormone therapy in women who have had a hysterectomy. Menopause. 2007;14(3 Pt 2):592-7; quiz 8-9.

13. Rossouw JE, Anderson GL, Prentice RL, LaCroix AZ, Kooperberg C, Stefanick ML, et al. Risks and benefits of estrogen plus progestin in healthy postmenopausal women: principal results From the Women's Health Initiative randomized controlled trial. JAMA. 2002;288(3):321-33.

14. Anderson GL, Limacher M, Assaf AR, Bassford T, Beresford SA, Black H, et al. Effects of conjugated equine estrogen in postmenopausal women with hysterectomy: the Women's Health Initiative randomized controlled trial. Jama. 2004;291(14):1701-12.

15. Manson JE, Hsia J, Johnson KC, Rossouw JE, Assaf AR, Lasser NL, et al. Estrogen plus progestin and the risk of coronary heart disease. N Engl J Med. 2003;349(6):523-34.

16. de Villiers TJ, Gass ML, Haines CJ, Hall JE, Lobo RA, Pierroz DD, et al. Global consensus statement on menopausal hormone therapy. Climacteric. 2013;16(2):203-4.

17. PEPI. Effects of estrogen or estrogen/progestin regimens on heart disease risk factors in postmenopausal women. The Postmenopausal Estrogen/Progestin Interventions (PEPI) Trial. The Writing Group for the PEPI Trial. JAMA. 1995;273(3):199-208.

18. Hulley S, Grady D, Bush T, Furberg C, Herrington D, Riggs B, et al. Randomized trial of estrogen plus progestin for secondary prevention of coronary heart disease in postmenopausal women. Heart and Estrogen/progestin Replacement Study (HERS) Research Group. JAMA. 1998;280(7):605-13.

19. Colditz GA, Egan KM, Stampfer MJ. Hormone replacement therapy and risk of breast cancer: results from epidemiologic studies. Am J Obstet Gynecol. 1993;168(5):1473-80.

20. Grodstein F, Stampfer MJ, Colditz GA, Willett WC, Manson JE, Joffe M, et al. Postmenopausal hormone therapy and mortality. N Engl J Med. 1997;336(25):1769-75.

21. Grodstein F, Manson JE, Stampfer MJ. Postmenopausal hormone use and secondary prevention of coronary events in the nurses' health study. a prospective, observational study. Ann Intern Med. 2001;135(1):1-8.

22. Grodstein F, Stampfer MJ, Goldhaber SZ, Manson JE, Colditz GA, Speizer FE, et al. Prospective study of exogenous hormones and risk of pulmonary embolism in women. Lancet. 1996;348(9033):983-7.

23. Finkelstein JS, Brockwell SE, Mehta V, Greendale GA, Sowers MR, Ettinger B, et al. Bone mineral density changes during the menopause transition in a multiethnic cohort of women. J Clin Endocrinol Metab. 2008;93(3):861-8.

24. Hess R, Colvin A, Avis NE, Bromberger JT, Schocken M, Johnston JM, et al. The impact of hormone therapy on health-related quality of life: longitudinal results from the Study of Women's Health Across the Nation. Menopause. 2008;15(3):422-8.

25. Avis NE, Crawford SL, Greendale G, Bromberger JT, Everson-Rose SA, Gold EB, et al. Duration of Menopausal Vasomotor Symptoms Over the Menopause Transition. JAMA Intern Med. 2015.

26. Beral V. Breast cancer and hormone-replacement therapy in the Million Women Study. Lancet. 2003;362(9382):419-27.

27. Vickers MR, MacLennan AH, Lawton B, Ford D, Martin J, Meredith SK, et al. Main morbidities recorded in the women's international study of long duration oestrogen after menopause (WISDOM): a randomised controlled trial of hormone replacement therapy in postmenopausal women. BMJ. 2007;335(7613):239.

28. Hodis HN, Mack WJ, Shoupe D, Azen SP, Stanczyk FZ, Hwang-Levine J, et al. Methods and baseline cardiovascular data from the Early versus Late Intervention Trial with Estradiol testing the menopausal hormone timing hypothesis. Menopause. 2014.

29. Cohen LS, Soares CN, Vitonis AF, Otto MW, Harlow BL. Risk for new onset of depression during the menopausal transition: the Harvard study of moods and cycles. Arch Gen Psychiatry. 2006;63(4):385-90.

30. Schierbeck LL, Rejnmark L, Tofteng CL, Stilgren L, Eiken P, Mosekilde L, et al. Effect of hormone replacement therapy on cardiovascular events in recently postmenopausal women: randomised trial. BMJ. 2012;345:e6409.

31. Freeman EW, Sherif K. Prevalence of hot flushes and night sweats around the world: a systematic review. Climacteric. 2007;10(3):197-214.

32. Grady D. Clinical practice. Management of menopausal symptoms. N Engl J Med. 2006;355(22):2338-47.

33. Kronenberg F. Hot flashes: epidemiology and physiology. Ann N Y Acad Sci. 1990;592:52-86; discussion 123-33.

34. Nelson HD. Commonly used types of postmenopausal estrogen for treatment of hot flashes: scientific review. JAMA. 2004;291(13):1610-20.

35. Utian WH, Shoupe D, Bachmann G, Pinkerton JV, Pickar JH. Relief of vasomotor symptoms and vaginal atrophy with lower doses of conjugated equine estrogens and medroxyprogesterone acetate. Fertil Steril. 2001;75(6):1065-79.

36. Good WR, John VA, Ramirez M, Higgins JE. Comparison of Alora estradiol matrix transdermal delivery system with oral conjugated equine estrogen therapy in relieving menopausal symptoms. Alora Study Group. Climacteric. 1999;2(1):29-36.

37. Utian WH. Psychosocial and socioeconomic burden of vasomotor symptoms in menopause: a comprehensive review. Health Qual Life Outcomes. 2005;3:47.

38. Woods NF, Mitchell ES. Symptoms during the perimenopause: prevalence, severity, trajectory, and significance in women's lives. Am J Med. 2005;118 Suppl 12B:14-24.

39. Freedman RR, Roehrs TA. Effects of REM sleep and ambient temperature on hot flash-induced sleep disturbance. Menopause. 2006;13(4):576-83.

40. Freedman RR, Roehrs TA. Sleep disturbance in menopause. Menopause. 2007;14(5):826-9.

41. Schiff I, Regestein Q, Tulchinsky D, Ryan KJ. Effects of estrogens on sleep and psychological state of hypogonadal women. JAMA. 1979;242(22):2405-4.

42. Harlow BL, Wise LA, Otto MW, Soares CN, Cohen LS. Depression and its influence on reproductive endocrine and menstrual cycle markers associated with perimenopause: the Harvard Study of Moods and Cycles. Arch Gen Psychiatry. 2003;60(1):29-36.

43. Bromberger JT, Meyer PM, Kravitz HM, Sommer B, Cordal A, Powell L, et al. Psychologic distress and natural menopause: a multiethnic community study. Am J Public Health. 2001;91(9):1435-42.

44. Bromberger JT, Matthews KA, Schott LL, Brockwell S, Avis NE, Kravitz HM, et al. Depressive symptoms during the menopausal transition: the Study of Women's Health Across the Nation (SWAN). J Affect Disord. 2007;103(1-3):267-72.

45. Freeman EW, Sammel MD, Liu L, Gracia CR, Nelson DB, Hollander L. Hormones and menopausal status as predictors of depression in women in transition to menopause. Arch Gen Psychiatry. 2004;61(1):62-70.

46. Bromberger JT, Kravitz HM, Matthews K, Youk A, Brown C, Feng W. Predictors of first lifetime episodes of major depression in midlife women. Psychol Med. 2009;39(1):55-64.

47. Avis NE, Brambilla D, McKinlay SM, Vass K. A longitudinal analysis of the association between menopause and depression. Results from the Massachusetts Women's Health Study. Ann Epidemiol. 1994;4(3):214-20.

48. Dennerstein L, Guthrie JR, Clark M, Lehert P, Henderson VW. A population-based study of depressed mood in middle-aged, Australian-born women. Menopause. 2004;11(5):563-8.

49. Schmidt PJ, Haq N, Rubinow DR. A longitudinal evaluation of the relationship between reproductive status and mood in perimenopausal women. Am J Psychiatry. 2004;161(12):2238-44.

50. Schmidt PJ, Rubinow DR. Sex hormones and mood in the perimenopause. Ann N Y Acad Sci. 2009;1179:70-85.

51. Callegari C, Buttarelli M, Cromi A, Diurni M, Salvaggio F, Bolis PF. Female psychopathologic profile during menopausal transition: a preliminary study. Maturitas. 2007;56(4):447-51.

52. Steiner M, Dunn E, Born L. Hormones and mood: from menarche to menopause and beyond. J Affect Disord. 2003;74(1):67-83.

53. Joffe H, Hall JE, Soares CN, Hennen J, Reilly CJ, Carlson K, et al. Vasomotor symptoms are associated with depression in perimenopausal women seeking primary care. Menopause. 2002;9(6):392-8.

54. Woods NF, Smith-DiJulio K, Percival DB, Tao EY, Mariella A, Mitchell S. Depressed mood during the menopausal transition and early postmenopause: observations from the Seattle Midlife Women's Health Study. Menopause. 2008;15(2):223-32.

55. Steinberg EM, Rubinow DR, Bartko JJ, Fortinsky PM, Haq N, Thompson K, et al. A cross-sectional evaluation of perimenopausal depression. J Clin Psychiatry. 2008;69(6):973-80.

56. Harlow BL, Signorello LB. Factors associated with early menopause. Maturitas. 2000;35(1):3-9.

57. Gyllstrom ME, Schreiner PJ, Harlow BL. Perimenopause and depression: strength of association, causal mechanisms and treatment recommendations. Best Pract Res Clin Obstet Gynaecol. 2007;21(2):275-92.

58. Soares CN, Almeida OP, Joffe H, Cohen LS. Efficacy of estradiol for the treatment of depressive disorders in perimenopausal women: a double-blind, randomized, placebo-controlled trial. Arch Gen Psychiatry. 2001;58(6):529-34.

59. Morrison MF, Kallan MJ, Ten Have T, Katz I, Tweedy K, Battistini M. Lack of efficacy of estradiol for depression in postmenopausal women: a randomized, controlled trial. Biol Psychiatry. 2004;55(4):406-12.

60. Harman M. Primary findings of the Kronos Early Prevention Study (KEEPS). Proceedings from the 23rd Anuual Meeting of the North American Menopause Society. Orlando Florida; 2012.

61. Parry BL. Perimenopausal depression. Am J Psychiatry. 2008;165(1):23-7.

62. Panay N, Studd J. Progestogen intolerance and compliance with hormone replacement therapy in menopausal women. Hum Reprod Update. 1997;3(2):159-71.

63. Sherwin BB, Gelfand MM. Sex steroids and affect in the surgical menopause: a double-blind, cross-over study. Psychoneuroendocrinology. 1985;10(3):325-35.

64. Sherwin BB. Affective changes with estrogen and androgen replacement therapy in surgically menopausal women. J Affect Disord. 1988;14(2):177-87.

65. Barlow DH, Samsioe G, van Geelen JM. A study of European womens' experience of the problems of urogenital ageing and its management. Maturitas. 1997;27(3):239-47.

66. Oriba HA, Maibach HI. Vulvar transepidermal water loss (TEWL) decay curves. Effect of occlusion, delipidation, and age. Acta Derm Venereol. 1989;69(6):461-5.

67. Pinkerton JV, Stovall DW, Kightlinger RS. Advances in the treatment of menopausal symptoms. Womens Health (Lond Engl). 2009;5(4):361-84; quiz 83-4.

68. Raymundo N, Yu-cheng B, Zi-yan H, Lai CH, Leung K, Subramaniam R, et al. Treatment of atrophic vaginitis with topical conjugated equine estrogens in postmenopausal Asian women. Climacteric. 2004;7(3):312-8.

69. Bachmann G, Lobo RA, Gut R, Nachtigall L, Notelovitz M. Efficacy of low-dose estradiol vaginal tablets in the treatment of atrophic vaginitis: a randomized controlled trial. Obstet Gynecol. 2008;111(1):67-76.

70. Biglia N, Peano E, Sgandurra P, Moggio G, Panuccio E, Migliardi M, et al. Low-dose vaginal estrogens or vaginal moisturizer in breast cancer survivors with urogenital atrophy: a preliminary study. Gynecol Endocrinol.

71. Santen RJ, Pinkerton JV, Conaway M, Ropka M, Wisniewski L, Demers L, et al. Treatment of urogenital atrophy with low-dose estradiol: preliminary results. Menopause. 2002;9(3):179-87.

72. Management of symptomatic vulvovaginal atrophy: 2013 position statement of The North American Menopause Society. Menopause. 2013;20(9):888-902; quiz 3-4.

73. Sherwin BB, Gelfand MM. The role of androgen in the maintenance of sexual functioning in oophorectomized women. Psychosom Med. 1987;49(4):397-409.

74. Cummings SR, Black DM, Rubin SM. Lifetime risks of hip, Colles', or vertebral fracture and coronary heart disease among white postmenopausal women. Arch Intern Med. 1989;149(11):2445-8.

75. Soyka LA, Fairfield WP, Klibanski A. Clinical review 117: Hormonal determinants and disorders of peak bone mass in children. J Clin Endocrinol Metab. 2000;85(11):3951-63.

76. Riggs BL, Wahner HW, Melton LJ, 3rd, Richelson LS, Judd HL, Offord KP. Rates of bone loss in the appendicular and axial skeletons of women. Evidence of substantial vertebral bone loss before menopause. J Clin Invest. 1986;77(5):1487-91.

77. Recker R, Lappe J, Davies K, Heaney R. Characterization of perimenopausal bone loss: a prospective study. J Bone Miner Res. 2000;15(10):1965-73.

78. FDA. Questions and answers for estrogen and estrogen with progestin therapies for postmenopausal women (updated) http://www.fda.gov/Drugs/DrugSafety/InformationbyDrugClass/ucm135339.htm accessed 3/29/2010. 2010.

79. Corson SL. A practical guide to prescribing estrogen replacement therapy. Int J Fertil Menopausal Stud. 1995;40(5):229-47.

80. Manson JE, Bassuk SS. Calcium supplements: do they help or harm? Menopause. 2014;21(1):106-8.

81. Ross AC, Manson JE, Abrams SA, Aloia JF, Brannon PM, Clinton SK, et al. The 2011 report on dietary reference intakes for calcium and vitamin D from the Institute of Medicine: what clinicians need to know. J Clin Endocrinol Metab. 2011;96(1):53-8.

82. Holick MF, Siris ES, Binkley N, Beard MK, Khan A, Katzer JT, et al. Prevalence of Vitamin D inadequacy among postmenopausal North American women receiving osteoporosis therapy. J Clin Endocrinol Metab. 2005;90(6):3215-24.

83. Anderson RN. Deaths: leading causes for 1999. Natl Vital Stat Rep. 2001;49(11):1-87.

84. Matthews KA, Meilahn E, Kuller LH, Kelsey SF, Caggiula AW, Wing RR. Menopause and risk factors for coronary heart disease. N Engl J Med. 1989;321(10):641-6.

85. Heckbert SR, Kaplan RC, Weiss NS, Psaty BM, Lin D, Furberg CD, et al. Risk of recurrent coronary events in relation to use and recent initiation of postmenopausal hormone therapy. Arch Intern Med. 2001;161(14):1709-13.

86. Alexander KP, Newby LK, Hellkamp AS, Harrington RA, Peterson ED, Kopecky S, et al. Initiation of hormone replacement therapy after acute myocardial infarction is associated with more cardiac events during follow-up. J Am Coll Cardiol. 2001;38(1):1-7.

87. Hodis HN, Mack WJ, Lobo RA, Shoupe D, Sevanian A, Mahrer PR, et al. Estrogen in the prevention of atherosclerosis. A randomized, double-blind, placebo-controlled trial. Ann Intern Med. 2001;135(11):939-53.

88. Rossouw JE, Prentice RL, Manson JE, Wu L, Barad D, Barnabei VM, et al. Postmenopausal hormone therapy and risk of cardiovascular disease by age and years since menopause. JAMA. 2007;297(13):1465-77.

89. Manson JE, Allison MA, Rossouw JE, Carr JJ, Langer RD, Hsia J, et al. Estrogen therapy and coronary-artery calcification. N Engl J Med. 2007;356(25):2591-602.

90. Prentice RL, Manson JE, Langer RD, Anderson GL, Pettinger M, Jackson RD, et al. Benefits and risks of postmenopausal hormone therapy when it is initiated soon after menopause. Am J Epidemiol. 2009;170(1):12-23.

91. Banks E, Canfell K. Invited Commentary: Hormone therapy risks and benefits--The Women's Health Initiative findings and the postmenopausal estrogen timing hypothesis. Am J Epidemiol. 2009;170(1):24-8.

92. Toh S, Hernandez-Diaz S, Logan R, Rossouw JE, Hernan MA. Coronary heart disease in postmenopausal recipients of estrogen plus progestin therapy: does the increased risk ever disappear? A randomized trial. Ann Intern Med;152(4):211-7.

93. Stevenson JC, Hodis HN, Pickar JH, Lobo RA. Coronary heart disease and menopause management: the swinging pendulum of HRT. Atherosclerosis. 2009;207(2):336-40.

94. Harman M. Primary findings of the Kronos Early Prevention Study (KEEPS). Proceedings from the 23rd Meeting of the North American Menopause Society. October 3-6 2012.

95. Hodis HN. Latest Data from the Elite Trial. International Menopause Society 14 World Congress on Menopause 2014. Cancun Mexico.

96. Heiss G, Wallace R, Anderson GL, Aragaki A, Beresford SA, Brzyski R, et al. Health risks and benefits 3 years after stopping randomized treatment with estrogen and progestin. JAMA. 2008;299(9):1036-45.

97. Bechlioulis A, Kalantaridou SN, Naka KK, Chatzikyriakidou A, Calis KA, Makrigiannakis A, et al. Endothelial function, but not carotid intima-media thickness, is affected early in menopause and is associated with severity of hot flushes. J Clin Endocrinol Metab;95(3):1199-206.

98. Matthews KA, Crawford SL, Chae CU, Everson-Rose SA, Sowers MF, Sternfeld B, et al. Are changes in cardiovascular disease risk factors in midlife women due to chronological aging or to the menopausal transition? J Am Coll Cardiol. 2009;54(25):2366-73.

99. Herrington DM, Reboussin DM, Brosnihan KB, Sharp PC, Shumaker SA, Snyder TE, et al. Effects of estrogen replacement on the progression of coronary-artery atherosclerosis. N Engl J Med. 2000;343(8):522-9.

100. Fahraeus L. The effects of estradiol on blood lipids and lipoproteins in postmenopausal women. Obstet Gynecol. 1988;72(5 Suppl):18S-22S.

101. Effects of estrogen or estrogen/progestin regimens on heart disease risk factors in postmenopausal women. The Postmenopausal Estrogen/Progestin Interventions (PEPI) Trial. The Writing Group for the PEPI Trial. JAMA. 1995;273(3):199-208.

102. Sarrel PM, Lindsay D, Rosano GM, Poole-Wilson PA. Angina and normal coronary arteries in women: gynecologic findings. Am J Obstet Gynecol. 1992;167(2):467-71.

103. Rosano GM, Sarrel PM, Poole-Wilson PA, Collins P. Beneficial effect of oestrogen on exercise-induced myocardial ischaemia in women with coronary artery disease. Lancet. 1993;342(8864):133-6.

104. Gangar KF, Vyas S, Whitehead M, Crook D, Meire H, Campbell S. Pulsatility index in internal carotid artery in relation to transdermal oestradiol and time since menopause. Lancet. 1991;338(8771):839-42.

105. Viscoli CM, Brass LM, Kernan WN, Sarrel PM, Suissa S, Horwitz RI. A clinical trial of estrogen-replacement therapy after ischemic stroke. N Engl J Med. 2001;345(17):1243-9.

106. Warren MP, Richardson, O, Chaundry,S.et al. The effect of Estrogen and Hormone Withdrawal on Health and Quality of Life after Publication of the Women's Health Initiative in New York City. Abstracts,Meeting of the North American Menopause Society Washington DC Sept. 2011;S-2:32.

107. Feuer EJ, Wun LM, Boring CC, Flanders WD, Timmel MJ, Tong T. The lifetime risk of developing breast cancer. J Natl Cancer Inst. 1993;85(11):892-7.

108. Anderson GL, Chlebowski RT, Aragaki AK, Kuller LH, Manson JE, Gass M, et al. Conjugated equine oestrogen and breast cancer incidence and mortality in postmenopausal women with hysterectomy: extended follow-up of the Women's Health Initiative randomised placebo-controlled trial. Lancet Oncol. 2012;13(5):476-86.

109. Anderson GL, Chlebowski RT, Rossouw JE, Rodabough RJ, McTiernan A, Margolis KL, et al. Prior hormone therapy and breast cancer risk in the Women's Health Initiative randomized trial of estrogen plus progestin. Maturitas. 2006;55(2):103-15.

110. Santen RJ, Song Y, Yue W, Wang JP, Heitjan DF. Effects of menopausal hormonal therapy on occult breast tumors. J Steroid Biochem Mol Biol. 2013;137:150-6.

111. Chlebowski RT, Hendrix SL, Langer RD, Stefanick ML, Gass M, Lane D, et al. Influence of estrogen plus progestin on breast cancer and mammography in healthy postmenopausal women: the Women's Health Initiative Randomized Trial. Jama. 2003;289(24):3243-53.

112. Breast cancer and hormone replacement therapy: collaborative reanalysis of data from 51 epidemiological studies of 52,705 women with breast cancer and 108,411 women without breast cancer. Collaborative Group on Hormonal Factors in Breast Cancer. Lancet. 1997;350(9084):1047-59.

113. Sellers TA, Mink PJ, Cerhan JR, Zheng W, Anderson KE, Kushi LH, et al. The role of hormone replacement therapy in the risk for breast cancer and total mortality in women with a family history of breast cancer. Ann Intern Med. 1997;127(11):973-80.

114. Gapstur SM, Morrow M, Sellers TA. Hormone replacement therapy and risk of breast cancer with a favorable histology: results of the Iowa Women's Health Study. JAMA. 1999;281(22):2091-7.

115. Bonnier P, Romain S, Giacalone PL, Laffargue F, Martin PM, Piana L. Clinical and biologic prognostic factors in breast cancer diagnosed during postmenopausal hormone replacement therapy. Obstet Gynecol. 1995;85(1):11-7.

116. Bergkvist L, Adami HO, Persson I, Bergstrom R, Krusemo UB. Prognosis after breast cancer diagnosis in women exposed to estrogen and estrogen-progestogen replacement therapy. Am J Epidemiol. 1989;130(2):221-8.

117. Boyd NF, Lockwood GA, Martin LJ, Knight JA, Byng JW, Yaffe MJ, et al. Mammographic densities and breast cancer risk. Breast Dis. 1998;10(3-4):113-26.

118. Greendale GA, Reboussin BA, Sie A, Singh HR, Olson LK, Gatewood O, et al. Effects of estrogen and estrogen-progestin on mammographic parenchymal density. Postmenopausal Estrogen/Progestin Interventions (PEPI) Investigators. Ann Intern Med. 1999;130(4 Pt 1):262-9.

119. Rutter CM, Mandelson MT, Laya MB, Seger DJ, Taplin S. Changes in breast density associated with initiation, discontinuation, and continuing use of hormone replacement therapy. JAMA. 2001;285(2):171-6.

120. Ross RK, Paganini-Hill A, Wan PC, Pike MC. Effect of hormone replacement therapy on breast cancer risk: estrogen versus estrogen plus progestin. J Natl Cancer Inst. 2000;92(4):328-32.

121. Holli K, Isola J, Cuzick J. Low biologic aggressiveness in breast cancer in women using hormone replacement therapy. J Clin Oncol. 1998;16(9):3115-20.

122. Rebbeck TR, Levin AM, Eisen A, Snyder C, Watson P, Cannon-Albright L, et al. Breast cancer risk after bilateral prophylactic oophorectomy in BRCA1 mutation carriers. J Natl Cancer Inst. 1999;91(17):1475-9.

123. Durna EM, Wren BG, Heller GZ, Leader LR, Sjoblom P, Eden JA. Hormone replacement therapy after a diagnosis of breast cancer: cancer recurrence and mortality. Med J Aust. 2002;177(7):347-51.

124. O'Meara ES, Rossing MA, Daling JR, Elmore JG, Barlow WE, Weiss NS. Hormone replacement therapy after a diagnosis of breast cancer in relation to recurrence and mortality. J Natl Cancer Inst. 2001;93(10):754-62.

125. Fournier A, Mesrine S, Boutron-Ruault MC, Clavel-Chapelon F. Estrogen-progestagen menopausal hormone therapy and breast cancer: does delay from menopause onset to treatment initiation influence risks? J Clin Oncol. 2009;27(31):5138-43.

126. Ravdin PM, Cronin KA, Howlader N, Berg CD, Chlebowski RT, Feuer EJ, et al. The decrease in breast-cancer incidence in 2003 in the United States. N Engl J Med. 2007;356(16):1670-4.

127. Jemal A, Siegel R, Ward E, Hao Y, Xu J, Thun MJ. Cancer statistics, 2009. CA Cancer J Clin. 2009;59(4):225-49.

128. Li CI, Malone KE, Porter PL, Lawton TJ, Voigt LF, Cushing-Haugen KL, et al. Relationship between menopausal hormone therapy and risk of ductal, lobular, and ductal-lobular breast carcinomas. Cancer Epidemiol Biomarkers Prev. 2008;17(1):43-50.

129. Stefanick ML, Anderson GL, Margolis KL, Hendrix SL, Rodabough RJ, Paskett ED, et al. Effects of conjugated equine estrogens on breast cancer and mammography screening in postmenopausal women with hysterectomy. JAMA. 2006;295(14):1647-57.

130. Harvey JA, Pinkerton JV, Herman CR. Short-term cessation of hormone replacement therapy and improvement of mammographic specificity. J Natl Cancer Inst. 1997;89(21):1623-5.

131. Biglia N, Mariani L, Ponzone R, Sismondi P. Oral contraceptives, salpingo-oophorectomy and hormone replacement therapy in BRCA1-2 mutation carriers. Maturitas. 2008;60(2):71-7.

132. Manson JE, Chlebowski RT, Stefanick ML, Aragaki AK, Rossouw JE, Prentice RL, et al. Menopausal hormone therapy and health outcomes during the intervention and extended poststopping phases of the Women's Health Initiative randomized trials. Jama. 2013;310(13):1353-68.

133. Hormone therapy for the prevention of chronic conditions in postmenopausal women: recommendations from the U.S. Preventive Services Task Force. Ann Intern Med. 2005;142(10):855-60.

134. Kotsopoulos J, Lubinski J, Neuhausen SL, Lynch HT, Rosen B, Ainsworth P, et al. Hormone replacement therapy and the risk of ovarian cancer in BRCA1 and BRCA2 mutation carriers. Gynecol Oncol. 2006;100(1):83-8.

135. Anderson GL, Judd HL, Kaunitz AM, Barad DH, Beresford SA, Pettinger M, et al. Effects of estrogen plus progestin on gynecologic cancers and associated diagnostic procedures: the Women's Health Initiative randomized trial. JAMA. 2003;290(13):1739-48.

136. Hulley S, Furberg C, Barrett-Connor E, Cauley J, Grady D, Haskell W, et al. Noncardiovascular disease outcomes during 6.8 years of hormone therapy: Heart and Estrogen/progestin Replacement Study follow-up (HERS II). JAMA. 2002;288(1):58-66.

137. Coughlin SS, Giustozzi A, Smith SJ, Lee NC. A meta-analysis of estrogen replacement therapy and risk of epithelial ovarian cancer. J Clin Epidemiol. 2000;53(4):367-75.

138. Persson I, Yuen J, Bergkvist L, Schairer C. Cancer incidence and mortality in women receiving estrogen and estrogen-progestin replacement therapy--long-term follow-up of a Swedish cohort. Int J Cancer. 1996;67(3):327-32.

139. Gambrell RD, Jr. The menopause: benefits and risks of estrogen-progestogen replacement therapy. Fertil Steril. 1982;37(4):457-74.

140. Pickar JH, Yeh IT, Wheeler JE, Cunnane MF, Speroff L. Endometrial effects of lower doses of conjugated equine estrogens and medroxyprogesterone acetate: two-year substudy results. Fertil Steril. 2003;80(5):1234-40.

141. Bartus RT, Dean RL, 3rd, Beer B, Lippa AS. The cholinergic hypothesis of geriatric memory dysfunction. Science. 1982;217(4558):408-14.

142. Ditkoff EC, Crary WG, Cristo M, Lobo RA. Estrogen improves psychological function in asymptomatic postmenopausal women. Obstet Gynecol. 1991;78(6):991-5.

143. Kampen DL, Sherwin BB. Estrogen use and verbal memory in healthy postmenopausal women. Obstet Gynecol. 1994;83(6):979-83.

144. Sherwin BB. Estrogen effects on cognition in menopausal women. Neurology. 1997;48(5 Suppl 7):S21-6.

145. Resnick SM, Metter EJ, Zonderman AB. Estrogen replacement therapy and longitudinal decline in visual memory. A possible protective effect? Neurology. 1997;49(6):1491-7.

146. Behl C, Skutella T, Lezoualc'h F, Post A, Widmann M, Newton CJ, et al. Neuroprotection against oxidative stress by estrogens: structure-activity relationship. Mol Pharmacol. 1997;51(4):535-41.

147. Singer CA, Figueroa-Masot XA, Batchelor RH, Dorsa DM. The mitogen-activated protein kinase pathway mediates estrogen neuroprotection after glutamate toxicity in primary cortical neurons. J Neurosci. 1999;19(7):2455-63.

148. Dubal DB, Wilson ME, Wise PM. Estradiol: a protective and trophic factor in the brain. J Alzheimers Dis. 1999;1(4-5):265-74.

149. Maki PM, Sundermann E. Hormone therapy and cognitive function. Hum Reprod Update. 2009;15(6):667-81.

150. Pefanco MA, Kenny AM, Kaplan RF, Kuchel G, Walsh S, Kleppinger A, et al. The effect of 3-year treatment with 0.25 mg/day of micronized 17beta-estradiol on cognitive function in older postmenopausal women. J Am Geriatr Soc. 2007;55(3):426-31.

151. Maki PM, Drogos LL, Rubin LH, Banuvar S, Shulman LP, Geller SE. Objective hot flashes are negatively related to verbal memory performance in midlife women. Menopause. 2008;15(5):848-56.

152. Evans DA, Funkenstein HH, Albert MS, Scherr PA, Cook NR, Chown MJ, et al. Prevalence of Alzheimer's disease in a community population of older persons. Higher than previously reported. JAMA. 1989;262(18):2551-6.

153. Berlinger WG, Potter JF. Low Body Mass Index in demented outpatients. J Am Geriatr Soc. 1991;39(10):973-8.

154. Mulnard RA, Cotman CW, Kawas C, van Dyck CH, Sano M, Doody R, et al. Estrogen replacement therapy for treatment of mild to moderate Alzheimer disease: a randomized controlled trial. Alzheimer's Disease Cooperative Study. JAMA. 2000;283(8):1007-15.

155. Kawas C, Resnick S, Morrison A, Brookmeyer R, Corrada M, Zonderman A, et al. A prospective study of estrogen replacement therapy and the risk of developing Alzheimer's disease: the Baltimore Longitudinal Study of Aging. Neurology. 1997;48(6):1517-21.

156. Zandi PP, Carlson MC, Plassman BL, Welsh-Bohmer KA, Mayer LS, Steffens DC, et al. Hormone replacement therapy and incidence of Alzheimer disease in older women: the Cache County Study. JAMA. 2002;288(17):2123-9.

157. Shumaker SA, Legault C, Rapp SR, Thal L, Wallace RB, Ockene JK, et al. Estrogen plus progestin and the incidence of dementia and mild cognitive impairment in postmenopausal women: the Women's Health Initiative Memory Study: a randomized controlled trial. JAMA. 2003;289(20):2651-62.

158. Grodstein F, Colditz GA, Stampfer MJ. Postmenopausal hormone use and cholecystectomy in a large prospective study. Obstet Gynecol. 1994;83(1):5-11.

159. Chen Z, Bassford T, Green SB, Cauley JA, Jackson RD, LaCroix AZ, et al. Postmenopausal hormone therapy and body composition--a substudy of the estrogen plus progestin trial of the Women's Health Initiative. Am J Clin Nutr. 2005;82(3):651-6.

160. Margolis KL, Bonds DE, Rodabough RJ, Tinker L, Phillips LS, Allen C, et al. Effect of oestrogen plus progestin on the incidence of diabetes in postmenopausal women: results from the Women's Health Initiative Hormone Trial. Diabetologia. 2004;47(7):1175-87.

161. Vingerling JR, Dielemans I, Witteman JC, Hofman A, Grobbee DE, de Jong PT. Macular degeneration and early menopause: a case-control study. BMJ. 1995;310(6994):1570-1.

162. Klein BE, Klein R, Jensen SC, Ritter LL. Are sex hormones associated with age-related maculopathy in women? The Beaver Dam Eye Study. Trans Am Ophthalmol Soc. 1994;92:289-95; discussion 95-7.

163. Risk factors for neovascular age-related macular degeneration. The Eye Disease Case-Control Study Group. Arch Ophthalmol. 1992;110(12):1701-8.

164. Haapasaari KM, Raudaskoski T, Kallioinen M, Suvanto-Luukkonen E, Kauppila A, Laara E, et al. Systemic therapy with estrogen or estrogen with progestin has no effect on skin collagen in postmenopausal women. Maturitas. 1997;27(2):153-62.

165. Chen L, Dyson M, Rymer J, Bolton PA, Young SR. The use of high-frequency diagnostic ultrasound to investigate the effect of hormone replacement therapy on skin thickness. Skin Res Technol. 2001;7(2):95-7.

166. Ashcroft GS, Dodsworth J, van Boxtel E, Tarnuzzer RW, Horan MA, Schultz GS, et al. Estrogen accelerates cutaneous wound healing associated with an increase in TGF-beta1 levels. Nat Med. 1997;3(11):1209-15.

167. Grodstein F, Colditz GA, Stampfer MJ. Post-menopausal hormone use and tooth loss: a prospective study. J Am Dent Assoc. 1996;127(3):370-7, quiz 92.

168. Simon JA, Portman DJ, Kaunitz AM, Mekonnen H, Kazempour K, Bhaskar S, et al. Low-dose paroxetine 7.5 mg for menopausal vasomotor symptoms: two randomized controlled trials. Menopause. 2013;20(10):1027-35.

169. Simon JA, Portman DJ, Kazempour K, Mekonnen H, Bhaskar S, Lippman J. Safety Profile of Paroxetine 7.5 mg in Women With Moderate-to-Severe Vasomotor Symptoms. Obstet Gynecol. 2014;123 Suppl 1:132s-3s.

170. Lobo RA, Pinkerton JV, Gass ML, Dorin MH, Ronkin S, Pickar JH, et al. Evaluation of bazedoxifene/conjugated estrogens for the treatment of menopausal symptoms and effects on metabolic parameters and overall safety profile. Fertil Steril. 2009;92(3):1025-38.

171. Pinkerton JV, Harvey JA, Lindsay R, Pan K, Chines AA, Mirkin S, et al. Effects of bazedoxifene/conjugated estrogens on the endometrium and bone: a randomized trial. J Clin Endocrinol Metab. 2014;99(2):E189-98.

172. Archer DF, Lewis V, Carr BR, Olivier S, Pickar JH. Bazedoxifene/conjugated estrogens (BZA/CE): incidence of uterine bleeding in postmenopausal women. Fertil Steril. 2009;92(3):1039-44.

173. Portman D, Palacios S, Nappi RE, Mueck AO. Ospemifene, a non-oestrogen selective oestrogen receptor modulator for the treatment of vaginal dryness associated with postmenopausal vulvar and vaginal atrophy: a randomised, placebo-controlled, phase III trial. Maturitas. 2014;78(2):91-8.

174. Constantine G, Graham S, Portman DJ, Rosen RC, Kingsberg SA. Female sexual function improved with ospemifene in postmenopausal women with vulvar and vaginal atrophy: results of a randomized, placebo-controlled trial. Climacteric. 2014:1-7.

175. Archer DF, Carr BR, Pinkerton JV, Taylor HS, Constantine GD. Effects of ospemifene on the female reproductive and urinary tracts: translation from preclinical models into clinical evidence. Menopause. 2014.

176. Winther K, Rein E, Hedman C. Femal, a herbal remedy made from pollen extracts, reduces hot flushes and improves quality of life in menopausal women: a randomized, placebo-controlled, parallel study. Climacteric. 2005;8(2):162-70.

177. Estrogen and progestogen use in postmenopausal women: 2010 position statement of The North American Menopause Society. Menopause;17(2):242-55.

178. Santen RJ, Allred DC, Ardoin SP, Archer DF, Boyd N, Braunstein GD, et al. Postmenopausal hormone therapy: an Endocrine Society scientific statement. J Clin Endocrinol Metab;95(7 Suppl 1):s1-s66.

179. Canonico M, Oger E, Plu-Bureau G, Conard J, Meyer G, Levesque H, et al. Hormone therapy and venous thromboembolism among postmenopausal women: impact of the route of estrogen administration and progestogens: the ESTHER study. Circulation. 2007;115(7):840-5.

180. Renoux C, Dell'aniello S, Garbe E, Suissa S. Transdermal and oral hormone replacement therapy and the risk of stroke: a nested case-control study. BMJ;340:c2519.

181. Fournier A, Berrino F, Clavel-Chapelon F. Unequal risks for breast cancer associated with different hormone replacement therapies: results from the E3N cohort study. Breast Cancer Res Treat. 2008;107(1):103-11.

182. Warren MP. Historical perspectives in postmenopausal hormone therapy: defining the right dose and duration. Mayo Clin Proc. 2007;82(2):219-26.

183. Ockene JK, Barad DH, Cochrane BB, Larson JC, Gass M, Wassertheil-Smoller S, et al. Symptom experience after discontinuing use of estrogen plus progestin. JAMA. 2005;294(2):183-93.

The Non-Thyroidal Illness Syndrome

 

ABSTRACT
NTIS refers to a syndrome found in seriously ill or starving patients with low fT3, usually elevated RT3, normal  or low TSH, and if prolonged, low fT4. It is found  in a high proportion of patients in the ICU setting, and correlates with a poor prognosis if TT4 is <4ug/dl. The patho-physiology includes suppression of TRH release, reducedT3 and T4 turnover, reduction in liver generation of T3, increased formation of RT3, and tissue specific down-regulation of deiodinases, transporters, and TH receptors. Although long debated, tissue TH levels are definitely reduced, and tissue hypothyroidism is presumably present. This is often not clinically evident because of the brief duration, and reduced but not absent tissue levels of TH. Although recognized for nearly 4 decades, interpretation of the syndrome is contested, because of lack of data. Some observes, totally without data, argue that it is a protective response and should not be treated. Other observers (as in this review) present available data suggesting, but not proving, that thyroid hormone replacement is appropriate, not harmful, and may be beneficial. The best form of treatment (TRH,TSH,or T3+T4) and possible accompanying treatments (GHRH, Cortisol, nutrition, insulin) lack consensus. In this review current data are laid out for reader’s review and judgment.

 

DEFINITIONS

Serum thyroid hormone levels drop during starvation and illness. In mild illness, this involves only a decrease in serum triiodothyronine (T3) levels. However, as the severity and length of the illness increases, there is a drop in both serum T3 and thyroxine (T4). This decrease of serum thyroid hormone levels is seen in starvation, sepsis, surgery, myocardial infarction, bypass, bone marrow transplantation, and in fact probably any severe illness.1-9 The condition has been called the euthyroid sick syndrome (ESS). An alternative designation, which does not presume the metabolic status of the patient, is nonthyroidal illness syndrome, or NTIS. For more than 3 decades the interpretation of these changes has been debated   Many observers have considered the changes in hormone level to be laboratory artifacts, or if valid, not representative of true hypothyroidism, or if hypothyroidism was present, that it was a beneficial response designed to “spare calories” (1-21). More recently evidence has accumulated that central hypothyroidism, and altered peripheral metabolism of T4 and T3, combine to produce a state marked by diminished serum and tissue supplies of thyroid hormones. Nevertheless, some observers accept the low hormone levels as valid, but maintain that this is a (unique) situation in which such lack of hormone is not truly hypothyroidism (i.e., the “euthyroid sick syndrome”). Lastly, there is even greater uncertainly about hormone replacement therapy, in considerable  part because the opinion that replacement treatment should not be given has been repeated so many times, even though there is effectively no factual support for that view. We need controlled clinical trials in order to answer the question. It can not be solved by oft-stated opinions.

Low T3 States

Starvation, and more precisely carbohydrate deprivation, appears to rapidly inhibit deiodination of T4 to T3 by type 1 iodothyronine deiodinase in the liver, thus inhibiting generation of T3 and preventing metabolism of reverse T3 (rT3).10 Consequently there is a drop in serum T3 and elevation of reverse T3. Since starvation induces a decrease in basal metabolic rate,11 it has been argued, teleologically, that this decrease in thyroid hormone represents an adaptive response by the body to spare calories and protein by inducing some degree of hypothyroidism. Patients who have only a drop in serum T3, representing the mildest form of the NTIS, do not show clinical signs of hypothyroidism. Nor has it been shown that this decrease in serum T3 (in the absence of a drop in T4) has an adverse physiologic effect on the body or that it is associated with increased mortality.

Nonthyroidal Illness Syndrome With Low Serum T4

As the severity of illness, and often associated starvation, progresses, there is the gradual development of a more complex syndrome associated with low T3 and usually low T4 levels. Generally thyroid-stimulating hormone (TSH) levels are low or normal despite the low serum hormone levels, and rT3 levels are normal or elevated. A large proportion of patients in an intensive care unit setting have various degrees of severity of NTIS with low T3 and T4. Plikat et al. found that 23% of patients admitted to an ICU during a 2-year period had low free T3, low free T4, and low or normal TSH, and that these findings gave a greatly increased risk of death.12 Girvent et al. note that NTIS is highly prevalent in elderly patients with acute surgical problems and is associated with poor nutrition, higher sympathetic response, and worse postoperative outcome.13 Surprisingly, during the past 4 decades, some endocrinologists have assumed that NTIS is a beneficial physiologic response,14-17 but factual evidence for this view is unavailable. However it seems illogical to consider  NTIS as an evolutionarily derived  physiologic response, since survival with the severity of illness seen in NTIS patients would be almost impossible except in modern ICUs.

A marked decrease in serum T3 and T4 in NTIS is associated with a high probability of death. NTIS was found in a group of 20 patients with severe trauma, among whom 5 died, and the drop in T3 correlated with the Apache II score.18 NTIS found in patients undergoing bone marrow transplantation was associated with a high probability of fatal outcome.19 NTIS was typical in elderly patients undergoing acute surgery and was associated with a worse prognosis.20 All of 45 non-dopamine-treated children with meningococcal septicemia had low T3, T4, and thyroxine-binding globulin (TBG), without elevated TSH. When serum T4 levels drop below 4 g/dL, the probability of death is about 50%, and with serum T4 levels below 2 g/dL, the probability of death reaches 80%.21-23 Obviously such associations do not prove that hypothyroidism is the cause of these complications or deaths, but the fact of hypothyroidism must at least raise the consideration of treatment.

Interpretations of NTIS

Several conceptual explanations of NTIS can be followed through the literature:

1.         The abnormalities represent test artifacts, and assays would indicate euthyroidism if proper tests were employed.

2.         The serum thyroid hormone abnormalities are due to inhibitors of T4 binding to proteins, and the tests do not appropriately reflect free hormone levels. Proponents of this concept may or may not take the position that a binding inhibitor is present throughout body tissues, rather than simply in serum, and that the binding inhibitor may also inhibit uptake of hormone by cells or prevent binding to nuclear T3 receptors and thus inhibit action of hormone.

3.         In NTIS, T3 levels in the pituitary are normal because of enhanced local deiodination. In this concept the pituitary is actually euthyroid, while the rest of the body is hypothyroid. This presupposes enhanced intrapituitary T4 > T3 deiodination as the cause.

4.         Serum hormone levels are in fact low, and the patients are biochemically hypothyroid, but this is (teleologically) a beneficial physiologic response and should not be altered by treatment.

5.         Lastly, NTIS is in part a form of secondary hypothyroidism, the patient’s serum and tissue hormone levels are truly low, tissue hypothyroidism is present, this is probably disadvantageous to the patient, and therapy should be initiated if serum thyroxine levels are depressed below the danger level of 4 μg/dL.

 

SERUM HORMONE LEVELS AND TISSUE HORMONE SUPPLIES IN NTIS

Serum T3 and Free T3

With few exceptions, reports on NTIS indicate that serum T3 and free T3 levels are low.24-30

Liver Iodothyronine D1 normally produces up to 80% of circulating T3 via T4>T3 deiodination, the remainder coming from the thyroid directly, or by a contribution from ID2 in muscle as noted below. ID1 in liver is down-regulated in severe illness, and this is certainly an important contributor to the low T3 in blood. One presumed cause is reduced nutrition, especially of carbohydrate, but direct effects of cytokines on liver may also be involved  The problem presumably is exacerbated by hypothyroidism, which also down-regulates ID1.

Chopra and co-workers reported that free T3 levels were low (Fig. 1),31 or in a second report, often normal.32 However, it is important to note that in the second report, the patients with “NTIS” actually had average serum T4 levels that were above the normal mean and did not have significant NTIS. Sapin et al. compared free T3 levels found in patients with NTIS by direct dialysis, microchromatography, analogue, two-step immune extraction, and a labeled antibody RIA method.30 Results were significantly below normal by five of the methods and low in the most severe cases by one method. Faber et al. evaluated thyroid hormone levels in 34 seriously ill patients, most of whom had low T4 and free T4 index values, and found generally normal free T3 and free T4 using an ultrafiltration technique.33 A point to consider is that some ultrafiltration techniques fail to exclude thyroid hormone–binding proteins from the filtrate and give spuriously high free hormone values.34

Figure 1. Free T3 concentrations in different groups of patients, as reported by Chopra et al, reference 32. In this report, patients with NTIS have significantly lowered Free T3 levels than do normal subjects.

Figure 1. Free T3 concentrations in different groups of patients, as reported by Chopra et al, reference 32. In this report, patients with NTIS have significantly lowered Free T3 levels than do normal subjects.

Serum rT3 may be reduced, normal, or elevated and is not a reliable indicator of abnormal thyroid hormone supply. While it may be expected that rT3 should always be elevated, this is not true, and often it is within the normal range. Peeters et al.35 found in patients with NTIS, serum TSH, T4, T3, and the T3/rT3 ratio were lower, whereas serum rT3 was higher than in normal subjects (P < 0.0001). Liver D1 is down-regulated, and D3 (which is not evident in liver and skeletal muscle of healthy individuals) is induced, particularly in disease states associated with poor tissue perfusion. The level of rT3 reflects the action of several enzymes and presumably, as well, tissue metabolic function. Induction of D3 would tend to increase rT3. Degradation of rT3 is reduced by decreased function of the same D1 enzyme that generates T3. However, formation of rT3 is limited by the low level of substrate (T4) in serum and in tissues and perhaps by inhibition of T4 entry into cells. Personal experience treating patients with NTIS (unpublished) shows that when T4 is given and repletes serum hormone levels, generation of rT3 rapidly increases, and levels often become significantly elevated.

Serum T4

Serum T4 levels are reduced in NTIS in proportion to the severity and, probably, length of the illness.24-35 In acute, short-term trauma such as cardiac bypass36 or in short-term starvation,37 there is no drop in serum T4. However, with increasing severity of trauma, illness, or infection, there is a drop in T4 which may become extreme. As indicated, serum T4 levels below 4 μg/dL are associated with a marked increased risk of death (up to 50%), and once T4 is below 2, prognosis becomes extremely guarded. In neonates, low total T4 and TSH are associated with a greater risk of death and severe intraventricular hemorrhage. It is suggested that thyroid hormone supplementation might be a potential benefit in infants with the lowest T4 values.27

Total serum T4 is reduced in part because of a reduction in TBG. One reason for this reduction appears to be because of cleavage of TBG. Schussler’s group recognized a rapid drop in TBG to 60% of baseline within 12 hours after bypass surgery, and their data suggest that this is due to cleavage of TBG by protease, which causes TBG to lose its T4-binding activity.38 Further studies by this group demonstrated the presence of a cleaved form of TBG present in serum of patients with sepsis.39

The impact of meningococcal sepsis on peripheral thyroid hormone metabolism and binding proteins was studied in 69 children with meningococcal sepsis. All children had decreased total T3 and total T3/rT3 ratios without elevated TSH. Lower total T4 levels were related to increased turnover of TBG by elastase. Lowered TBG is a partial explanation for lower total T4 and T3 in NTIS.40

Serum Free Thyroxine

A major problem in understanding NTIS is in analyzing data on the level of free T4. Free T4 is believed by most workers to represent hormone availability to tissues, although it is in fact intracellular T3 that binds to the receptors. The results of free T4 assays in NTIS are definitely method dependent. They may be influenced by a variety of variables, including (alleged) inhibitors present in serum or the effect of agents such as drugs, metabolites, or free fatty acids in the serum or assay. Assays which include an estimate of TBG capacity to estimate free hormone typically return low values for calculated free thyroxine in NTIS. Methods using T3 analogs in the assay also give levels that are depressed. The free T4 level determined by dialysis varies widely, as does T4 measured by ultrafiltration25-29; the majority of reports are of low values, but in some samples nnormal or rarely elevated values.25,26,41-43

In theory, methods utilizing equilibrium dialysis may allow dilution of dialyzable inhibitors. Compounds such as 3-carboxy-4-methyl-5-propyl-2-furan-propanoic acid, indoxyl sulfate, and hippuric acid, can accumulate in severe renal failure.44 However, these compounds probably do not interfere with serum hormone assays. Free fatty acids, if elevated to 2 to 5 mmol/L, can displace T4 binding to TBG and elevate free T4. Free fatty acids almost never reach such levels in vivo.45,46 However, even small quantities of heparin (0.08 units/kg given IV, or 5000 units given SC), commonly given to patients in an ICU, can lead to in vitro generation of free fatty acids during extended serum dialysis for “free T4“ assay and falsely augment apparent free hormone levels.47 This is probably a common and serious problem, which explains many instances of apparently elevated free T4 levels in patients with acute illness.

Results obtained using ultrafiltration also are variable. Wang et al.48 found that in patients with NTIS, free T4 measured by ultrafiltration was uniformly low (average of 11.7 ng/L), but when measured by equilibrium dialysis, free T4 was near normal, at 18 ng/L. By ultrafiltration, free T3 was also (not surprisingly) found to be low and similar to free T3 by radioimmune assay. Chopra32 found levels below the normal mean, ±2 SD, when measured by dialysis; 6 of 9 were low when measured by ultrafiltration, and 7of 9 were low when measured by standard resin-uptake-corrected free T4. The means of the NTIS patients in this study were clearly below the mean of normals.

Thus, although free T4 is low in most assays that involve a correction for TBG levels, there is still some question as to the true free T4 in patients with NTIS. It is of interest that this problem does not carry over to estimates of free T3, which are depressed in most studies. There might be two reasons for this difference. Firstly, the depression of total T3 is proportionately greater than of total T4. Secondly, factors which affect thyroid hormone binding are more apt to alter T4 assays than T3, since T4 is normally more tightly bound to TBG than is T3.

 

IS THERE EVIDENCE FOR SUBSTANCES IN SERUM WHICH CAN AFFECT T4 BINDING TO PROTEINS?

Mendel et al.49 carefully review the studies that have claimed the presence of dialyzable inhibitors of binding and point out that many of these studies must be viewed with caution44,45,50-53 .Numerous artifacts are present in both dialysis assays and ultrafiltration assays. They also point out that while the low free T4 by resin uptake assays found in NTIS generally do not agree with the clinical status of the patient, it is equally true that clinical assessment generally does not fit with the high free T4 results found by some equilibrium dialysis assays in NTIS.

An argument that completely refutes the importance of factors in serum inhibiting binding of thyroid hormone is provided in the clinical study of Brent and Hershman (Fig. 2).54 These

Figure 2. Patients with severe NTIS were randomized and left untreated or given T4 iv over two weeks. Serum T3, T4, and TSH concentrations are shown for the survivors of the control filled circles), and T4-treated empty circles), groups during the study period and at the time of follow-up. Upper and lower lines designate the normal range. Note the prompt recovery of T4 values to the normal range immediately following i.v. treatment with T4. Also note the elevated TSH levels in some patients. T3 levels did not return to normal following T4 treatment for up to two weeks. (Reference 54)

Figure 2. Patients with severe NTIS were randomized and left untreated or given T4 iv over two weeks. Serum T3, T4, and TSH concentrations are shown for the survivors of the control filled circles), and T4-treated empty circles), groups during the study period and at the time of follow-up. Upper and lower lines designate the normal range. Note the prompt recovery of T4 values to the normal range immediately following i.v. treatment with T4. Also note the elevated TSH levels in some patients. T3 levels did not return to normal following T4 treatment for up to two weeks. (Reference 54)

researchers gave 1.5 μg of T4 per kg body weight daily to 12 of 24 patients with severe NTIS and followed serum hormone levels over 14 days. T4 levels returned to the normal range within 3 days of therapy. Thus the serum thyroxine pool was easily replenished, and T4 levels reached normal values. Not surprisingly, because of reduced T4>T3 deiodination, T3 levels did not return to the normal range until the end of the study period in the few patients who survived. However, the ability of intravenous thyroxine in replacement doses to promptly restore the plasma pool to normal clearly shows that neither a loss of serum TBG nor an inhibitor of binding could be the main cause of low serum T4 in this group of severely ill patients.

With growing acceptance of decreased thyroid secretion and decreased peripheral t3 production as causes of low T4 and T3, there has been little emphasis on serum T4 binding inhibitors in recent literature. Some contribution by low TBG levels may, or may not (see below) play a role, but any role for binding inhibitors in producing this syndrome must be marginal

TSH LEVELS

Serum TSH in NTIS is typically normal or reduced and may be markedly low, although usually not less than 0.05 μU/mL.16,24,25,28,29,31,55 However, to use usual endocrinology logic, these TSH levels are almost always inappropriately low for the observed serum T4 and T3. Third-generation assays with sensitivity down to 0.001 U/mL may allow differentiation of patients with hyperthyroidism from those with NTIS, although there can be overlap in these very disparate conditions.56 Serum TSH in patients with NTIS may have reduced biological activity, perhaps because of reduced thyrotropin-releasing hormone (TRH) secretion and reduced glycosylation. Some patients are found with a TSH level above normal, and elevation of TSH above normal commonly occurs transiently if patients recover from NTIS (Fig. 3).29,54 This elevation of TSH strongly suggests that the patients are recovering from a hypothyroid state, during which the ability of the pituitary to respond had been temporarily inhibited.

Figure 3. T3 and TSH concentrations are shown in patients with nonthyroidal illness who were eventually discharged from hospital (left panels). The broken line indicates ± 2 SD of the mean value in the normal subjects. The right panel displays T3 and TSH concentrations in patients with NTIS who died. Subjects are indicated by numbers. Note the elevated TSH in some patients who recovered, and the generally dropping T3 and low TSH levels in patients who died. (Reference 29)

Figure 3. T3 and TSH concentrations are shown in patients with nonthyroidal illness who were eventually discharged from hospital (left panels). The broken line indicates ± 2 SD of the mean value in the normal subjects. The right panel displays T3 and TSH concentrations in patients with NTIS who died. Subjects are indicated by numbers. Note the elevated TSH in some patients who recovered, and the generally dropping T3 and low TSH levels in patients who died. (Reference 29)

 

Responsiveness of the pituitary to TRH during NTIS is variable: many patients respond less than normal,57 and others respond normally.58 “Normal” responsiveness in the presence of low TSH may suggest that there is a hypothalamic abnormality as a cause of the low TSH and low T4. There is also a diminution or loss of the diurnal rhythm of TSH,59 and in some studies, there is evidence for reduction of TSH glycosylation, with lower TSH bioactivity.60 A logical explanation is that the low TSH is in fact the proximate cause of the low thyroid hormone levels. Hypothalamic function is impaired in patients with NTIS and TRH mRNA is low,  resulting in low TSH and thus low output of thyroid hormones by the thyroid.

There is other evidence of diminished hypothalamic function in patients with serious illness. Serum testosterone drops rapidly, as do follicle-stimulating hormone (FSH) and luteinizing hormone (LH).61,62 Typically serum cortisol is elevated as part of a stress response, and because metablism of corticol is reduced. Some patients develop hypotension in association with apparent transient central hypoadrenalism, have low or normal serum ACTH, and cortisol levels under 20 μg/dL. Some of these patients respond dramatically to cortisol replacement and may manifest normal adrenal function at a later time if they recover.

Centrally mediated hyposomatotropism, hypothyroidism, and pronounced hypoandrogenism were observed in a study of patients in the catabolic state of critical illness. In these patients, pulsatile LH secretion and mean LH secretions are very low, even in the presence of extremely low circulating total testosterone and low estradiol. Pulsatile growth hormone (GH) and TSH secretion are also suppressed. Interleukin 1 β (IL-1β) levels are normal, whereas IL-6 and tumor necrosis factor α (TNF-α) are elevated. Exogenous IV gonadotropin-releasing hormone (GnRH) partially return serum testosterone levels toward normal but do not completely overcome hypoandrogenism, suggesting that combined deficiency of GH, GnRH, and TSH secretagogues may be important in this low androgen syndrome.63

THYROID HORMONE TURNOVER

Kaptein et al.64,65 studied a group of patients who were critically ill, all of whom had total T4 below 4 μg/dL, low fT4 index, low normal free T4 by dialysis, and TSH which was normal or slightly elevated. In these patients, the mean T4 by dialysis was significantly below the normal mean. There was on average a 35% decrease in thyroxine disposal per day (Table 1). The T4 production rate in NTIS was significantly below the mean of 17 normal subjects (p < 0.005). In a similar study of T3 kinetics,65 free T3 was found to be 50% of normal serum values. The production rate of T3 was reduced by 83% (Table 2). These two studies document a dramatic reduction in provision of T4 and T3 to peripheral tissues, which would logically indicate that the effects of hormone lack (hypothyroidism) should be present. A third study reported dramatically reduced total T4 and T3 turnover, with normal thyroidal secretion of T3 in patients with NTIS due to uremia.66 However, this was a calculated rather than directly measured value of T3 secretion, was highly variable, and does not negate the extreme reduction in T3 supply due to diminished T4 >T3 conversion in peripheral organs.

T4 ENTRY INTO CELLS AND GENERATION OF T3

Thyroid hormone is transported actively into tissues by several specific transporters including MCT8, and in the pituitary OATP1C1. In the cell it is metabolized by enzymes which activate it to T3, or inactivate it to rT3, or promote excretion via sulfation or glucuronidation. Iodotyrosine deiodinase type 1 (ID1) is found in liver, kidney and thyroid, and the enzyme present in liver is considered a main source of T3, possibly providing 80% of the total, the remainder coming largely from the thyroid. ID1 is down-regulated in hypothyroidism, and in NTIS, reducing serum T3 levels.  ID2 is present in brain and pituitary, and is responsible for local production of T3 in those tissues. Recent data show that D2 present in muscle may also contribute to serum T3. ID2 is up-regulated by hypothyroidism, and is up-regulated in  NTIS. The third enzyme, ID3, deiodinates the inner thyronine ring, converting T4 to rT3 and T3 to T2. It’s activity in liver is up-regulated in NTIS.

Using deiodination of T4 as an index of cellular transport of T4 into rat hepatocytes, Lim et al.67 and Vos et al.68 found that serum from patients with NTIS inhibited T4 uptake. Sera from critically ill NTIS patients caused reduced T4 uptake compared to control sera in one study, and the authors considered elevated nonesterified fatty acids (NEFA) and bilirubin and reduced albumin to play a role. Serum from patients with mild NTIS did not cause impaired deiodination of T4 and T3.69 Inhibition of uptake of T4 into hepatocytes caused by sera of patients with NTIS also was observed by Sarne and Refetoff.70 There is a diminution in the “reducing equivalents” available for the deiodination of T4 to T3 in liver, and presumably elsewhere, thus lowering transport and the function of the type 1 iodothyronine deiodinase.71 In animals, and probably in man, there is also a drop in the level of type 1 iodothyronine deiodinase enzyme, apparently due to hypothyroidism, since it can be reversed by giving T3. Recently a study was performed on blood, liver, and skeletal-muscle biopsies of patients immediately after death in intensive care unit settings. Liver T4 deiodinase 1 was found to be down-regulated, and deiodinase 3 was induced in liver and muscle, especially in situations associated with poor tissue perfusion. These changes contribute to the low generation of T3 and its increased metabolism in NTIS, thus lowering the intracellular T3 levels.35

Table 1. T4 Kinetics in the Low T4 State of Nonthyroidal Illness64

Case Number

TT4 (µg/dL)

FT4 (ng/dL)

PR (µg/d/m2)

Normal Subjects (n = 19)

     

Mean

7.1

2.21

50.3

±SE

0.4

0.13

3.4

Sick Patients

     

1

2.7

2.05

32.4

2

3.0

1.23

51.1

3

1.2

0.48

39.0

4

1.4

1.04

23.7

5

1.3

0.75

22.2

6

3.0

1.35

34.6

7

1.9

1.33

36.6

8

2.0

1.88

25.3

9*

0.4

0.28

10.0

10*

1.5

1.50

13.7

11*

1.6

1.70

18.4

Mean

1.8

1.24

27.9

±SE

0.2

0.17

3.7

P

<0.001

<0.001

<0.001

FT4, Free thyroxine; PR, production rate; TT4, total thyroxine.

*Patients receiving dopamine.

All P values are for unpaired t tests.

Table 2. T3 Kinetics in the Low-T4 State of Nonthyroidal Illness65

Case Number

TT3 (ng/dL)

FT3 (pg/dL)

PR (µg/d/m2)

Normal Subjects (n = 12)

     

Mean

162

503

23.47

±SE

5

46

2.12

Sick Patients

     

3

30

272

6.18

5

42

247

5.67

6

25

151

5.41

7

34

266

8.39

12*

45

282

6.07

Mean

35

244

6.34

±SE

4

24

0.53

P

<0.001

<0.001

<0.005

FT3, Free triiodothyronine; PR, production rate; TT3, total triiodothyronine. *Patient receiving dopamine.

In theory, reduced cellular uptake (acting  alone) would cause tissue hypothyroidism, reduced T3 generation and serum T3 levels, and elevated serum T4, which is not observed. It is likely that reduced hormone supply in NTIS is caused by multiple factors, and that reduced cell uptake, if present, is one of the factors. T4 is converted to T3, although at a reduced rate. In addition, T4 is rapidly converted to rT3 by an intracellular process, suggesting that entry into cells is not seriously impaired, but the pathways of intracellular deiodination are abnormal.

 

THYROID HORMONE IN TISSUES

There are increasing data on thyroid hormone in tissues of patients with NTIS.72 In one study, there was of a dramatically reduced level of T3 in tissues (Table 3). While most samples had very low levels of T3 compared to normal tissues, some patients with NTIS showed sporadically and inexplicably high levels of T3 in certain tissues, especially skeletal muscle and heart.

Table 3. Tissue T3 Concentrations in Nonthyroidal Illness Syndrome (nmol of T3/kg of Wet Weight)72

 

Control Group

NTI Group

Tissue

Mean

SD

P

Mean

SD

Cerebral cortex

2.2

0.9

<.05

1.2

1.1

Hypothalamus

3.9

2.2

<.01

1.4

1.2

Anterior pituitary

6.8

2.5

<.005

3.7

1.1

Liver

3.7

2.3

<.01

0.9

0.9

Kidney

12.9

4.3

<.001

3.7

2.8

Lung

1.8

0.8

<.01

0.8

0.5

Skeletal muscle

2.3

1.2

NS

> 10.9

 

Heart

4.5

1.5

NS

> 16.3

 

NS, Not significantly different; NTI, nonthyroidal illness; T3, triiodothyronine.*Patients receiving dopamine.

Peeters et al.73 investigated 79 patients who died after intensive care, some of whom received thyroid hormone treatment. Tissue iodothyronine levels were positively correlated with serum levels, indicating that the decrease in serum T3 during illness is associated with decreased levels of tissue T3. Higher serum T3 levels in patients who received thyroid hormone treatment were accompanied by higher levels of liver and muscle T3, with evidence for tissue-specific regulation. Tissue rT3 and the T3/rT3 ratio were correlated with tissue deiodinase activities. Monocarboxylate transporter 8 expression was not related to the ratio of the serum/tissue concentrations of the different iodothyronines.73

TR LEVELS

Information on expression of TRs in human tissues during illness is limited. Increased expression of the messenger ribonucleic acid (mRNA) for thyroid hormone receptors α1, α2, and β1 has been reported in cardiac tissue of patients with dilated cardiomyopathy; α1 and α2 isoforms also had increased expression in ischemic heart disease.74 Rodriguez-Perez et al. studied subcutaneous fat and skeletal muscle in patients with septic shock.75 In muscle, mRNA for TRβl and RXR gamma was reduced, and mRNA for RXR alpha was increased, compared to normals. In adipose tissue, MCT8, TRβ1, TRα1, and RXR gamma mRNAs were lower. The authors conclude that in these patients, tissue responses were related to decreased hormone levels and decreased hormone action. In animals, starvation and illness are associated with a reduction in thyroid hormone receptor levels. In experimental studies in mice, LPS induces NTIS, and this is associated with an early decrease in binding of the RXR/TR dimer to DNA due to limiting amounts of RXR, and later an up to 50% decrease in levels of RXR and TR protein.- Lado-Abeal  and co workers found in humans with prolonged NTIS that expression of TRbeta1, TRalpha1, and RXRgamma in striated muscle were reduced compared to normals, and that these changes were unrelated to expression of NFkB1(Figure 4)(78). .

Figure 4 . THRA, THRB1, RXRG, RXRA, AND PPARG PROTEIN levels were evaluated by western blotting.

Figure 4 . THRA, THRB1, RXRG, RXRA, AND PPARG PROTEIN levels were evaluated by western blotting.

 

 

 

ORGAN SPECIFIC RESPONSES IN NTIS

In contrast to a uniform whole body response, there are wide variations in responses to thyroid hormone supply and action in different tissues, and between the response to acute illness and chronic illness. These multifactorial systems involve serum TH levels, TH transporters, deiodinases, TH receptors, and enzyme responses which are under different regulation in individual tissues (79). In reviewing these data it is useful to note that meaning of data often must be “interpreted”, and that this can depend on the mind-set of the reviewer. In the hypothalamus in an animal model of NTIS, TH transporters MCT10 and OATP1C1 (but not MCT8) were increased, and hypothalamic D2, considered a major source of local T3, was up-regulated, but there was no corresponding increase in tissue T3..  TRalpha and beta mRNA expression  levels were not altered (Figure 4)(80). Tissue levels of T3 and T4 are reduced in chronic NTIS in both experimental models and in humans. Thus the know low TRH mRNA levels in the hypothalamus reflect the action of neural signals, and not an hypothesized local tissue hyperthyroidism. On the other hand, the “low normal” T3 levels in the hypothalamus, in the presence of low serum T3 levels, presumably reflect the actionof D2, and may provide a partial explanation of why TSH is not usually elevated in NTIS. Data on the pituitary are few. D2 levels have not shown consistent changes. In animal models TRbeta2 levels are reported to be reduced in acute NTIS. In humans with fatal illness, pituitary T3 levels were low. The overall picture is of central down-regulation of the hypothalamus and pituitary and low levels of tissue T3 despite increments in transporter activity and D2 deiodinase(Figure 5). These findings fit with the observed correction of TSH and TH levels in human NTIS through administration of TRH.

Figure 5 . T3 and T4 content in Hypothalamus of chronically ill animals with NTIS.(From Mebis et al(80).

Figure 5 . T3 and T4 content in Hypothalamus of chronically ill animals with NTIS.(From Mebis et al(80).

Liver D1 and D3 activity are reduced in acute NTIS. In man MCT8 and MCT10 may be reduced in acute but not chronic NTIS. Several enzymes that are responsive to TH have reduced activity in acute NTIS in animals. The data fit with reduced metabolic activity in this organ. In chronic  NTIS in man, reduced  serum T3 and T4 and normal or elevated RT3  are characteristic. Liver T3 is low in chronic human NTIS, and is directly related to serum T3. MCT8, but not MCTt10, are increased in liver and muscle in prolonged NTIS (81). Expression levels of TRalpha and beta mRNA are reported to be reduced, or increased. Metabolic activity is probably reduced, but in relation to oxygenation and nutrients as well as TH activity. In a rabbit model of chronic NTIS, serum and liver T3  and liver D1 activity were low. Interestingly replacement with basal levels T4 or T3 did not reverse these abnormalities, but 3-5 fold increments, or TRH administration, did so (82). This study offers several interesting points. There was a very strong correlation between serum  and tissue hormone levels, so tissue entry was snot a problem. The requirement for more than replacement dose T4 to restore tissue hormone is mainly due to repletion of very diminished stores of T4, in this 6 day treatment  protocol. No evidence was found for an important role of T4-sulfo  conjugates. The response in muscle is less well defined. D2 increases and D3 is decreased in acute NTIS in an animal model, and there is evidence for decreased TR expression. Changes in enzyme responses do not show a consistent pattern.  In chronic NTIS, human muscle D3 is augmented,  while D2 has been found low, or increased. Muscle MCT8 is increased (81), and this has been proposed as a response to hypothyroidism, since TH treatment in a rabbit model of NTIS returns the transporters to normal. Enzymatic activity is presumed low but good data are lacking.  Prolonged infusion of lipopolysaccharide in pigs induced a severe NTIS state, associated with generally low tissue TH levels, reduced TH transporters, and low TR-beta levels, suggesting reduced TH sensitivity and hypothyroidism(83).

 

ARE PATIENTS WITH NTIS CLINICALLY HYPOTHYROID?

It is straightforward that the typical clinical parameters of severe hypothyroidism are absent in patients with NTIS. However, these patients usually present with a serious illness and are diagnostically challenging in view of their complicated state. Many are febrile, have extensive edema, have sepsis or pneumonia, may have hyper metabolism associated with burns, have severe cardiac or pulmonary disease, and in general have features that could easily mask evidence of hypothyroidism. Further, the common clinical picture of hypothyroidism does not develop within 2 to 3 weeks of complete thyroid hormone deprivation, but rather requires a much longer period for expression. General laboratory tests are also suspect. Thus starvation or disease-induced alterations in cholesterol, liver enzymes, TBG, creatine kinase, and even basal metabolic rate generally rule out the use of these associated markers for evidence of hypothyroidism. Angiotensin-converting enzyme levels are low,84 as seen in hypothyroidism, while high-affinity testosterone-binding globulin (TeBG) and osteocalcin levels are not altered.85Antithrombin III levels are reduced in a septic rat model of NTIS. T3 supplementation returned the sepsis-induced decrease in antithrombin III levels toward normal.86

 

MECHANISM OF THYROID HORMONE SUPPRESSION IN NTIS

It is probable that the cause of NTIS is multifactorial and may differ in different groups of patients. Specifically, the changes in liver disease and renal disease are probably somewhat different from those occurring in other forms of illness. Certainly one important cause of the drop in serum T3 is a decreased generation of T3 by type 1 iodothyronine deiodinase.87 Reduced entry of T4 into cells is not a major problem. Some studies have suggested that individuals with NTIS may have selenium deficiency, and this may contribute to a malfunction of the selenium-dependent iodothyronine deiodinase.However, supplements of 500 μg of selenium given to patients in a surgical ICU during the first 5 days after serious injury caused only modest changes in thyroid hormones. The data did not suggest a major role for selenium deficiency in this condition.88

The overall daily metabolic consumption of thyroid hormone, both thyroxine and T3, is radically diminished in the NTIS syndrome in the presence of low hormone serum levels. The reduced degradation cannot produce the lowering of serum hormone levels; a primary reduction in degradation would increase serum hormone. The change in degradation must be due to the low hormone supply, and other factors. Schussler and co-workers have observed a sharp drop in TBG levels during cardiac bypass surgery, which their studies indicate is due to some selective consumption of TBG. It is possible that this occurs because of activation of serine protease inhibitors (serpins) at sites of inflammation, which cleave the TBG into an inactive form.38

Considerable evidence suggests that an alteration in hypothalamic and pituitary function causes the low production of T4, which in turn causes the low production of T3. In rats, starvation reduces hypothalamic mRNA for TRH, reduces portal serum TRH, and lowers pituitary TSH content.89A recent study documents low TRH mRNA in hypothalamic paraventricular nuclei90 in NTIS patients (Fig. 6). Responses to administered TRH vary in different reports, being suppressed or even augmented.57,58 Administration of TRH has been suggested as an effective

 Figure 6. In situ hybridization study demonstrating mRNA for TRH in the periventricular nuclei of a subject who died with NTIS in Panel A, and a subject who died accidentally in Panel B. mRNA for TRH is significantly reduced in patients with NTIS. (Reference 90)

Figure 6. In situ hybridization study demonstrating mRNA for TRH in the periventricular nuclei of a subject who died with NTIS in Panel A, and a subject who died accidentally in Panel B. mRNA for TRH is significantly reduced in patients with NTIS. (Reference 90)

means of restoring serum hormone levels to normal in individuals with NTIS. A recent report by Van den Berghe and co-workers proves that administration of TRH to patients with severe NTIS leads directly to increased TSH levels, increased T4 levels, and increased T3 levels.91 This data is strong support (albeit not proof) for the role of diminished hypothalamic function as a crucial factor in NTIS.

Quite possibly the production of TRH, and responses to TRH, are reduced by cytokines (to be discussed later) or by glucocorticoids.92 The diurnal variation in glucocorticoid levels at least in part controls the normal diurnal variation in TSH levels, perhaps by affecting pituitary responsiveness to TRH.93 High levels of glucocorticoids in Cushing’s disease suppress TSH and cause a modest reduction in serum hormone levels.94 High levels of glucocorticoids are known to suppress pituitary response to TRH in man.92Stress-related elevation of glucocorticoids in animals causes suppression of TSH and serum T4 and T3 hormone levels.95 Thus stress-induced glucocorticoid elevation may be one factor affecting TRH and TSH production.

Why should pituitary production of TSH be diminished in the presence of low serum thyroid hormone levels? One idea was that augmented intrapituitary conversion of T4 to T3 allowed the pituitary to remain suppressed while the rest of the body was actually hypothyroid. While some data supported  this idea in a uremic rat model of NTIS96, careful studies in both experimental  animals(82,83) and man, described above, disprove this concept.

 Another suggestion is that some other metabolite of thyroxine may be involved in control of pituitary responsiveness. For example, possibly triiodothyroacetic acid (triac) or tetraiodothyroacetic acid (tetrac) generated by metabolism of thyroxine could control pituitary responsiveness,92 but there is no experimental proof of this idea, and even if true, it would mean that the pituitary was normal but the rest of the body hypothyroid. As suggested earlier, elevated serum cortisol levels could play a role. The most obvious possibility is that low TSH stems from diminished TRH production, as previously described. It must also be remembered that the defect in pituitary function is not restricted to TSH, but that LH and FSH are also suppressed in seriously ill patients, and testosterone is reduced, in contrast to the generally augmented glucocorticoid levels. Quite possibly these changes are the effect on the hypothalamus of neural integration of multiple factors including stress, starvation, glucocorticoids, and cytokines.

Van den Berghe has stressed that the changes in endocrine function seen during severe illness have a biphasic course. Possibly the initial suppression of T3 levels represents a genetically engineered adaptive response of the organism, allowing reduced metabolic rate and conservation of energy and protein stores for a longer period of time, while the animal or man goes through a period of starvation. However, the circumstances surrounding severe illness, and the resuscitative efforts applied in an intensive care unit over 1 or more weeks, seem to be a different reponse. This second phase of the syndrome, with associated suppression of thyroid hormone and other pituitary hormones and a variety of other changes, may represents a maladaptive response. Patients in this situation tend to have elevated insulin levels, nitrogen wasting, retention of fats if calories are made available, and a variety of other metabolic abnormalities that include neuropathy and cardiomyopathy. These authors consider that provision of multiple hormonal support, including thyroid hormone, growth hormone, and androgens, may be beneficial.97,98

 

CYTOKINES IN NTIS

In a series of septic patients studied shortly after admission to an ICU, total T4, free T4, total T3, and TSH were depressed, and IL-1β, soluble interleukin-2 receptor (sIL-2R), IL-6, and TNF-α were elevated.99 The hypothalamo-pituitary-adrenal axis was activated as expected. The data suggest central suppression of TSH as the cause of the problem, but the relation to cytokines is unclear, as seen in the following reports. Hermus et al.100 showed that continuous infusion of IL-1 in rats causes suppression of TSH, T3, and free T4. Higher doses of IL-1 were accompanied by a febrile reaction and suppression of food intake, which presumably played some role in the altered thyroid hormone economy. IL-1 did not reproduce the diminution in hepatic 5′-deiodinase activity believed to be so characteristic of NTIS. IL-1 is also known to impair thyroid hormone synthesis by human thyrocytes and is enhanced in many diseases associated with NTIS.101 van der Poll et al.102 studied the effect of IL-1 receptor blockade in human volunteers to determine if it could alter the NTIS induced by endotoxin. Blockade of IL-1 activity was achieved by infusing recombinant human IL-1 receptor antagonist, but this did not prevent the drop in T4, free T4, T3, and TSH or the rise in rT3 caused by endotoxin. This is evidence against an important role for IL-1.

Interferon γ

Interferon-γ (IFN-γ) 100 μg/m2 administered subcutaneously to normal volunteers did not alter TNF-α levels, caused a small elevation of IL-6 levels, and thus did not support a role for IFN-γ in the pathogenesis of the euthyroid sick syndrome in humans.103

Tumor Necrosis Factor

TNF is another proinflammatory cytokine that is thought to be involved in many of the illnesses associated with NTIS.Infusion of recombinant TNF in man by van der Poll et al.104 produced a decrease in serum T3 and TSH and an increase in rT3. Free T4 was transiently elevated in association with a significant rise in FFA levels. These studies suggest that TNF could be involved in producing NTIS. Recombinant IL-6 given to humans activates the hypothalamic pituitary-adrenal axis, and this could play a role in suppressing TSH production. However, Chopra et al.105 did not find TNF to be closely correlated with hormone changes in NTIS. van der Poll et al.1063 gave human subjects endotoxin, which caused lowering of T4, free T4, T3, and TSH. TNF blockade by a recombinant TNF receptor-IgG fusion protein did not alter the response, indicating that TNF did not cause the changes in hormone economy induced by administration of endotoxin. Nagaya et al.107 proposed a mechanism through which TNF could reduce serum T3. TNF-α was found during in vitro studies to activate nuclear factor kB (NF-kB), which in turn inhibits the T3-induced expression of 5′-DI, which would lower T3 generation in liver. However, as noted above, activation of liver NFkB is not seen in NTIS.

Interleukin 6

Serum IL-6 is often elevated in NTIS,108 and its level is inversely related to T3 levels. Stouthard et al.109 gave recombinant human IL-6 chronically to human volunteers. Short-term infusion of IL-6 caused a suppression of TSH, but daily injections over 42 days caused only a modest decrease in T3 and a transient increase in rT3 and free T4 concentrations. IL-6 could be involved in the NTIS syndrome, although the mechanism was not defined. In an animal model of NTIS studied by Wiersinga and collaborators,110 antibody blockade of IL-6 failed to prevent the induced changes in thyroid hormone economy typical of NTIS. Boelen et al. studied the levels of IFN, IL-8, and IL-10 in patients with NTIS and found no evidence that they had a pathogenic role.111 Short-term administration of recombinant IFN-γ to normal subjects caused a minimal elevation of IL-6, no alteration in TNF, and did not significantly alter thyroid hormone levels. Michalaki et al. observed that serum T3 drops early after abdominal surgery as an early manifestation of the NTIS syndrome, prior to an increase in serum IL-6 or TNF-α, suggesting that these changes in cytokines do not induce the drop in T3.113

The potential interaction between cytokines and the hypothalamic-pituitary-thyroid axis is certainly complicated, and cytokines themselves operate in a network. For example, IL-1 and TNF can stimulate secretion of IL-6. Activation of TNF and IL-1 production is associated with the occurrence of cytokine inhibitors in serum, which are actually fragments of the cytokine receptor or actual receptor antagonists. Soluble TNF receptor and IL-1 RA are receptor antagonists, which can inhibit the function of the free cytokines. These molecules are increased in many infectious, inflammatory, and neoplastic conditions. Boelen et al.113 found evidence that the NTIS is “an acute phase response” generated by activation of a cytokine network. Soluble TNF, soluble TNF receptor, soluble IL-2 receptor antagonist, and IL-6 all inversely correlated with serum T3 levels.

While the studies noted fail to pinpoint one cytokine as the crucial mediator, we can be convinced that striking changes in cytokines co-occur during NTIS and probably play a pathogenic role by mechanisms yet undefined.

 

OTHER FACTORS ALTERING SERUM T4 SUPPLY

Altered CNS Metabolism

In healthy men going through two 4.5-hour-long sessions of induced hypoglycemia, TSH, fT3 and fT4 are significantly reduced.114 Perinatal asphyxia, recognized by low Apgar scores, is associated with a depression of TSH, T4, and T3, and the reductions are greatest in infants with hypoxic/ischemic encephalopathy. In this study, 6 of 11 infants with fT4 < 2ng/dL died. These data suggest, not surprisingly, that reduced substrate or O2 supply to the CNS could induce hypothalamic/pituitary dysfunction.114,115

Glucagon

Administration of glucagon to dogs caused a significant fall in serum T3, suggesting that stress-induced hyperglucagonemia may be a contributor to the NTIS syndrome by altering intracellular metabolism of T4.116

Dopamine

Dopamine given in support of renal function and cardiac function must play a role in many patients who develop low hormone levels while in an intensive care unit setting. Dopamine inhibits TSH secretion directly, depresses further the already abnormal thyroid hormone production, and induces significant worsening of the low hormone levels. Withdrawal of dopamine infusion is followed by a prompt dramatic elevation of TSH, a rise in T4 and T3, and an increase of the T3/rT3 ratio. All of these changes suggested to Van den Berghe et al.117 that dopamine makes some patients with NTIS hypothyroid, inducing a condition of iatrogenic hypothyroidism, and that treatment (presumably by administering thyroid hormone), “should be evaluated.”

Leptin

Leptin plays a key role in control of thyroid hormone levels during starvation in animals. During starvation, leptin levels drop. With this there is diminished stimulation of TRH, thus diminished secretion of TSH, and lowered thyroid hormone levels. Administration of leptin appears to work via the arcuate nucleus of the hypothalamus to induce production of pro-opiomelanocortin (POMC), and thus α-melanocyte-stimulating hormone (αMSH), and reduce Agouti-related protein (AgRP). Normally αMSH stimulates the melanocortin 4 receptor (MC4R), whereas AgRP suppresses it. Presumably through these actions, a lack of leptin during starvation leads to diminished stimulation of the MC4R receptor on the TRH neurons in ventricular nuclear centers and thus diminished TRH secretion. Administration of leptin partially reverses this sequence.118 These actions appear to be part of an energy-conserving scheme related to thyroid changes during starvation and are associated with leptin-induced increase in appetite, decreased energy expenditure, and modified neuroendocrine function. The relevance of this to human physiology is as yet unclear, but the data strongly suggest that leptin is involved in the down-regulation of thyroid function during acute starvation.118-120 In clinical trials, stimulation of growth hormone secretion by GH secretogogues lead to increased insulin and leptin levels in severely ill ICU patients. To date, studies of leptin levels in patients with NTIS have indicated they are normal or elevated, not low.121

Atrial Natriuretic Peptides

Atrial natriuretic peptides, including amino acids 1 to 30, amino acids 31 to 67 (known as vessel dilator), 79 to 98 (kaliuretic hormone), and 99 to 126 (atrial natriuretic hormone), derived from the ANH prohormone, significantly decreased circulating concentrations of total T4, free T4, and free T3, when given to healthy humans for 60 minutes. A reciprocal increase in TSH lasted for 2 or 3 hours after cessation of the administration of these hormones, suggesting that the effect was a direct inhibition of thyroid hormone release from the thyroid gland rather than an action of the hormones upon the hypothalamus or pituitary. No data are available on these factors in NTIS122 (Table 4).

 

DIAGNOSIS

Typically the endocrinologist is presented with a severely ill patient in whom there is no prior history suggestive of pituitary disease, in whom clinical findings of hypothyroidism are either absent or masked by other disorders, with a T4 and FTI (by an index method) that are low, a low or normal TSH, and, if measured, a low T3. If T4 is below 4 μg/dL in this setting, the diagnosis of NTIS, associated with a potentially fatal outcome, may be assumed; rT4 may be normal or elevated and is not diagnostic. An elevated TSH suggests the presence of prior hypothyroidism, which should be treated. Finding positive antithyroid antibody titers supports the diagnosis of primary hypothyroidism but does not prove it.

Serum cortisol should be measured. Transient, apparently central, hypoadrenalism may occur in severe illness.123-125 Cortisol should be above 20 μg/dL, and commonly is above 30. If below 20, ACTH should be drawn, and the patient may be given supportive cortisol therapy. Serum cortisol should certainly be determined if thyroid hormone is to be given. Since CBG may be reduced, it is advisable to measure serum free cortisol if possible. It is useful to determine FSH in postmenopausal women as a sign of pituitary function, but this is less clearly valuable in men. If there is a reason to consider hypopituitarism, a CAT scan of the pituitary is appropriate, or at least a skull film.

Use of aspirin, dilantin, and carbamazepine should be noted, since they can lower T4 and FTI as measured by several “index” methods. Dopamine used in the setting of severe illness can induce clear-cut hypothyroidism. Hyperthyroidism is the typical cause of suppression of TSH below 0.1 μU/mL, but it is rarely difficult to exclude this diagnosis in the setting of severely depressed T4 and T3.

 

IS THYROID HORMONE TREATMENT OF NTIS ADVANTAGEOUS OR DISADVANTAGEOUS?

Two valuable studies are available on replacement therapy using thyroid hormone in patients with NTIS. In the study by Brent and Hershman,54 replacement with 1.5 μg T4 IV per kilogram body weight daily, in 12 patients, promptly returned serum T4 levels to normal (thereby proving that a binding defect was not the cause of the low T4) but did not normalize T3 levels over a period of 2 to 3 weeks. However, in both the treated and control group, mortality was 80%.54 Clearly this excellent small study, which used for primary therapy what would now be considered the wrong hormone, failed to show either an advantageous or disadvantageous effect. It is possible that the failure to show a positive effect was due to the failure of T3 levels to be restored to normal. In a study of severely burned patients given 200 μg T3 daily, again there was no evidence of a beneficial or disadvantageous effect.126 Mortality was not so great, as in the Brent and Hershman study, but it is entirely possible that the high levels of T3 given worsened the hypermetabolism known to be present in burn patients and could have, at these levels, been disadvantageous.

An important study by Acker et al. certainly advises caution regarding T4 therapy in patients with acute renal failure. Numerous studies in animals have documented a beneficial effect of T4 therapy in experimental acute renal failure.127 In a randomized, controlled prospective study of patients with acute renal failure (ARF), treated patients received 150 μg of thyroxine a total of four times intravenously over 2 days.128 The single difference recognized in the subsequent laboratory data was a suppression of TSH. T4 treatment had no effect on any measure of ARF severity. Among other questions, it is not clear that serum T3 levels were ever altered. However, mortality was higher in the thyroxine group (43% versus 13%) than in the control group. It is of interest that, as the authors state, “the observed mortality in the controls in this study was less than that typically seen in our institution in ARF and ICU patients, whereas the 43% mortality noted in the thyroid group better approximates both our experience and that reported in the literature for ICU patients.” It will be difficult to replicate this study (although this reader believes it should be replicated). But it is uncertain whether the small dose of thyroxine administered over 2 days actually is related to the mortality, considering that the mortality in the treated group was that usually observed, whereas the control happened to have a much lower mortality.128 The same group has also studied the effect of thyroid hormone treatment on posttransplant acute tubular necrosis. T3 treatment during the posttransplant period did not alter outcome in a beneficial or derogatory manner.129

Studies from animals are often quoted in the literature as an argument against treatment of NTIS or for the therapy. A study of sepsis induced in animals showed no difference in mortality with treatment, but some animals treated with thyroid hormone died earlier than did those that were untreated.130 Chopra et al. induced NTIS in rats by injection of turpentine oil. The reduction in T4, T3, free T4 index, and TSH were associated with no clear evidence of tissue hypothyroidism, and urinary nitrogen excretion was normal. Thyroid hormone replacement with T4 or T3 did not significantly alter enzyme activities or urinary nitrogen excretion.131 Healthy pigs were subjected to 20 minutes of regional myocardial ischemia by Hsu and collaborators,132 and this was associated with a drop in T3, free T3, and elevated rT3. Some animals were treated with 0.2 μg T3 per kilogram for five doses over 2 hours. While myocardial infarction size was not altered, the pigs treated with T3 showed a more rapid improvement in cardiac index. Oxygen consumption did not alter. It should be noted that the T3 levels fell back to normal levels within 4 hours of the last T3 dose, suggesting that more prolonged therapy might have been beneficial. Katzeff et al.133 studied a model of NTIS induced by caloric restriction in young rats. In these animals, T3 was reduced, and there was a decrease in LV relaxation time, SERCA2 mRNA, and αMHC mRNA. All changes were reversed to normal values by supplementation with T3, suggesting that the low-T3 syndrome was related to the pathologic cardiac changes. Sepsis and multisystem organ failure are often associated with disseminated intravascular coagulation and consumption of coag inhibitors such as antithrombin III. Chapital studied a model of sepsis in rats and showed that T3 supplementation reduced the decrease in antithrombin III levels, which presumably would reflect a beneficial effect.134 Dogs subjected to hemorrhagic shock recovered more cardiovascular function when given T3 intravenously than did untreated animals.135 Neurologic outcome after anoxia is improved in dogs by T3 treatment.136

Short-term studies on T3 replacement of patients in shock, in patients with respiratory disease, in subjects who are brain dead and potential organ donors, and in patients undergoing coronary artery bypass grafts all suggest modest cardiovascular benefits from the administration of T3. One study reports benefit by replacing T3 to elevate the depressed T3 levels in premature infants.137 Other studies found no apparent effects. Children treated with T3 postoperatively when they have undergone cardiac surgery also require less cardiac support(138). T3 administration (one dose of approximately 6 μg IV) did not alter cardiac performance in brain-dead transplant donors.139 Coronary artery bypass, as studied by Klemperer and collaborators,36 was associated with a drop in serum T3; IV administration of T3 elevated T3 above normal, augmented cardiac output, and reduced the need for pressor support but had no other effect. In this study, however, the patients had a very favorable prognosis and minimal NTIS, so the study primarily shows that administration of T3 had no adverse effect under these circumstances. In a study of patients after heart transplant, patients with the low T3 syndrome (NTIS) had higher mortality, higher incidence of acute rejection, highest number of re-operations, and higher incidence of infections, compared to those without NTIS (140). In a study reported several years ago, T3 administration to critically ill neonates with severe respiratory distress appeared to improve survival. Infants of less than 37 weeks gestational age or weighing less than 220 grams were given prophylactic doses of thyroxine and T3 daily and had a lower mortality rate than untreated infants.137 Use of thyroid hormone replacement in children after cardiac surgery has been extensively reviewed by Haas et al., with the conclusion that it is a desirable treatment option, especially in high-risk patients.141 Goarin et al. studied the effect of T3 administration in brain-dead organ donors and found that although it returned T3 levels to normal, it did not improve hemodynamic status or myocardial function.142 Pingitore et al. gave T3 by IV infusion for 3 days to patients with chronic heart failure. Heart rate, plasma nor-epinephrine (down 52%), natriuretic peptide, and aldosterone (down 23%) were all significantly diminished, and ventricular performance improved, without side effects.143 In a randomized study of patients for 24 hours after coronary bypass, correction of the usual drop in serum T3 by IV T3 infusion had no beneficial or deleterious effect on cardiac parameters.144 Of interest, it also did not affect leucine flux or urinary nitrogen excretion, contrary to the usual assumption that a drop in serum T3 should spare body protein. Novitsky (145) studied three conditions in which NTIS and myocardial functional depression have been documented - i) transient regional myocardial ischemia and reperfusion, ii) transient global myocardial ischemia in patients undergoing cardiac surgery on cardiopulmonary bypass, and iii) transient inadequate global myocardial perfusion in brain-dead potential organ donors. Under all three conditions, in models and in man, following administration of T3/T4, the myocardial dysfunction was rapidly reversed. Cautiously use of thyroid hormonal therapy to any patient with the ESS and/or a stunned myocardium was advocated. The general outcome of these studies is that they weakly support the use of T3, and none of the studies found evidence of damage caused by treatment.144-150

In summary, it can be stated that there is no clear evidence that thyroxine or triiodothyronine treatment of NTIS in animals or man is disadvantageous, but no certain proof that it is advantageous.  In the acute NTIS syndrome associated with operations, short term treatment with T3 augments cardiac function, but has not been shown to alter the already low mortality (150). However, what evidence there is suggests TH may be beneficial. The argument has been raised that administration of thyroid hormone in NTIS would prevent the elevation in TSH commonly seen in recovering patients. This seems rather specious. More objectively, the elevation of TSH is another suggestion that the few patients who survive the ordeal were hypothyroid and left untreated. Lastly, it is unlikely that administration of replacement hormone during NTIS would be harmful, even if all of the evidence presented suggesting hypothyroidism was erroneous, and the patients were in fact euthyroid.

 

IF THYROID HORMONE REPLACEMENT IS GIVEN, WHAT SHOULD IT BE?

Clearly the high mortality rate in patients in the chronic phase of NTIS, with T4 under 4 μg/dL, suggests that this is a target group in whom thyroid hormone administration should be considered. In this group of patients, there appears to be no obvious contraindication to replacement therapy, with the possible exception of people who have cardiac decompensation or arrhythmias. Even here, the evidence is uncertain. There is no clear evidence that administration of replacement doses of T3 to patients with low cardiac output is disadvantageous, and in fact current studies using intravenous T3 in these patients indicate it is well tolerated and may be beneficial.151 Arrhythmias obviously also raise a question, but again, there is no evidence that replacement of thyroid hormone to a normal level would cause trouble in control of arrhythmias. Low free T3 levels are reported to be associated with an increased incidence of fibrillation after cardiac surgery in elderly patients.152 Thus even in this group of patients, it is reasonable to suggest therapy. It should also be noted that among patients with NTIS, there will certainly be patients who are clearly hypothyroid—based on known disease, treatment with dopamine, or elevated TSH—who need replacement therapy by any standard.

If therapy is to be given, it cannot be thyroxine alone, since this would fail to promptly elevate T3 levels.54 Treatment should include oral, or if this is impractical, intravenous T3, and probably should be at the replacement level of approximately 50 μg/day given in divided doses. It may be appropriate to give slightly higher doses, such as 75 μg/day for 3 to 4 days to increase the body pool more rapidly, followed by replacement doses as described. Coincidentally, it is appropriate to start replacement with T4. Serum levels of T4 and T3 should be followed at frequent intervals (every 48 hours) and dosages adjusted to achieve a serum T3 level at least low normal (70 to 100 ng/dL) prior to the next scheduled dose. If treatment is successful, T3 administration can gradually be reduced, and thyroxine administration can be increased to replacement levels as deiodination increases. Because of the marked diminution in T4 to T3 deiodination, and shunting of T4 toward r T3, replacement with T4 may initially only lead to elevation of rT3 and have very little effect upon T3 levels, or physiologic action. In this situation, continued administration of T3 would be preferred. An alternate therapy, giving TRH to stimulate TSH production and TH release, has been shown to be effective in raising TH levels during short term treatment. This is described below. No prolonged treatment, or effects on survival have so far been reported.

 

ADDITIONAL SUPPORTIVE HORMONAL THERAPY TO CONSIDER

Although this discussion concentrates on the potential value of treating patients with NTIS with replacement thyroid hormone, several important recent studies expand the concept to other areas, including treatment of the associated hyperglycemia, relative adrenal insufficiency, use of beta blockers in burn patients, and possible use of GHRH and testosterone. Van den Berghe and co-workers have suggested that the acute and prolonged critical illness responses are entirely different neuroendocrine conditions. In protracted severe illness, patients are kept alive with conditions that previously caused death. However, this process has unmasked a variety of nonspecific wasting syndromes that include protein loss, accumulation of fat stores, hyperglycemia and insulin resistance, hypoproteinemia, hypercalcemia, potassium depletion, and hypertriglyceridemia. In prolonged illness, cortisol values are elevated, although ACTH levels are low, indicating that other mechanisms are driving the steroid response. Growth hormone secretory pulses are reduced, and the mean GH concentration is low in prolonged critical illness. FSH and LH are reduced, and testosterone levels are reduced. These authors maintain that the reduced neuroendocrine drive, present in the chronic phase of illness in an intensive care setting, is unlikely to be an evolutionary preserved beneficial process. They suggest that the administration of hypothalamic physiotropic releasing peptides may be a safer strategy than the administration of peripherally active hormones.153-156

NTIS is typically associated with poor nutrition unless supportive measures are taken, and undernutrition is a known inducer of NTIS. Yet nutritional support is not uncomplicated. Parenteral feeding of rabbits (in contrast to fasting) in a model of chronic NTIS resulted in a normalization of low T3 levels, but did not correct low T4 levels(157) .Use of early parenteral alimentation for patients in the ICU with NTIS was recommended by Perez-Guisado et al (158), who found it decreased length of hospital stay.  However a study by Langouche et al (159) found, in contrast, that adding parenteral nutrition sooner than one week in the ICU increased complications and delayed recovery.

  Hyperglycemia and insulin resistance are common in critically ill patients, even if they have not previously had diabetes. Van den Berghe et al carried out a prospective randomized study on ICU patients on mechanical ventilation, maintaining blood glucose at a level between 80 and 110 mg/dl, versus allowing glucose to range between a level of 180 – 200 mg/dl. Intensive insulin therapy reduced overall in-hospital mortality by 34 percent, bloodstream infections by 46 percent, acute renal failure requiring dialysis or hemofiltration by 41 percent, the median number of red-cell transfusions by 50 percent, and critical-illness polyneuropathy by 44 percent, and patients receiving intensive therapy were less likely to require prolonged mechanical ventilation and intensive care (98). This strict blood sugar control can lead to bouts of hypoglycemia, and higher controlled values are currently aimed for. In isolated brain injury patients, intensive insulin therapy reduced mean and maximal intracranial pressure while identical cerebral perfusion pressures were obtained with eightfold less vasopressors. Seizures and diabetes insipidus occurred less frequently. At 12 months follow-up, more brain-injured survivors in the intensive insulin group were able to care for most of their own needs. Preventing even moderate hyperglycemia with insulin during intensive care protected the central and peripheral nervous systems, with clinical consequences such as shortening of intensive care dependency and possibly better long-term rehabilitation. Prevention of catabolism, acidosis, excessive inflammation, and impaired innate immune function may explain previously documented beneficial effects of intensive insulin therapy on outcome of critical illness.   Severe burns are known to be associated with a hypermetabolic state and a strong sympathetic response. Beta blockade given as propranolol to reduce the resting heart rate by 20% decreased resting energy expenditure and increased net muscle protein balance  significantly in a group of burn patients. It is logical that this would be a significant benefit (160). Severe sepsis, which is of course associated with NTIS, is frequently associated with relative adrenal insufficiency, and possibly systemic inflammation-induced glucocorticoid receptor resistance. In a prospective randomized study, Annane et al studied a seven day treatment of patients with septic shock, by giving hydrocortisone, 50 mg q6h, and 9-alpha-fludrocortisone, 50 mg once daily. The risk of death in this treated group was significantly reduced without increasing any adverse effects. The treatment was most beneficial in individuals who responded poorly to a 250 mg ACTH test, which was conducted prior to the therapy.  Non-response was defined as a response of 9 mg/dl or less, between the lowest, and highest concentration taken after the ACTH injection. Samples were taken in this study at 30 and 60 minutes (161). The severity of the illness was suggested by the statistics that 63% died in the placebo group, and 53% in the corticosteroid treatment group. The authors recommend that all patients with catecholamine dependent septic shock should be given a combination of hydrocortisone and fludrocortisone as soon as a short corticotrophin stimulation test is performed, and the treatment should be continued for seven days in non-responders. Hamrahian et al advise caution in using total serum cortisol measurements in patients with serum albumin levels below 21.5gm/dl. They observed that these patients may have low total cortisol because of low CBG, but have normal or elevated free cortisol levels (162) In contrast to the generally beneficial effects of hormonal therapy described above, high levels of growth hormone given to critically ill patients were found by Takala et al to augment mortality. The dosage used was 0.1 mg/kg bw, for up to 21 days. Mortality rate was nearly double. These authors suggest that GH may have an adverse effect upon immunity, cause fluid retention, and cause hyperglycemia (163).

TREATMENT WITH HYPOTHALAMIC RELEASING HORMONES

Van Den Berghe and collaborators have pioneered studies on the effects of hypothalamic releasing hormones in patients with severe NTIS. The logic supporting this approach is that it corrects a major cause of the low hormonal state, and may allow normal feed-back control and peripheral regulation of hormones, thus being more physiological than replacing the peripheral hormone deficit directly. Extensive studies document restoration of T4 and T3 levels following administration of TRH and GH secretagaugue (153). In a rabbit model of NTIS treatment with GHRP-2 and TRH reactivated the GH and TSH axes and altered liver deiodinase activity, driving T4 to T3 conversion. In NTIS there are suppressed pulsatile GH, TSH, LH secretion in the face of low serum concentrations of IGF-I, IGFBP-3 and the acid-labile subunit (ALS), thyroid hormones, and total and estimated free testosterone levels, whereas free estradiol (E2) estimates are normal. Ureagenesis and breakdown of bone tissue are increased. Baseline serum TNF-alpha, IL-6 and C-reactive protein level and white blood cell (WBC) count are elevated; serum lactate is normal. Coadministration of GHRP-2, TRH and GnRH reactivated the GH, TSH and LH axes in prolonged critically ill men and evoked beneficial metabolic effects which were absent with GHRP-2 infusion alone and only partially present with GHRP-2 + TRH. These data underline the importance of correcting the multiple hormonal deficits in patients with prolonged critical illness to counteract the hypercatabolic state (154. Contrary to expectation, intensive insulin therapy suppressed serum IGF-I, IGFBP-3, and acid-labile subunit concentrations. This effect was independent of survival of the critically ill patient. Concomitantly, serum GH levels were increased by intensive insulin therapy. The data suggest that intensive insulin therapy surprisingly suppressed the somatotropic axis despite its beneficial effects on patient outcome. GH resistance accompanied this suppression of the IGF-I axis. To what extent and through which mechanisms the changes in the GH-IGF-IGFBP axis contributed to the survival benefit under intensive insulin therapy remain elusive (155). While outcome studies using this approach are not available, it is quite possible that treatment of NTIS by use of hypothalamic releasing hormones may be a preferred approach.

CONCLUSIONS

This review has presented the arguments for administration of replacement T3 and T4 hormone in patients with NTIS. However, it is impossible to be certain at this time that it is beneficial to replace hormone, or whether this could be harmful. Other recent reviews on this topic are available (164). Only a prospective study will be adequate to prove or disprove the value of hormone replacement, and probably this would need to involve hundreds of patients. Tragically, many ICU patients continuing to die with NTIS (we do not know if this is from NTIS) and we have now waited over 40 years for the proper controlled study to be done. One cannot envisage that replacement of thyroxin or T3 can “cure” patients with NTIS. The probable effect, if any is achieved, will be a modest increment in overall physiologic function and a decrease in mortality. Perhaps this would be 5%, 10%, or 20%. If effective, thyroid hormone replacement will be one of many beneficial treatments given the patient, rather than a single magic bullet which would reverse all the metabolic changes going wrong in these severely ill patients. Ongoing studies document the beneficial effects of hormone replacement in these acutely and severely ill patients. Possibly therapy will ultimately involve replacement of peripheral hormones, or may instead be via growth hormone-releasing peptide (GHRP), TRH, GnRH, insulin, adrenal steroids, and leptin.

  REFERENCES

1.         McIver B, Gorman CA: Euthyroid sick syndrome: An overview, Thyroid 7:125–132, 1997.

2.         DeGroot LJ: “Non-thyroidal illness syndrome” is functional central hypothyroidism, and if severe, hormone replacement is appropriate in light of present knowledge, J Endocrinol Invest 26:1163–1170, 2003.

3.         Stathatos N, Wartofsky L: The euthyroid sick syndrome: is there a physiologic rationale for thyroid hormone treatment? J Endocrinol Invest 26:1174–1179, 2003.

4.         Hennemann G, Docter R, Krenning EP: Causes and effects of the low T3 syndrome during caloric deprivation and non-thyroidal illness: an overview, Acta Med Kaust 15:42–45, 1988.

5.         Phillips RH, Valente WA, Caplan ES, et al: Circulating thyroid hormone changes in acute trauma: prognostic implications for clinical outcome, J Trauma 24:116–119, 1984.

6.         Vardarli I, Schmidt R, Wdowinski JM, et al: The hypothalamo-hypophyseal-thyroid axis, plasma protein concentrations and the hypophyseogonadal axis in low T3 syndrome following acute myocardial infarct, Klin Wochenschrift 65:129–133, 1987.

7.         Eber B, Schumacher M, Langsteger W, et al: Changes in thyroid hormone parameters after acute myocardial infarction, Cardiology 86:152–156, 1995.

8.         Holland FW, Brown PS, Weintraub BD, et al: Cardiopulmonary bypass and thyroid function: a “euthyroid sick syndrome.” Ann Thorac Surg 52:46–50, 1991.

9.         Vexiau P, Perez-Castiglioni P, Socie G, et al: The “euthyroid sick syndrome”: Incidence, risk factors and prognostic value soon after allogeneic bone marrow transplantation, Br J Hematol 85:778–782, 1993.

10.       Harris ARC, Fang SL, Vagenakis AG, et al: Effect of starvation, nutriment replacement, and hypothyroidism on in vitro hepatic T4 to T3 conversion in the rat, Metabolism 27:1680–1690, 1978.

11.       Welle SL, Campbell RG: Decrease in resting metabolic rate during rapid weight loss is reversed by low-dose thyroid hormone treatment, Metabolism 35:289–291, 1986.

12.       Plikat K, Langgartner J, Buettner R, et al: Frequency and outcome of patients with nonthyroidal illness syndrome in a medical intensive care unit, Metabolism 56(2):239–244, 2007 Feb.

13.       Girvent M, Maestro S, Hernandez R, et al: Euthyroid sick syndrome, associated endocrine abnormalities, and outcome in elderly patients undergoing emergency operation, Surgery 123:560–567, 1998.

14.       Wartofsky L, Burman KD: Alterations in thyroid function in patients with systemic illnesses: the “Euthyroid Sick Syndrome”, Endocrine Rev 3:164–217, 1982.

15.       Kaptein EM: Clinical relevance of thyroid hormone alterations in nonthyroidal illness, Thyroid International 4:22–25, 1997.

16.       Docter R, Krenning EP, de Jong M, et al: The sick euthyroid syndrome: changes in thyroid hormone serum parameters and hormone metabolism, Clin Endocrinol 39:499–518, 1993.

17.       Chopra IJ, Huang TS, Boado R, et al: Evidence against benefit from replacement doses of thyroid hormones in nonthyroidal illness: studies using turpentine oil–injected rat, J Endocrinol Invest 10:559–564, 1987.

18.       Schilling JU, Zimmermann T, Albrecht S, et al: Low T3 syndrome in multiple trauma patients – a phenomenon or important pathogenetic factor? Medizinische Klinik 3:66–69, 1999.

19.       Schulte C, Reinhardt W, Beelen D, et al: Low T3-syndrome and nutritional status as prognostic factors in patients undergoing bone marrow transplantation, Bone Marrow Transplantation 22:1171–1178, 1998.

20.       Girvent M, Maestro S, Hernandez R, et al: Euthyroid sick syndrome, associated endocrine abnormalities, and outcome in elderly patients undergoing emergency operation, Surgery 123:560–567, 1998.

21.       Maldonado LS, Murata GH, Hershman JM, et al: Do thyroid function tests independently predict survival in the critically ill? Thyroid 2:119, 1992.

22.       Vaughan GM, Mason AD, McManus WF, et al: Alterations of mental status and thyroid hormones after thermal injury, J Clin Endocrinol Metab 60:1221, 1985.

23.       De Marinis L, Mancini A, Masala R, et al: Evaluation of pituitary-thyroid axis response to acute myocardial infarction, J Endocrinol Invest 8:507, 1985.

24.       Surks MI, Hupart KH, Pan C, et al: Normal free thyroxine in critical nonthyroidal illnesses measured by ultrafiltration of undiluted serum and equilibrium dialysis, J Clin Endocrinol Metab 67:1031–1039, 1988.

25.       Melmed S, Geola FL, Reed AW, et al: A comparison of methods for assessing thyroid function in nonthyroidal illness, J Clin Endocrinol Metab 54:300–306, 1982.

26.       Kaptein EM, MacIntyre SS, Weiner JM, et al: Free thyroxine estimates in nonthyroidal illness: comparison of eight methods, J Clin Endocrinol Metab 52:1073–1077, 1981.

27.       Kantor M, et al: Admission thyroid evaluation in very-low-birth-weight infants: association with death and severe intraventricular hemorrhage, Thyroid 13:965, 2003.

28.       Chopra IJ, Solomon DH, Hepner GW, et al: Misleadingly low free thyroxine index and usefulness of reverse triiodothyronine measurement in nonthyroidal illnesses, Ann Int Med 90:905–912, 1979.

29.       Bacci V, Schussler GC, Kaplan TB: The relationship between serum triiodothyronine and thyrotropin during systemic illness, J Clin Endocrinol Metab 54:1229–1235, 1982.

30.       Sapin R, Schlienger JL, Kaltenbach G, et al: Determination of free triiodothyronine by six different methods in patients with nonthyroidal illness and in patients treated with amiodarone, Ann Clin Biochem 32:314–324, 1995.

31.       Chopra IJ, Taing P, Mikus L: Direct determination of free triiodothyronine (T3) in undiluted serum by equilibrium dialysis/radioimmunoassay, Thyroid 6:255–259, 1996.

32.       Chopra IJ: Simultaneous measurement of free thyroxine and free 3,5,3’-triiodothyronine in undiluted serum by direct equilibrium dialysis/radioimmunoassay: evidence that free triiodothyronine and free thyroxine are normal in many patients with the low triiodothyronine syndrome, 1998.

33.       Faber J, Kirkegaard C, Rasmussen B, et al: Pituitary-thyroid axis in critical illness, J Clin Endocrinol Metab 65(2):315–320, 1987 Aug.

34.       Fritz KS, Wilcox RB, Nelson JC: Quantifying spurious free T4 results attributable to thyroxine-binding proteins in serum dialysates and ultrafiltrates, Clin Chem 53(5):985–988, 2007 May.

35.       Peeters RP, Wouters PJ, Kaptein E, et al: Reduced activation and increased inactivation of thyroid hormone in tissues of critically ill patients, J Clin Endocrinol Metab 88(7):3202–3211, 2003 Jul.

36.       Klemperer JD, Klein I, Gomez M, et al: Thyroid hormone treatment after coronary-artery bypass surgery, N Engl J Med 333:1522–1527, 1995.

37.       Osburne RC, Myers EA, Rodbard D, et al: Adaptation to hypocaloric feeding: physiologic significance of the fall in T3, Metabolism 32:9–13, 1983.

38.       Afandi B, Schussler GC, Arafeh A-H, et al: Selective consumption of thyroxine-binding globulin during cardiac bypass surgery, Metabolism 49:270–274, 2000.

39.       Jirasakuldech B, Schussler GC, Yap MG, et al: A characteristic serpin cleavage product of thyroxine-binding globulin appears in sepsis sera, J Clin Endocrinol Metab 85:3996–3999, 2000.

40.       den Brinker M, Joosten KF, Visser TJ, et al: Euthyroid sick syndrome in meningococcal sepsis: the impact of peripheral thyroid hormone metabolism and binding proteins, J Clin Endocrinol Metab 90(10):5613–5620, 2005.

41.       Uchimura H, Nagataki S, Tabuchi T, et al: Measurements of free thyroxine: comparison of percent of free thyroxine in diluted and undiluted sera, J Clin Endocrinol Metab 42:561–566, 1976.

42.       Nelson JC, Weiss RM: The effect of serum dilution on free thyroxine concentration in the low T4 syndrome of nonthyroidal illness, J Clin Endocrinol Metab 61:239–246, 1985.

43.       Wang R, Nelson JC, Weiss RM, et al: Accuracy of free thyroxine measurements across natural ranges of thyroxine binding to serum proteins, Thyroid 10:31–39, 2000.

44.       Kaptein EM: Thyroid hormone metabolism and thyroid diseases in chronic renal failure, Endocrine Rev 17:45–63, 1996.

45.       Liewendahl K, Helenius T, Naveri H, et al: Fatty acid-induced increase in serum dialyzable free thyroxine after physical exercise: implication for nonthyroidal illness, J Clin Endocrinol Metab 74:1361–1365, 1992.

46.       Mendel CM, Frost PH, Cavalieri RR: Effect of free fatty acids on the concentration of free thyroxine in human serum: the role of albumin, J Clin Endocrinol Metab 63:1394–1399, 1986.

47.       Jaume JC, Mendel CM, Frost PH, et al: Extremely low doses of heparin release lipase activity into the plasma and can thereby cause artifactual elevations in the

48.       Wang Y-S, Hershman JM, Pekary AE: Improved ultrafiltration method for simultaneous measurement of free thyroxine and free triiodothyronine in serum, Clin Chem 31:517–522, 1985.

49.       Mendel CM, Laufhton CW, McMahon FA, et al: Inability to detect an inhibitor of thyroxine-serum protein binding in sera from patients with nonthyroidal illness, Metabolism 40:491–502, 1991.

50.       Chopra IJ, Huang T-S, Beredo A, et al: Evidence for an inhibitor of extrathyroidal conversion of thyroxine to 3,5,3’-triiodothyronine in sera of patients with nonthyroidal illnesses, J Clin Endocrinol Metab 60:666, 1985.

51.       Wang R, Nelson JC, Wilcox RB: Salsalate administration—a potential pharmacological model of the sick euthyroid syndrome, J Clin Endocrinol Metab 83:3095–3099, 1998.

52.       Csako G, Zweig MH, Benson C, et al: On the albumin-dependence of the measurement of free thyroxine. II. Patients with nonthyroidal illness, Clin Chem 33:87–92, 1987.

53.       Chopra IJ, Chua Teco GN, Mead JF, et al: Relationship between serum free fatty acids and thyroid hormone binding inhibitor in nonthyroid illnesses, J Clin Endocrinol Metab 60:980–984, 1985.

54.       Brent GA, Hershman JM: Thyroxine therapy in patients with severe nonthyroidal illnesses and lower serum thyroxine concentration, J Clin Endocrinol Metab 63:1–8, 1986.

55.       Chopra IJ: Nonthyroidal illness syndrome or euthyroid sick syndrome? Endocrine Pract 2:45–52, 1996.

56.       Franklyn JA, Black EG, Betteridge J, et al: Comparison of second and third generation methods for measurement of serum thyrotropin in patients with overt hyperthyroidism, patients receiving thyroxine therapy, and those with nonthyroidal illness, J Clin Endocrinol Metab 78:1368–1371, 1994.

57.       Vierhapper H, Laggner A, Waldhausl W, et al: Impaired secretion of TSH in critically ill patients with “low T4-syndrome”. Acta Endocrinol 101:542–549, 1982.

58.       Faber J, Kirkegaard C, Rasmussen B, et al: Pituitary-thyroid axis in critical illness, J Clin Endocrinol Metab 65:315–320, 1987.

59.       Arem R, Deppe S: Fatal nonthyroidal illness may impair nocturnal thyrotropin levels, Am J Med 88:258–262, 1990.

60.       Lee H-Y, Suhl J, Pekary AE, et al: Secretion of thyrotropin with reduced concanavalin-A-binding activity in patients with severe nonthyroid illness, J Clin Endocrinol Metab 65:942, 1987.

61.       Spratt DI, Bigos ST, Beitins I, et al: Both hyper- and hypogonadotropic hypogonadism occur transiently in acute illness: bio- and immunoactive gonadotropins, J Clin Endocrinol Metab 75:1562–1570, 1992.

62.       Spratt DI, Cox P, Orav J, et al: Reproductive axis suppression in acute illness is related to disease severity, J Clin Endocrinol Metab 76:1548–1554, 1993.

63.       Van den Berghe G, Weekers F, Baxter RC, et al: Five-day pulsatile gonadotropin-releasing hormone administration unveils combined hypothalamic-pituitary-gonadal defects underlying profound hypoandrogenism in men with prolonged critical illness, J Clin Endocrinol Metab 86:3217–3226, 2001.

64.       Kaptein EM, Grieb DA, Spencer CA, et al: Thyroxine metabolism in the low thyroxine state of critical non-thyroidal illnesses, J Clin Endocrinol Metab 53:764–771, 1981.

65.       Kaptein EM, Robinson WJ, Grieb DA, et al: Peripheral serum thyroxine, triiodothyronine and reverse triiodothyronine kinetics in the low thyroxine state of acute nonthyroidal illnesses. A noncompartmental analysis, J Clin Invest 69:526–535, 1982.

66.       Lim VS, Fang VS, Katz AI, et al: Thyroid dysfunction in chronic renal failure. A study of the pituitary-thyroid axis and peripheral turnover, Kinetics of thyroxine and triiodothyronine, J Clin Invest 60(3):522–534, 1997.

67.       Lim C-F, Docter R, Visser TJ, et al: Inhibition of thyroxine transport into cultured rat hepatocytes by serum of nonuremic critically ill patients: effects of bilirubin and nonesterified fatty acids, J Clin Endocrinol Metab 76:1165–1172, 1993.

68.       Vos RA, de Jong M, Bernard BF, et al: Impaired thyroxine and 3,5,3’-triiodothyronine handling by rat hepatocytes in the presence of serum of patients with nonthyroidal illness, J Clin Endocrinol Metab 80:2364–2370, 1995.

69.       Lim C-F, Docter R, Krenning EP, et al: Transport of thyroxine into cultured hepatocytes: effects of mild nonthyroidal illness and calorie restriction in obese subjects, Clinical Endocrinol 40:79–85, 1994.

70.       Sarne DH, Refetoff S: Measurement of thyroxine uptake from serum by cultured human hepatocytes as an index of thyroid status: reduced thyroxine uptake from serum of patients with nonthyroidal illness, J Clin Endocrinol Metab 61:1046–1052, 1985.

71.       Krenning D, et al: Decreased transport of thyroxine (T4), 3, 33′, 5-triiodothyronine (T3) and 3,33′,5′-triiodothyronine (rT3) into rat hepatocytes in primary culture due to a decrease of cellular ATP content and various drugs, FEBS Lett 140:229–233, 1982.

72.       Arem R, Wiener GJ, Kaplan SG, et al: Reduced tissue thyroid hormone levels in fatal illness, Metabolism 42:1102–1108, 1993.

73.       Peeters RP, van der Geyten S, Wouters PJ, et al: Tissue thyroid hormone levels in critical illness, J Clin Endocrinol Metab 90(12):6498–6507. Epub 2005 Sep 20, 2005.

74.       d’Amati G, di Gioia CRT, Mentuccia D, et al: Increased expression of thyroid hormone receptor isoforms in end-stage human congestive heart failure, J Clin Endocrinol Metab 86:2080–2084, 2001.

75.       Rodriguez-Perez A, Palos-Paz F, Kaptein E, et al: Identification of molecular mechanisms related to nonthyroidal illness syndrome in skeletal muscle and adipose tissue from patients with septic shock, Clin Endocrinol (Oxf) 68(5):821–827, 2008.

76.       Sanchez B, Jolin T: Triiodothyronine-receptor complex in rat brain: effects of thyroidectomy, fasting, food restriction, and diabetes, Endocrinology 129:361–367, 1991.

77.       Beigneux A, et al: Sick euthyroid syndrome is associated with decreased TR expression and DNA binding in mouse liver, Am J Physiol Endocrinol Metab 284:E228-E236, 2003.

78        -Lado-Abeal J, Romero A, Castro-Piedras I, Rodriguez-Perez A, Alvarez-Escudero J     Thyroid hormone receptors are down-regulated in skeletal muscle of patients with non-thyroidal illness syndrome secondary to non-septic shock. Eur J Endocrinol. 2010 Nov;163(5):765-73. 79        -Boelen A, Kwakkel J, Fliers E  Beyond low plasma T3: local thyroid hormone metabolism during inflammation and infection.Endocr Rev. 2011 Oct;32(5):670-93.

80.       .Mebis L, Debaveye Y, Ellger B, Derde S, Ververs EJ, Langouche L, Darras VM, Fliers E, Visser TJ, Van den Berghe G. Changes in the central component of the hypothalamus-pituitary-thyroid axis in a rabbit model of prolonged critical illness.Crit Care. 2009;13(5):R147

81.       Mebis L, Paletta D, Debaveye Y, Ellger B, Langouche L, D'Hoore A, Darras VM, Visser TJ, Van den Berghe G. Expression of thyroid hormone transporters during critical illness. Eur J Endocrinol. 2009 Aug;161(2):243-50.

82.       Debaveye Y, Ellger B, Mebis L, Visser TJ, Darras VM, Van den Berghe G.--Effects of substitution and high-dose thyroid hormone therapy on deiodination, sulfoconjugation, and tissue thyroid hormone levels in prolonged critically ill rabbits.Endocrinology. 2008 Aug;149(8):4218-28. .

83.       Castro I, Quisenberry L, Calvo RM, Obregon MJ, Lado-Abeal J. Septic shock non-thyroidal illness syndrome causes hypothyroidism and conditions for reduced sensitivity to thyroid hormone. J Mol Endocrinol. 2013 Mar 18;50(2):255-66.

84.       Brent GA, Hershman JM, Reed AW, et al: Serum angiotensin converting enzyme in severe nonthyroidal illness associated with low serum thyroxine concentration, Ann Intern Med 100:680–683, 1986.

84.       Seppel T, Becker A, Lippert F, et al: Serum sex hormone-binding globulin and osteocalcin in systemic nonthyroidal illness associated with low thyroid hormone concentrations, J Clin Endocrinol Metab 81:1663–1665, 1996.

86.       Chapital AD, Hendrick SR, Lloyd L, et al: The effects of triiodothyronine augmentation on antithrombin III levels in sepsis, Am Surg 67(3):253–255, 2001 Mar.

87.       Kaplan MM: Subcellular alterations causing reduced hepatic thyroxine-53-monodeiodinase activity in fasted rats, Endocrinology 104:58–64, 1979.

88.       Berger MM, Reymond MJ, Shenkin A, et al: Influence of selenium supplements on the post-traumatic alterations of the thyroid axis: a placebo-controlled trial, Intensive Care Med 27:91–100, 2001.

89.       Blake NG, Eckland JA, Foster OJF, et al: Inhibition of hypothalamic thyrotropin-releasing hormone messenger ribonucleic acid during food deprivation, Endocrinology 129:2714–2718, 1991.

90.       Fliers E, Guldenaar SEF, Wiersinga WM, et al: Decreased hypothalamic thyrotropin-releasing hormone gene expression in patients with nonthyroidal illness, J Clin Endocrinol Metab 82:4032–4036, 1997.

91.       Van den Berghe G, De Zegher F, Baxter RC, et al: Neuroendocrinology of prolonged critical illness: effects of exogenous thyrotropin-releasing hormone and its combination with growth hormone secretagogues, J Clin Endocrinol Metab 83:309–319, 1998.

92.       Nicoloff JT, Fisher DA, Appleman MD Jr: The role of glucocorticoids in the regulation of thyroid function in man, J Clin Invest 49:1922, 1970.

93.       Brabant G, Brabant A, Ranft U: Circadian and pulsatile thyrotropin secretion in euthyroid man under the influence of thyroid hormone and glucocorticoid administration, J Clin Endocrinol Metab 65:83, 1987.

94.       Benker G, Raida M, Olbricht T, et al: TSH secretion in Cushing’s syndrome: relation to glucocorticoid excess, diabetes, goiter, and the “sick euthyroid syndrome”, Clin Endocrinol 33:777–786, 1990.

95.       Bianco AC, Nunes MT, Hell NS, et al: The role of glucocorticoids in the stress-induced reduction of extrathyroidal 3,5,3′-triiodothyronine generation in rats, Endocrinology 120:1033–1038, 1987.

96.       Lim VS, Passo C, Murata Y, et al: Reduced triiodothyronine content in liver but not pituitary of the uremic rat model: demonstration of changes compatible with thyroid hormone deficiency in liver only, Endocrinology 114:280–286, 1984.

97.       Van den Berghe G, de Zegher F, Bouillon R: Acute and prolonged critical illness as different neuroendocrine paradigms, J Clin Endocrinol Metab 83:1827–1834, 1998.

94.       Van den Berghe G, et al: Intensive insulin therapy in critically ill patients, N Engl J Med 345:1359–1367, 2001.

98.       Van den Berghe G, Schoonheydt K, Becx P, et al: Insulin therapy protects the central and peripheral nervous system of intensive care patients, Neurology 64(8):1348–1353, 2005 Apr 26.

99.       Mönig H, Arendt T, Meyer M, et al: Activation of the hypothalamo-pituitary-adrenal axis in response to septic or non-septic diseases—implications for the euthyroid sick syndrome, Intensive Care Med 25:1402–1406, 1999.

100.     Hermus RM, Sweep CG, van der Meer MJ, et al: Continuous infusion of interleukin-1 beta induces a non-thyroidal illness syndrome in the rat, Endocrinology 131:2139–2146, 1992.

101.     Cannon JG, Tompkins RG, Gelfand JA, et al: Circulating interleukin-1 and tumor necrosis factor in septic shock and experimental endotoxin fever, J Infect Dis 161:79–84, 1990.

102.     van der Poll T, Van Zee KJ, Endert E, et al: Interleukin-1 receptor blockade does not affect endotoxin-induced changes in plasma thyroid hormone and thyrotropin concentrations in man, J Clin Endocrinol Metab 80:1341–1346, 1995.

103.     de Metz J, Romijn JA, Endert E, et al: Administration of interferon-γ in healthy subjects does not modulate thyroid hormone metabolism, Thyroid 10:87–91, 2000.

104.     van der Poll T, Romijn JA, Wiersinga WM, et al: Tumor necrosis factor: a putative mediator of the sick euthyroid syndrome in man, J Clin Endocrinol Metab 71:1567–1572, 1990

105.     Chopra IJ, Sakane S, Chua Teco GN: A study of the serum concentration of tumor necrosis factor—in thyroidal and nonthyroidal illnesses, J Clin Endocrinol Metab 72:1113–1116, 1991.

106.     van der Poll T, Endert E, Coyle SM, et al: Neutralization of TNF does not influence endotoxin-induced changes in thyroid hormone metabolism in humans, Am J Physiol 276:R357-R362, 1999.

107.     Nagaya T, Fujieda M, Otsuka G, et al: A potential role of activated NF-kappa B in the pathogenesis of euthyroid sick syndrome, J Clin Invest 106(3):393–402, 2000.

108.     Bartalena L, Brogioni S, Grasso L, et al: Relationship of the increased serum interleukin-6 concentration to changes of thyroid function in nonthyroidal illness, J Endocrinol Invest 17:269–274, 1994.

109.     Stouthard JML, van der Poll T, Endert E, et al: Effects of acute and chronic interleukin 6 administration on thyroid hormone metabolism in humans, J Clin Endocrinol Metab 79:1342–1346, 1994.

110.    Boelen A1, Platvoet-ter Schiphorst MC, Wiersinga WM Immunoneutralization of interleukin-1, tumor necrosis factor, interleukin-6 or interferon does not prevent the LPS-induced sick euthyroid syndrome in mice. J Endocrinol. 1997 Apr;153(1):115-22.

111.     Boelen A, Platvoet-ter Schiphorst MC, Wiersinga WM: Relationship between serum 3,5,3′-triiodothyronine and serum interleukin 8, interleukin 10 or interferon gamma in patients with nonthyroidal illness, J Endocrinol Invest 19:480–483, 1996.

112.     Michalaki M, Vagenakis AG, Makri M, et al: Dissociation of the early decline in serum T3 concentration and serum IL-6 rise and TNFα in nonthyroidal illnss syndrome induced by abdominal surgery, J Clin Endocrinol Metab 86:4198–4205, 2001.

113.     Boelen A, Platvoet-ter Schiphorst MC, Wiersinga WM: Soluble cytokine receptors and the low 3,5,3′-triiodothyronine syndrome in patients with nonthyroidal disease, J Clin Endocrinol Metab 80:971–976, 1995.

114.     Schultes B, et al: Acute and prolonged effects of insulin-induced hypoglycemia on the pituitary-thyroid axis in humans, Metabolism 51:1370–1374, 2002.

115.     Pereira D, Procianoy R: Effect of perinatal asphyxia on thyroid-stimulating hormone and thyroid hormone levels, Acta Pediatr 92:339–345, 2003.

116.     Custro N, Scafidi V, Costanzo G, et al: Role of high blood glucagon in the reduction of serum levels of triiodothyronine in severe nonthyroid diseases, Minerva Endocrinol 14:221–226, 1989.

117.     Van den Berghe G, de Zegher F, Lauwers P: Dopamine and the sick euthyroid syndrome in critical illness, Clin Endocrinol 41:731–737, 1994.

118.     Legradi G, Emerson CH, Ahima RS, et al: Leptin prevents fasting-induced suppression of prothyrotropin-releasing hormone messenger ribonucleic acid in neurons of the hypothalamic paraventricular nucleus, Endocrinology 138:2569–2576, 1997.

119.     Legradi G, Emerson CH, Ahima RS, et al: Arcuate nucleus ablation prevents fasting-induced suppression of proTRH mRNA in the hypothalamic paraventricular nucleus, Neuroendocrinology 68:89–97, 1998.

120.     Flier JS, Harris M, Hollenberg AN: Leptin, nutrition, and the thyroid: the why, the wherefore, and the wiring, J Clin Invest 105:859–861, 2000.

121.     Bornstein SR, et al: Leptin levels are elevated despite low thyroid hormone levels in the “euthyroid sick” syndrome, J Clin Endocrinol Metab 82(12):4278–4279, 1997.

122.     Vesely DL, San Miguel GI, Hassan I, et al: Atrial natriuretic hormone, vessel dilator, long-acting natriuretic hormone, and kaliuretic hormone decrease the circulating concentrations of total and free T4 and free T3 with reciprocal increase in TSH, J Clin Endocrinol Metab 86:5438–5442, 2001.

123.     Kidess AI, Caplan RH, Reynertson RH, et al: Transient corticotropin deficiency in critical illness, Mayo Clin Proc 68:435–441, 1993.

124.     Merry WH, Caplan RH, Wickus GG, et al: Acute adrenal failure due to transient corticotropin deficiency in postoperative patients, Surgery 116:1095–1100, 1994.

125.     Lambert SWJ, Bruining HA, DeLong FH: Corticosteroid therapy in severe illness, N Engl J Med 337:1285–1292, 1997.

126.     Becker RA, Vaughan GM, Ziegler MG, et al: Hyper-metabolic low triiodothyronine syndrome of burn injury, Critical Care Med 10:870–875, 1982.

127.     Siegel N, et al: Beneficial effect of thyroxine on recovery from toxic acute renal failure, Kidney Int 25:906–911, 1984.

128.     Acker CG, Singh AR, Flick RP, et al: A trial of thyroxine in acute renal failure, Kidney Int 57:293–298, 2000.

129.     Acker CG, Flick R, Shapiro R, et al: Thyroid hormone in the treatment of post-transplant acute tubular necrosis (ATN), Am J Transplant 2:57–61, 2002.

130.     Little JS: Effect of thyroid hormone on survival after bacterial infection, Endocrinology 117:1431–1435, 1985.

131.     Chopra IJ, Huang TS, Boado R, et al: Evidence against benefit from replacement doses of thyroid hormones in nonthyroidal illness: studies using turpentine oil-injected rat, J Endocrinol Invest 10:559, 1987.

132.     Hsu R-B, Huang T-S, Chen Y-S, et al: Effect of triiodothyronine administration in experimental myocardial injury, J Endocrinol Invest 18:702–709, 1995.

133.     Katzeff HL, Powell SR, Ojamaa K: Alterations in cardiac contractility and gene expression during low T3 syndrome: prevention with T3, Amer J Physiol 273:E951–E956, 1997.

134.     Chapital AD, Hendrick SR, Lloyd L, et al: The effects of triiodothyronine augmentation on antithrombin III levels in sepsis, Am Surg 67:253–255, 2001.

135.     Shigematsu H, Shatney CH: The effect of triiodothyronine and reverse triiodothyronine on canine hemorrhagic shock, Nippon Geka Gakkai Zasshi 89:1587–1593, 1988.

136.     Facktor MA, Mayor GH, Nachreiner RF, et al: Thyroid hormone loss and replacement during resuscitation from cardiac arrest in dogs, Resuscitation 26:141–162, 1993.

137.     Schoenberger W, Grimm W, Emmrich P, et al: Thyroid administration lowers mortality in premature infants, Lancet 2:1181, 1979.

138      Choi YS, Kwak YL, Kim JC, Chun DH, Hong SW, Shim JK. Peri-operative oral triiodothyronine replacement therapy to prevent postoperative low triiodothyronine state following valvular heart surgery. Anaesthesia. 2009 Aug;64(8):871-7.

139.     Novitzky D, et al: Improved cardiac allograft function following triiodothyronine therapy to both donor and recipient, Transplantation 49:311–316, 1990.

140.     Kowalczuk-Wieteska A, Baranska-Kosakowska A, Zakliczynski M, Lindon S, Zembala M. Do thyroid disorders affect the postoperative course of patients in the early post-heart transplant period? Ann Transplant. 2011 Jul-Sep;16(3):77-81.

141.     Haas NA, Camphausen CK, Kececioglu D: Clinical review: thyroid hormone replacement in children after cardiac surgery: is it worth a try? Crit Care 10(3):213, 2006.

142.     Goarin JP, Cohen S, Riou B, et al: The effects of triiodothyronine on hemodynamic status and cardiac function in potential heart donors, Anesth Analg 83:41–47, 1996.

143.     Pingitore A, Galli E, Barison A, et al: Acute effects of triiodothyronine (T3) replacement therapy in patients with chronic heart failure and low-T3 syndrome: a randomized, placebo-controlled study, J Clin Endocrinol Metab 93(4):1351–1358, 2008.

144.     Spratt DI, Frohnauer M, Cyr-Alves H, et al: Physiological effects of nonthyroidal illness syndrome in patients after cardiac surgery, Am J Physiol Endocrinol Metab 293(1):E310-E315, 2007.

145.      Novitzky D1, Cooper DK2. Thyroid hormone and the stunned myocardium. J Endocrinol.        2014 Oct;223(1):R1-8. doi: 10.1530/JOE-14-0389.

146.     Hesch R, et al: Treatment of dopamine-dependent shock with triiodothyronine, Endocr Res Commun 8:299–301, 1981.

147.     Dulchavsky S, et al: T3 preserves respiratory function in sepsis, J Trauma 31:753–759, 1991.

148.     Dulchavsky S, Hendrick S, Dutta S: Pulmonary biophysical effects of triiodothyronine (T3) augmentation during sepsis-induced hypothyroidism, Trauma 35: 104–109, 1993.

149.     Bennett-Guerro E, et al: Duke T3 Study Group, Cardiovascular effects of intravenous triiodothyronine in patients undergoing coronary artery bypass graft surgery. A randomized, double-blind, placebo-controlled trial, J Am Med Assoc 275:687–692, 1996.

150      Kaptein EM, Sanchez A, Beale E, Chan LS. Clinical review: Thyroid hormone therapy for postoperative nonthyroidal illnesses: a systematic review and synthesis. J Clin Endocrinol Metab. 2010 Oct;95(10):4526-34.

151.     Hamilton MA, Stevenson LW: Thyroid hormone abnormalities in heart failure: possibilities for therapy, Thyroid 6:527–529, 1996.

152.     Kokkonen L, Majahalme S, Kööbi T, et al: Atrial fibrillation in elderly patients after cardiac surgery: postoperative hemodynamics and low postoperative serum triiodothyronine, J Cardiothorac Vasc Anesth 19(2):182–187, 2005 Apr.

153.       Weekers F, Michalaki M, Coopmans W, Van Herck E, Veldhuis JD, Darras VM, Van den Berghe G Endocrine and metabolic effects of growth hormone (GH) compared with GH-releasing peptide, thyrotropin-releasing hormone, and insulin infusion in a rabbit model of prolonged critical illness.Endocrinology. 2004 Jan;145(1):205-13.

154.       Van den Berghe G, Baxter RC, Weekers F, Wouters P, Bowers CY, Iranmanesh A,Veldhuis JD, Bouillon R: Clin Endocrinol (Oxf). 2002 May;56(5):655-69. The combined administration of GH-releasing peptide-2 (GHRP-2), TRH and GnRH tomen with prolonged critical illness evokes superior endocrine and metabolic effects compared to treatment with GHRP-2 alone

155. .      Mesotten D, Wouters PJ, Peeters RP, Hardman KV, Holly JM, Baxter RC, Van den Berghe G.: J Clin Endocrinol Metab. 2004 Jul;89(7):3105-13. Regulation of the somatotropic axis by intensive insulin therapy during protracted critical illness.

156.        Van den Berghe G, Weekers F, Baxter RC, Wouters P, Iranmanesh A, Bouillon R, Veldhuis JD. Five-day pulsatile gonadotropin-releasing hormone administration unveils combined hypothalamic-pituitary-gonadal defects underlying profound hypoandrogenism in men with prolonged critical illness. J Clin Endocrinol Metab 86:3217-3226, 2001

157.      Mebis L, Eerdekens A, Güiza F, Princen L, Derde S, Vanwijngaerden YM, Vanhorebeek I, Darras VM, Van den Berghe G, Langouche L Contribution of nutritional deficit to the pathogenesis of the nonthyroidal illness syndrome in critical illness: a rabbit model study. Endocrinology. 2012 Feb;153(2):973-84.

158.    Pérez-Guisado J, de Haro-Padilla JM, Rioja LF, Derosier LC, de la Torre JI.The potential  association of later initiation of oral/enteral nutrition on euthyroid sick syndrome in burn patients.Int J Endocrinol. 2013;2013:707360

159.      Langouche L, Vander Perre S, Marques M, Boelen A, Wouters PJ, Casaer MP, Van den Berghe G. Impact of early nutrient restriction during critical illness on the nonthyroidal illness syndrome and its relation with outcome: a randomized, controlled clinical study. J Clin Endocrinol Metab. 2013 Mar;98(3):1006-13.

160.      Herndon DN, Hart DW, Wolf SE, Chinkes DL, Wolfe RR. Reversal of catabolism by beta-blockade after severe burns. N Engl J Med 345:1223-1229, 2001.

161.      Annane D, Sebille V, Charpentier C, Bollaert P-E, Francois B, Korach J-M, Capellier G, Cohen Y, Azoulay E, Troche G, Chaumet-Riffaut P, Bellissant E. Effect of treatment with low doses of hydrocortisone and fludrocortisone on mortality in patients with septic shock. J Amer Med Assn 288:862-871, 2002.

162.      Hamrahian AH, Oseni TS, Arafah BM. Measurements of serum free cortisol in critically ill patients.N Engl J Med. 2004 Apr 15;350(16):1629-38

163.      Takala J, Ruokonen E, Webster NR, Nielsen MS, Zandstra DF, Vundelinckx G, Hinds CJ. Increased mortality associated with growth hormone treatment in critically ill adults. N Engl J Med:341:785-792, 1999. 164        Van den Berghe G. Non-thyroidal illness in the ICU: a syndrome with different faces. Thyroid. 2014 Oct;24(10):1456-65. doi: 10.1089

AGE-RELATED CHANGES IN THE MALE REPRODUCTIVE AXIS

Revised 1  Nov 2014

Outline     

Aging of male mammals is a very recent evolutionary event observed mostly in humans and animals in captivity.  Most animal species in the wild do not live beyond their reproductive years; during periods of food deprivation, many small animals may not even live beyond puberty. Even among humans, only the men and women of the past two generations have enjoyed a life expectancy of greater than fifty years. With increasing life expectancies on all continents except Africa, the reproductive problems of aging men and women have begun to receive the attention that they deserve. Aging of humans is associated with functional alterations at all levels of the reproductive axis that affect both the steroidogenic and gametogenic compartments.  Even after forty years of investigation, the controversy surrounding the use of hormone replacement in postmenopausal women has grown only louder (1-4); in contrast, the issue of testosterone replacement in older men has been shrouded in acerbic debate from its very inception, even though not a single, adequately powered, long term, randomized trial of testosterone replacement has yet been conducted. There is agreement that in young men with classical hypogonadism due to known diseases of the testis, pituitary and the hypothalamus, testosterone replacement is relatively safe and has many beneficial effects in improving lean body mass, sexual function, energy, mood, and sense of well being and reducing fat mass (5-8). However, the data from otherwise healthy, young, men with classical hypogonadism should not be directly extrapolated to older men with age-related decline in serum testosterone concentrations (9-10). As discussed in this chapter, there is agreement that serum testosterone levels decline as a function of age; however, the effects of testosterone supplementation on health-related outcomes in older men have not been studied. Long term data on the effects of testosterone supplementation on the risk of prostate disease and cardiovascular events are lacking. Thus, the risks and benefits of long term testosterone replacement in older men remain unknown.

  1. I.              CHANGES IN THE STEROIDOGENIC COMPARTMENT OF THE TESTIS
  2. A.           Age Related Changes in Circulating Concentrations of Reproductive Hormones

For many years, there was considerable controversy over whether serum total testosterone levels were lower in healthy older men; it was argued that older men had lower testosterone levels because of the confounding influence of chronic illness and medications. However, a number of cross-sectional studies are in agreement that even after accounting for the potential confounding factors such as time of sampling, concomitant illness and medications, and technical issues related to hormone assays, serum total testosterone levels are lower in older men in comparison to younger men (11-33). Several longitudinal studies (11, 13, 14, 16, 17) have confirmed a gradual but progressive decrease in serum testosterone concentrations from age 20 to 80. In contrast to the sharp reduction in ovarian estrogen production at menopause, the age-related decline in men does not start at a discrete coordinate in old age; rather, total testosterone concentrations, after reaching a peak in the second and third decade, decline inexorably throughout a man’s life (Figure 1). Because of the absence of an identifiable inflection point at which testosterone levels begin to decline abruptly or more rapidly, many investigators have questioned the validity of the concept of “andropause”, which misleadingly implies an abrupt cessation of androgen production in men (20, 34).The term ‘late-onset hypogonadism’ has been proposed to reflect the view that in some middle-aged and older men (> 65 years), the age-related decline in testosterone concentration is associated with a cluster of symptoms and signs in a syndromic constellation which resembles that observed in men with classical hypogonadism (28, 35).

Most studies of age-related change in testosterone levels included healthy, older men. Adiposity, chronic illness, weight gain, medications and genetic factors affect testosterone levels and the trajectory of the age-related decline in testosterone levels in men (12-13, 36-38).  The rate of age-related decline is greater in older men with chronic illness and adiposity than in healthy, non-obese older men (12, 36-37).

Sex-hormone binding globulin concentrations are higher in older men than younger men (13, 24, 29). Thus, the age-related decline in free testosterone levels is of a greater magnitude than that in total testosterone levels. Similarly, there is a greater percent decline in bioavailable testosterone concentrations (the fraction of circulating testosterone which is not bound to SHBG) than in total testosterone concentrations.

B.        Age-Related Decline in Testosterone Levels in Middle-Aged and Older Men

An Expert Panel of the Endocrine Society defined androgen deficiency as a syndrome resulting from reduced production of testosterone and characterized by a set of signs and symptoms in association with unequivocally low testosterone levels (5). Many epidemiologic studies have defined androgen deficiency solely in terms of serum testosterone concentrations below the lower limit of the normal range for healthy, young men leading to exaggerated estimates of the prevalence of androgen deficiency in older men. Additionally, serum testosterone levels in most studies were measured using direct immunoassays, whose accuracy in the low range has been questioned.  Not surprisingly, the estimates of the prevalence of androgen deficiency in older men have varied greatly among different studies. In the Baltimore Longitudinal Study of Aging (BLSA) (11), 30% of men over the age of 60 and 50% of men over the age of 70 had total testosterone concentration below the lower limit of normal range for healthy young men (325 ng/dL, 11.3 nmol/L). The prevalence was even higher when these investigators used a free testosterone index to define androgen deficiency (11). Several other studies have also reported a similarly high prevalence of low total and free testosterone levels in older men. In contrast, more recent studies, using liquid chromatography tandem mass found the prevalence of androgen deficiency to be significantly lower than that observed in the MMAS and BLSA (20-21, 28-31). Although 10–15% of men aged ≥65 years have low total testosterone levels (28-31), the prevalence of late-onset hypogonadism defined by symptoms and a total testosterone level <8 nmol/l in the EMAS was 3.2% for men aged 60–69 years and 5.1% for those aged 70–79 years (28).  A cross-sectional survey performed in Finland (20), which did not use a random probability sample, found that only 27% of those who had high andropausal symptom score had androgen deficiency, defined as serum testosterone less than 287 ng/dL (10 nmol/L).  In this study, most of the older men with low testosterone levels had a systemic disease; less than 3% of healthy, older men had low testosterone levels. The Healthy Man Study in Australia also found no significant age-related decline in testosterone or dihydrotestosterone in men who reported being in good health (39). These authors have argued that ill health, rather than aging itself, is the major contributor to androgen deficiency in older men.

C.        Mechanisms of Age-Related Decline in Testosterone Levels

Circulating testosterone concentrations are a function of testosterone production and clearance rates; the age-related decline in serum testosterone concentrations is primarily a consequence of decreased production rates in older men (9, 10, 24-26, 29). Plasma clearance rates of testosterone are, in fact, lower in older men than in younger men (37-38). The decline in testosterone production in older men is the result of abnormalities at all levels of the hypothalamic-pituitary-testicular axis (23-25, 39-50)

C.1. Gonadotropin Secretion and Regulation in Older Men.  There is considerable heterogeneity in circulating LH and FSH concentrations in individual older men; both hypogonadotropic and hypergonadotropic hypogonadism have been reported (35, 37). As a group, serum LH and FSH concentrations are higher in older men than in young men (13-14). Serum LH and FSH levels show an age-related increase in longitudinal studies. However, serum LH concentrations do not increase in proportion to the age-related decline in circulating testosterone levels, probably due to the impairment of GnRH secretion and alterations in gonadal steroid feedback and feedforward relationships (39-50); both of these mechanisms are operative in older men.

The data on LH response to GnRH are somewhat contradictory. Urban et al (44) used an interstitial cell bioassay to measure serum concentrations of bioactive LH and found that although basal bioactive LH concentrations were similar in this sample of young and older men, older men demonstrated diminished LH response to GnRH administration. However, in a subsequent study, Zwart et al (45) found greater gonadotropin responsiveness to GnRH in older men than younger men; the maximal and incremental LH and FSH secretory masses in response to graded doses of GnRH were significantly higher in healthy, older men than in younger men. The estimated half-lives of LH, FSH, or alpha-subunit were not significantly different between young and older men.

The Brown Norway rat has been widely used as a model of reproductive aging. In this experimental model, the prepro-GnRH mRNA content and the number of neurons expressing prepro-GnRH mRNA are lower in older male rats in comparison to young rats (46-47). The GnRH content of several hypothalamic areas is also lower in intact older rats than younger rats (46). Older Brown Norway rats exhibit significant reductions in glutamate and -aminobutyric acid (GABA) levels in the hypothalamus compared to young rats (47). These observations suggest that the decreased hypothalamic excitatory amino acid expression and the reduced responsiveness of GnRH neurons to NMDA may contribute to the altered LH pulsatile secretion observed in old rats (47).

Infusions of testosterone and DHT are associated with greater reductions in mean serum LH and FSH levels and the frequency of LH pulses in older men in comparison to young men (48). Winters et al (43) reported that the degree of LH inhibition during testosterone replacement of older, hypogonadal men was significantly greater than in young, hypogonadal men suggesting that older men are more sensitive to the feedback inhibitory effects of testosterone on LH. Deslypere et al (48) also found decreased LH pulse frequency and a greater degree of LH inhibitory response to estradiol administration in older men than young controls. Age-related increase in FSH levels is not associated with a progressive or proportionate decrease in inhibin B levels (49).  Thus the mechanistic basis of FSH increase in not fully understood, although the lack of change in inhibin B levels suggests that Sertoli cell function is relatively preserved in older men.

Pulsatile GnRH secretion is attenuated in older men. In addition, there are disturbances of the feedback and feed-forward relationships between testosterone and LH secretion (42, 50-51). Thus, the sensitivity of pituitary LH secretion to androgen-mediated feedback inhibition is increased; in addition, the ability of LH to stimulate synchronously testicular testosterone secretion (feedforward) is attenuated (42, 50-51). This insight has emerged largely from the research of Veldhuis who used novel algorithms to quantitate the orderliness of pulsatile hormone secretion, and the synchrony between secretion of related hormones (e.g., LH and testosterone, and LH and FSH) (42, 50-51). This research has revealed that the orderliness of LH pulses and the synchrony between LH and testosterone pulses are decreased in older men (50-51); in addition, there is greater variability in LH pulse frequency, amplitude, and secretory mass in older men, in comparison to younger men (50-51).

C.2. Testicular Testosterone Production in Older Men.  Testosterone secretion in healthy, young men is characterized by a diurnal rhythm with higher concentrations in the morning and lower levels in later afternoon. Many studies have revealed that the diurnal rhythm of testosterone secretion is dampened in older men (22, 32). Testosterone response to LH and human chorionic gonadotropin is decreased in older men in comparison to younger men (23-25).

D.        Physiological and Clinical Correlates of Age-Related Decline in Circulating Testosterone in Epidemiological Studies

Many of the physiological changes that occur with advancing age, such as loss of bone and muscle mass, increased fat mass, impairment of physical, sexual and cognitive functions, loss of body hair, and decreased hemoglobin levels, are similar to those associated with androgen deficiency in young men. Aging is associated with loss of skeletal muscle mass (Figure 2), muscle strength and power, and progressive impairment of physical function (52-76). Epidemiological studies of older men have reported associations between low testosterone levels and health outcomes, although these associations are weak. For instance, in a number of epidemiologic studies, such as the St. Louis Inner City Study of Aging  Men (55), the Olmsted County Epidemiological Study (54), and the New Mexico Elderly Health Study (57-58), low bioavailable testosterone levels were associated with low appendicular skeletal muscle mass. Low bioavailable testosterone levels also have been associated with decreased strength of upper as well as lower extremity muscles (55-56) and decreased performance in self-reported as well as performance-based measures of physical function (77-81). Low free testosterone levels have also been associated with the development of mobility limitation and the frailty syndrome (82-85).

The association of testosterone levels with sexual dysfunction has been inconsistent across studies because of the heterogeneity of instruments used to define sexual dysfunction, varying quality of instruments used to assess sexual function, problems of testosterone assay quality, and failure to distinguish among various categories of sexual dysfunction (86-91). Androgen deficiency and erectile dysfunction are two independently distributed clinical disorders; in general, serum total and bioavailable testosterone levels are not significantly different between men who report erectile dysfunction and those who do not (90-91). In the MMAS, decreased libido, as assessed by a single question, was associated only with very low testosterone levels (86). In the EMAS, total and free testosterone levels were associated with overall sexual function in middle-aged and older men (28). This relationship was observed more robustly at testosterone concentrations <8 nmol/L, but not at higher testosterone concentrations (92). Men deemed to have low total and free testosterone levels in EMAS, using reference ranges generated in healthy young men, were more likely to report decreased morning erections, erectile dysfunction, and decreased frequency of sexual thoughts than those with normal testosterone levels (29).In another study of men over the age of 50 who had benign prostatic hyperplasia, sexual dysfunction, assessed by the Sexual Function Inventory, was reported only in men with serum total testosterone levels less than 225 ng/dL (88).

Aging of humans is attended by a decline in several aspects of cognitive function; of these multiple domains of cognition that decline with aging, declines in verbal memory, visual memory, spatial ability, and executive function are associated with the age-related decline in testosterone(87-101).

The relationship of testosterone levels with depression has been inconsistent across epidemiologic studies (102-106). Low testosterone levels in older men appear to be associated more with subsyndromal depression and dysthymia than with major depression (105-106). In one study, testosterone levels were lower in older men with dysthymic disorder than in those without any depressive symptoms (106). In another study (107), men with low testosterone levels had higher Carrol Rating Index scores, indicating more depressive symptoms than those who had normal testosterone levels.

Several epidemiologic studies of older men (108-112), including MrOS(108), Rancho Bernardo Study (109), Framingham Heart Study (110), and the Olmsted County Study (111) – have found bioavailable testosterone levels to be associated with bone mineral density, bone geometry, and bone quality (112); the associations are stronger with bioavailable testosterone and estradiol levels than with total testosterone levels. In the MrOS Study, the odds of osteoporosis in men with a total testosterone less than 200 ng/dL were 3.7 fold higher than in men with normal testosterone level (108); free testosterone was an independent predictor of prevalent osteoporotic bone fractures (115).

Several studies have evaluated the association of testosterone levels and mortality (116-119)(118). Some, but not all, studies found higher all-cause mortality and cardiovascular mortality in men with low testosterone levels than in those with normal testosterone levels. In a meta-analyses of epidemiologic studies of community-dwelling men, low testosterone levels were associated with an increased risk of all-cause and CVD death (Figure 3) (120-121). However, the strength of the inferences of these meta-analyses was limited by considerable heterogeneity in study populations; it is possible that effects may have been driven by differences in the age distribution and the health status of the study populations (120-124).

Testosterone levels are not correlated with aging-related symptoms assessed by the Aging Male Symptom (AMS) score or with lower urinary tract symptoms assessed by the IPSS/AUA prostate symptom questionnaire (122). Some cross-sectional studies found no difference in serum testosterone levels between men who had coronary artery disease and those who did not have coronary artery disease; other studies have reported testosterone levels to be lower in men with coronary artery disease than in men without coronary artery disease (123-128).

  1. E.            
  2. F.            Potential Beneficial Effects of Testosterone Therapy in Older Men with Low Testosterone Levels

It has been hypothesized that increasing serum testosterone concentrations in older men with low testosterone levels into a range that is mid-normal for healthy, young men would improve physical function and mobility, some domains of sexual and cognitive functions, energy and sense of wellbeing, and reduce the risk of falls and fractures, and improve overall quality of life. A number of randomized trials have demonstrated improvements in surrogate markers, such as lean and fat mass; however, there has been a paucity of long-term, placebo-controlled, randomized trials that are adequately powered to detect clinically meaningful changes in health outcomes such as fracture rates, physical function, disability, progression to dementia, and overall quality of life. Furthermore, none of the previously published studies had sufficient power to address the long term risks of prostate and cardiovascular disease.

The following section describes the effects of testosterone supplementation on multiple organs systems focusing on muscle mass and performance, physical function, bone mineral density and fracture risk, sexual function, mood, and cognitive function.

E.1.     Effects of Testosterone Supplementation on Muscle Mass and Performance and Physical Function in Older Men with Low Testosterone Levels.

E.1.a. Age-Related Changes in Muscle Mass and Performance. Sarcopenia, the loss of muscle mass and function, is an important consequence of aging (53-57). The prevalence of sarcopenia, depending on the definition used, varies from 10-30% in men over the age of 70 (53-57). The principal component of the decrease in fat-free mass is the loss of muscle mass; there is little change in non-muscle lean mass (59-65). Between 20 and 80 years of age, the skeletal muscle mass decreases by 35-40% in men (63), in part due to decreased muscle protein synthesis (70). Although there is a loss of both type I and type II fibers, there is a disproportionate decrease in the number of type II muscle fibers that are important for generation of power (71-72). In spite of the significant depletion of muscle mass, body weight does not decrease, and may even increase because of the corresponding accumulation of body fat (59-65) (Figure 3).

The loss of muscle mass that occurs with aging is associated with a reduction in muscle strength (73-76). There is a substantial decrease in muscle strength and power between 50 and 70 years of age, primarily due to muscle fiber loss and selective atrophy of type II fibers (71-76).  The loss of muscle strength is even greater after the age of 70; 28% of men over the age of 74 could not lift objects weighing more than 4.5 kg (75). With increasing age, there is a progressive reduction in muscle power (129-130), the speed of strength generation, and fatigability, the ability to persist in a task.

Loss of muscle mass and strength leads to impairment of physical function, as indicated by the impaired ability to arise from a chair, climb stairs, generate gait speed, and maintain balance (129-132). The impairment of physical function contributes to loss of independence, depression, and increased risk of falls and fractures in older men. Therefore, function promoting therapies that can reverse or prevent aging-associated sarcopenia are desirable.

E.1.b. Anabolic Effects of Testosterone in Humans: Testosterone Trials in Healthy, Hypogonadal Men, Men with Chronic Illness, and Older Men.   The anabolic effects of testosterone on the muscle have been a source of intense controversy for over sixty years. The athletes and recreational body builders abuse large doses of androgenic steroids with the belief that these compounds increase muscle mass and strength. Until recently, the academic community was skeptical about such claims because of the problems of study design. However, a number of studies in healthy hypogonadal men, men with chronic illness, and in healthy older men have established that testosterone administration increases skeletal muscle mass, maximal voluntary strength, and leg power (132-139).

In a systematic review of testosterone trials in healthy, hypogonadal men, testosterone therapy increased fat-free mass and body weight (Figure 4) (132-139). Some studies of testosterone replacement therapy have reported significant improvements in maximal voluntary strength (135, 138), and decreases in whole body fat mass (134, 137-139). The administration of supraphysiologic doses of testosterone in eugonadal men also increases fat-free mass, muscle size, and maximal voluntary strength (140-143). Resistance exercise training augments the anabolic response to androgens; thus men receiving testosterone and resistance exercise training together experience greater gains in fat-free mass and muscle strength than those receiving either intervention alone (143).

The anabolic effects of testosterone on fat-free mass, muscle size, and maximal voluntary strength are related to the administered testosterone dose and the circulating testosterone concentrations (140-142) (Figure 5). Testosterone effects on muscle performance are domain-specific: testosterone administration increases maximal voluntary strength and leg power, but does not affect fatigability or specific tension (141). The gains in maximal voluntary strength during testosterone administration are proportional to the increase in muscle mass; unlike resistance exercise training, testosterone does not improve the contractile properties of the human skeletal muscle (141).

Testosterone replacement of young, hypogondal men increases muscle protein synthesis (136, 144-145); the effects of testosterone replacement on muscle protein degradation need further investigation.

Systematic reviews(133, 146-147) of randomized, placebo-controlled trials in HIV-infected men with weight loss (147-152)have revealed that testosterone therapy for 3 to 6 months was associated with greater gains in lean body mass than placebo administration (difference in lean body mass change between placebo and testosterone arms 1.22 kg, 95% CI 0.23-2.22 for the random effect model). In two (147, 151) out of three trials that measured muscle strength (147, 151-152), testosterone administration was associated with significantly greater improvements in maximal voluntary strength than placebo. Testosterone therapy had a moderate effect on depression indices (-0.6, 95% CI -1.0, -0.2) (153)and fatigue (154), but did not improve overall quality of life (153-154). Changes in CD4+ T lymphocyte counts, HIV copy number, PSA, plasma HDL cholesterol, and adverse event rates were not significantly different between the placebo and testosterone-treatment groups (147-154). Overall, short-term (3-6 months) testosterone use in HIV-infected men with low testosterone levels and weight loss can induce modest gains in body weight and lean body mass with minimal changes in quality of life and mood. This inference is weakened by inconsistency of results across trials, heterogeneity in inclusion and exclusion criteria, disease status, testosterone formulations and doses, treatment duration, and methods of body composition analysis (133). There are no data on testosterone effects on physical function, risk of disability, or long term safety.

Testosterone administration increases fat-free mass and decreases fat mass in older men with low testosterone levels. Meta-analyses (133, 154) of randomized trials (156-160) that included middle-aged and older men with low or low normal testosterone levels, and that used testosterone or its esters in replacement doses for >90 days, have confirmed that testosterone administration is associated with a significantly greater increase in fat-free mass, hand grip strength, and a greater reduction in whole body fat mass than placebo (Figure 5). The average gains in fat-free mass generally were greater in trials that used injectable testosterone esters than in those which used transdermal testosterone gel. The change in body weight did not differ significantly between the testosterone and placebo groups. Testosterone administration improves stair climbing speed and power, and self-reported physical function, as assessed by the SF-36 questionnaire. Changes in other performance-based measures of physical function, such as gait speed have been inconsistent across trials (157, 160-162).

One reason for the failure to demonstrate improvements in physical function is that the measures of physical function used in previous studies had low ceilings. The widely used measures such as 0.625 m stair climb, standing up from a chair, and  20-meter walk are tasks that require only a small fraction of an individual’s maximal voluntary strength. In most healthy, older men, the baseline maximal voluntary strength is far higher than the threshold below which these measures would detect impairment. Another confounder of the effects of anabolic interventions on muscle function is the learning effect. For instance, subjects who are unfamiliar with weight lifting exercises often demonstrate improvements in measures of muscle performance because of increased familiarity with the exercise equipment and technique. Because of the considerable test-to-test variability in tests of physical function, it is possible that previous studies did not have adequate power to detect meaningful differences in measures of physical function between the placebo and testosterone-treated groups. It is also possible that neuromuscular adaptations needed to translate strength gains into functional improvements require a lot longer than the 3 to 6 month duration of most of the previous trials.

Although most randomized testosterone trials have been conducted in healthy older men, three recent trials were conducted in older men with functional limitations (163-166).In a trial of pre-frail or frail men (166), administration of 50 mg testosterone gel daily for 6 months induced greater improvements in lean mass, knee extension peak torque and sexual symptoms than did placebo gel (166). Performance-based measures of physical function did not differ significantly between groups, but they improved in the subgroup of frail elderly men (166). In Testosterone in Older Men (TOM) Trial, older men with mobility limitation were randomly assigned to either placebo or 10 g testosterone gel daily for 6 months (163-164). The testosterone dose was adjusted to achieve testosterone levels between 17.4 nmol/l and 34.7 nmol/L (500 to 1000 ng/dL).The improvements in leg-press strength, chest-press strength and power, and loaded stair-climbing speed and power were significantly greater in men assigned to testosterone arm than in those receiving placebo (Figure 6). A greater proportion of men in the testosterone arm improved more than the minimal clinically important difference for leg-press and chest-press strength and stair-climbing speed than that in the placebo arm. Because of a higher frequency of cardiovascular-related events in the testosterone arm compared with the placebo arm, the trial’s data and safety monitoring board stopped further administration of study medication (163-164).

Therefore, while there is strong evidence that testosterone supplementation increases skeletal muscle mass and strength, the clinically important improvements in health outcomes - physical function, falls, fractures and disability -  in men with clinical conditions such as mobility limitation or fall propensity have been difficult to demonstrate. Innovative strategies to translate gains in muscle mass and strength induced by testosterone into functional improvements are needed (167). Adjunctive exercise training might be required to induce the neuromuscular and behavioural adaptations that are necessary to translate the gains in muscle mass and strength into functional improvements (167). The findings of the TOM trial and other epidemiologic studies have heightened the concern that frail elderly men with a high burden of chronic co-morbidities may be at an increased risk of adverse events (164), providing the impetus to develop, strategies to achieve increased selectivity and a more favourable risk to benefit ratio (164).

E.1.c. Mechanisms of Androgen Action on the Muscle Testosterone-induced increase in muscle mass is associated with hypertrophy of both type I and II muscle fibers(168). The absolute number and the relative proportion of type I and type II fibers do not change during testosterone administration. Testosterone-induced muscle fiber hypertrophy is associated with dose-dependent increases in myonuclear number and satellite cell number (169).

Testosterone administration has been shown to increase fractional muscle protein synthesis and improve the reutilization of amino acids (144-145). The effects of testosterone on muscle protein degradation have not been well studied. However, the muscle protein synthesis hypothesis does not explain the reciprocal decrease in fat mass or the increases in myonuclear and satellite cell number that occur during testosterone administration (169). Testosterone promotes the differentiation of mesenchymal multipotent muscle progenitor cells into the myogenic lineage and inhibits the differentiation of these precursor cells into the adipogenic lineage (170-171). Thus, testosterone promotes the formation of myosin heavy chain II positive myotubes in multipotent cells and up-regulates markers of myogenic differentiation, such as MyoD and myosin heavy chain (170-171). Testosterone and DHT inhibit adipogenic differentiation and downregulate markers of adipogenic differentiation, such as PPAR-¡ and C/EBPµ.

Testosterone’s effects on myogenic differentiation are mediated largely through its binding to the classical androgen receptor, which induces a conformational change in the androgen receptor protein, promoting its association with its co-activator, beta-catenin, causing the complex to translocate into the nucleus (171-172). The androgen receptor – beta-catenin complex associates with TCF-4 and activates a number of Wnt target genes (171-172), including follistatin. Follistatin cross-communicates the signal from the AR-beta- catenin pathway to the TGF-beta signaling pathway, blocking signaling through the TGF-beta / Smad 2/3 (173-174). Follistatin plays an essential role in mediating the effects of testosterone on myogenic differentiation (174-175). In a remarkable discovery, Jasuja et al (175) found that the administration of recombinant follistatin selectively increased muscle mass and decreased fat mass but had no effect on prostate growth. Recombinant follistatin and testosterone each regulated the expression of a large number of common genes in the skeletal muscle, but they differed substantially in the expression profile of genes activated in the prostate (175). Among the genes activated differentially by testosterone but not by follistatin in the prostate, Jasuja et al (175) identified polyamine pathway as an important signaling pathway. The polyamine pathway has been known to be involved in regulating prostate growth. Administration of testosterone in combination with an inhibitor of ornithine decarboxylase, a key enzyme in the polyamine pathway, to castrated male mice restored levator ani muscle mass but not prostate mass, indicating that ODC1 plays an important role in mediating the effects of testosterone on the prostate(Figure 7) (175). Therefore, combined administration of testosterone plus ODC1 inhibitor provides a novel approach for achieving selectivity of testosterone’s anabolic effects on the muscle while sparing the prostate (175).

E.1.d. The Role of Steroid 5-alpha Reductase and DHT in Mediating Androgen Effects in the Muscle. Although the enzyme steroid 5-alpha-reductase is expressed at low concentrations within the muscle (176-177), we do not know whether conversion of testosterone to dihydrotestosterone is required for mediating the androgen effects on the muscle. Men with benign prostatic hypertrophy who are treated with a 5-alpha reductase inhibitor do not experience muscle loss (178). Similarly, individuals with congenital 5-alpha-reductase deficiency have normal muscle development at puberty (178). These data suggest that 5-alpha reduction of testosterone to DHT is not obligatory for mediating its effects on the muscle. However, all the kindred with 5-alpha reductse deficiency that have been published to-date have had mutations of type 2 isoform of the enzyme. Similarly, finasteride is a weak inhibitor of only the type 2 isoform of the enzyme. The circulating concentrations of DHT in male patients with congenital mutation of type 2 5-alpha reductase enzyme or in men treated with finasteride are lower than eugonadal men; however, these patients still produce significant amounts of DHT and their circulating DHT concentrations are often in the lower end of the male range. It is reassuring that long term administration of dutasteride, a dual and potent inhibitor of both 5-alpha reductase isoforms, has not been associated with significant reductions in bone mineral density (178); the data on the effects of duatasteride on muscle mass are not available. This issue is important because if 5-alpha reduction of testosterone to DHT were not obligatory for mediating its anabolic effects on the muscle, then it might bebeneficial to administer testosterone with an inhibitor of 5-alpha reductase or to develop selective androgen receptor modulators that do not undergo 5-alpha reduction.

To determine whether testosterone’s effects on muscle mass and strength, sexual function, hematocrit, prostate, sebum production, and lipids are attenuated when its conversion to DHT is blocked, we administered to healthy men, 18-50 years, a long-acting GnRH-agonist to suppress endogenous testosterone. We randomized them to placebo or dutasteride (dual inhibitor of steroid 5-alpha reductase type 1 and 2) 2.5-mg daily, plus 50, 125, 300, or 600-mg testosterone enanthate weekly for 20-weeks (179). Changes in lean and fat mass, leg-press and chest-press strength, were related to testosterone dose but did not differ between placebo and dutasteride groups (179). The relationship between testosterone concentrations and the change in lean mass, muscle strength, hematocrit, sebum production and PSA were similar between dutasteride and placebo arms(Figure 8) (179). Changes in sexual-function scores, bone markers, prostate volume, and PSA did not differ between groups (179).  These data indicate that testosterone’s conversion to DHT is not essential for mediating its effects on muscle mass and strength, sexual function, hematocrit, or sebum in men over the range of testosterone concentrations achieved in this trial (179). These data are consistent with studies that have reported that administration of 5α-reductase inhibitors has little or no effect on muscle or bone mass (180-182).

E.1.e. The Role of CYP19aromatase in Mediating Testosterone Effects on the Muscle. Studies of aromatase knockout mice have revealed higher fat mass and lower muscle mass in mice that are null for the P450-linked CYP19 aromatase gene (183). Similarly, humans with CYP19 aromatase mutations have decreased muscle mass and increased fat mass, and they exhibit insulin resistance (184). Data from these gene-targeting experiments suggest that aromatization of testosterone to estradiol might also be important in mediating androgen effects on body composition. Finkelstein et al (185) have recently examined the relative roles of testosterone and estradiol in regulation of muscle and fat mass, and sexual function. These investigators found that testosterone’s effects on lean mass, muscle size, and strength were not significantly attenuated when its conversion to estradiol was blocked by administration of an aromatase inhibitor (185). However, testosterone’s effects on fat mass and sexual desire appeared to be mediated by estradiol (185).

E.2. Testosterone and Sexual Function in Older Men

E.2.a. Regulation of Sexual Function by Testosterone   Sexual function in men is a complex process that includes central mechanisms for regulation of sexual desire and arousability, and local mechanisms for penile tumescence, orgasm, and ejaculation (186) . Primary effects of testosterone are on sexual interest and motivation (186-191). Testosterone replacement of young, androgen deficient men improves a wide range of sexual behaviors including frequency of sexual activity, sexual daydreams, sexual thoughts, feelings of sexual desire, spontaneous erections, and attentiveness to erotic stimuli (186-194).  Kwan et al (190) demonstrated that androgen-deficient men have decreased frequency of sexual thoughts and lower overall sexual activity scores; however, these men can achieve erections in response to visual erotic stimuli. Hypogonadal men have lower frequency and duration of the episodes of nocturnal penile tumescence; testosterone replacement increases both the frequency and duration of sleep-entrained, penile erections (192-194). Although both orgasm and ejaculation are believed to be androgen-independent, hypogonadal men have decreased ejaculate volume and their orgasm may be delayed.

Although hypogonadal men can achieve erections, it is possible that achievement of optimal penile rigidity might require physiologic testosterone concentrations. Testosterone regulates nitric oxide synthase activity in the cavernosal smooth muscle (195). Testosterone administration in orchidectomized rats increases penile blood flow and has trophic effects on cavernosal smooth muscle (196-198).

In male rodents, all measures of mating behavior are normalized by relatively low testosterone levels that are insufficient to maintain prostate and seminal vesicle weight (199-200). Similarly, in men, sexual function is maintained at relatively low normal levels of serum testosterone(185, 191, 201). Testosterone’s effects on libido may be mediated through its conversion to estradiol (185).

E.2.b. Relationship of Androgen Deficiency and Erectile Dysfunction in Middle-Aged and Older Men   Erectile dysfunction and androgen deficiency are two common but independently distributed, clinical disorders that sometimes co-exist in the same patient (186, 202-204). Hypogonadism is a clinical syndrome that results from androgen deficiency (5); in contrast, erectile dysfunction is usually a manifestation of a systemic vasculopathy, often of atherosclerotic origin. Thus androgen deficiency and erectile dysfunction have distinct pathophysiology. Eight to ten percent of men presenting with erectile dysfunction have low testosterone levels (203-206). The prevalence of low testosterone levels is not significantly different between middle aged and older men with impotence and those without impotence (203-206). Testosterone administration does not improve sexual function in men with erectile dysfunction who have normal testosterone levels (207-210). In men with sexual dysfunction who have unequivocally low testosterone levels, testosterone therapy improves libido and overall sexual activity (209-210). The response to testosterone supplementation in this group of men is variable because of the co-existence of other disorders such as diabetes mellitus, hypertension, cardiovascular disease, and psychogenic factors. Several meta-analyses of the usefulness of androgen replacement therapy concluded that testosterone administration is associated with greater improvements in sexual function compared to placebo treatment in men with sexual dysfunction and unequivocally low testosterone levels (208-210).

It had been speculated that testosterone administration might improve erectile response of men with ED to selective phosphodiesterase inhibitors (211-213). To determine whether the addition of testosterone to a phosphodiesterase-5-inhibitor improves erectile response, we conducted a randomized, placebo-controlled trial (214), in men, 40-to-70 years, with erectile dysfunction and low total testosterone< 11.5 nmol/L (330ng/dL) and/or free testosterone<173.5 pmol/L (50pg/mL). All participants were initially started on sildenafil alone and the sildenafil dose was optimized based on their response during a 3-7 week run-in period (214). The participants were then randomized to 10-g testosterone or placebo gel for 14-weeks in combination with the optimized sildenafil dose (214).  The administration of sildenafil alone was associated with substantial increases in erectile function domain (EFD) score and total and satisfactory sexual encounters (214). However, the change in EFD score in men assigned to testosterone plus sildenafil did not differ significantly from that in men assigned to placebo plus sildenafil (214).  Changes in total and successful sexual encounters, quality-of-life, and marital-intimacy did not differ between testosterone and placebo groups. Even among the subsets of men with baseline testosterone<250ng/dL or those without diabetes, there were no significant differences in EFD scores between the two arms (214).  Another placebo-controlled trial of men with erectile dysfunction who were non-responders to tadalafil also did not show a greater improvement in erectile function in men assigned to the testosterone arm than in those assigned to the placebo arm (213). Thus, in randomized trials, the addition of testosterone to PDE5Is has not been shown to improve erectile function in men with erectile dysfunction (213-214).

Androgen deficiency is an important cause of low sexual desire disorder (186). Therefore, serum testosterone concentrations should be measured in the diagnostic evaluation of hypoactive sexual desire disorder, recognizing that low sexual desire is often multifactorial; systemic illness, relationship and differentiation (the ability of individuals in a relationship to maintain their distinct identities) issues, depression, and many medications can be important antecedents or contributors to low sexual desire and sexual dysfunction.

E.3. Testosterone Effects on Bone Mineral Metabolism  

E.3.a.  The Effects of Androgen Deficiency on Bone Mass.  Testosterone deficiency is associated with a progressive loss of bone mass (215-218). In one study performed in sexual offenders (215), surgical orchiectomy was associated with a progressive decrease in bone mineral density of a magnitude similar to that seen in women after menopause. Similarly, androgen deficiency induced by the administration of a GnRH agonist, surgical orchiectomy, or an androgen antagonist for the treatment of prostate cancer leads to loss of bone mass (216-218). In male rats, surgical orchiectomy or androgen blockade by administration of an androgen receptor antagonist is associated with loss of bone mass (219).

Androgen deficiency that develops before the completion of pubertal development is associated with reduced cortical and trabecular bone mass (220-221). During the pubertal years, significant bone accretion occurs under the influence of sex steroids; therefore, individuals with sex-steroid deficiency before or during peri-pubertal years may end up with suboptimal peak bone mass. Similarly, men with acquired androgen deficiency have lower bone mineral density than age-matched controls(133).

E.3.b. Effects of Testosterone Administration in Young, Androgen-Deficient Men.  Testosterone therapy of healthy, young, hypogonadal men is associated with significant increases in vertebral bone mineral density (134, 222-226). However, bone mineral density is typically not normalized after 1-2 years of testosterone replacement therapy (134). The reasons for the failure of testosterone replacement therapy to normalize bone mineral density in androgen-deficient men are not entirely clear. Some hypogonadal participants patients included in these testosterone trials had panhypopituitarism and also suffered from growth hormone deficiency. It is possible that concomitant GH replacement might be necessary for restoration of normal bone mineral density. Excessive glucocorticoid replacement might also contribute to bone loss in these patients. In addition, some participants had experienced testosterone deficiency before the onset and completion of pubertal development. Because maximal bone mass is achieved in part through bone accretion during the peripubertal period under the influence of sex-steroid hormones, the individuals who develop androgen deficiency during the critical pubertal developmental window of bone accretion, may end up with decreased peak bone mass, and testosterone administration may not be able to restore bone mass to levels seen in eugonadal age-matched controls. Many testosterone replacement trials were less than 3 years in duration, and it is possible that a longer period of testosterone administration might be necessary to achieve maximal improvements in bone mineral density. Indeed, Behre et al (222) reported that bone mineral density in some hypogonadal men continued to increase even after many years of testosterone treatment using a scrotal transdermal patch and reached the levels expected for age-matched eugonadal controls.

E.3.c. Cross-sectional Studies of the Relationship Between Sex-Hormone Concentrations and Osteoporosis in Older Men.  The age-related decline in sex hormones is associated with age-related changes in bone mineral density and increased risk of osteoporotic fractures (108-115). Older men with hip fractures have lower testosterone levels than age-matched controls (227). In general, epidemiologic studies have reported bioavailable testosterone and estradiol levels to be more strongly associated with bone mineral density of the spine, hip, and distal radius than total testosterone levels (109, 111-112, 114).

E.3.d. Testosterone Replacement Studies in Older Men.  Three long-term studies of testosterone replacement of relatively healthy older men have examined the effects of testosterone on bone mineral density but have reported inconsistent results (162, 228-230). One study found greater increases in vertebral bone mineral density in the testosterone arm of the trial than in the placebo arm, while another study did not find any significant differences between the change in vertebral or femoral bone mineral density between testosterone and placebo groups (230). The third study reported greater gains in bone mineral density of the femoral neck but not of other regions in men randomized to receive testosterone compared to those who received placebo. A meta-analysis of randomized trials found a significantly greater increase in lumbar bone mineral density but not in femoral bone mineral density in the testosterone arms of trials that used intramuscular testosterone than in placebo arms (231); transdermal testosterone had no significant impact.

E.3.e. Mechanisms of Androgen Action on the Bone.  Testosterone increases bone mass by several mechanisms (232). Short-term studies of androgen replacement have shown inconsistent increases in markers of bone formation, but a more consistent reduction in markers of bone resorption (232-235). These observations suggest that testosterone increases bone mineral density in part through its aromatization to estrogen, which inhibits bone resorption. Estrogen deficiency contributes to increased bone resorption and remodeling by multiple mechanisms. Estrogens regulate the activation frequency of bone functional basic multicellular units, the duration of the resorption phase and the formation phase, and osteoclast recruitment (236). The protective effects of estrogen on bone in both male and female mice during growth and maturation are mediated largely through estrogen receptor-alpha (237-243). In men androgens and estrogens both play independent roles in regulating bone resorption (236). In addition, there is increasing evidence that testosterone might also directly stimulate osteoblastic bone formation. Androgen receptors have been demonstrated on osteoblasts and on mesenchymal stem cells (244). Testosterone stimulates cortical bone formation (245). Testosterone also stimulates the production of several growth factors within the bone, including IGF-1; these growth factors may contribute to bone formation (246). Testosterone increases muscle mass, which may indirectly increase bone mass by increased loading. Testosterone might inhibit apoptosis of osteoblasts through non-genotropic mechanisms (247-248). In addition to its effects on bone mineral density, testosterone might reduce fall propensity because of its effects on muscle strength and reaction time.

E.3.f. Synopsis.  Testosterone replacement has been shown to increase vertebral bone mineral density in young and older men with unequivocally low testosterone levels (5). Testosterone increases bone mass by multiple mechanisms. However, testosterone effects on fracture risk have not been studied.

E.4. Testosterone Effects on Cognitive Function

E.4.a. Cross-sectional Studies Correlating Sex-Hormone Levels and Cognitive Function.  Androgens effects on cognitive function are domain-specific. For instance, observations that men outperform women in a variety of visuo-spatial skills suggest that androgens enhance visuo-spatial skills (249). In !Kung San Bushmen of Southern Africa, testosterone, but not estradiol, levels correlated with better spatial ability and with worse verbal fluency (250).  Circulating levels of dihydrotestosterone, a metabolite of testosterone that is not converted to estrogen, positively correlated with verbal fluency (250). Barrett-Conner et al (251) found positive associations between total and bioavailable testosterone levels, and global cognitive functioning and mental control, but not with visuospatial skills. In the Baltimore Longitudinal Study of Aging (252), higher free testosterone index was associated with better scores on visual and verbal memory, visuospatial functioning, and visuomotor scanning. Men with low testosterone levels had lower scores on visual memory and visuospatial performance (252). Neither total testosterone level nor the free testosterone index was correlated with verbal knowledge, mental status, or depressive symptoms (252). Other studies have reported a complex relationship between androgen levels and spatial ability (253--256). Women with congenital adrenal hyperplasia with high androgen levels score higher on tests of spatial cognition than their age- and gender-matched siblings (257). 46, XY rats with androgen insensitivity perform worse on tests of spatial cognition than their age-matched controls (258).

E.4.b. Intervention Trials of the Effects of Testosterone Supplementation on Cognitive Function. Several small clinical trials in elderly hypogonadal men have provided conflicting results (259-265); not surprisingly, a systematic review of clinical trials revealed no significant effects of testosterone on cognition (5). Janowsky et al (259) tested verbal and visual memory, spatial cognition, motor speed and cognitive flexibility in a group of healthy older men who received 3 months of testosterone supplementation. Testosterone replacement was associated with a significant improvement in spatial cognition only. Serum testosterone levels were not significantly correlated with spatial performance, but estradiol levels showed a significant inverse relationship with spatial performance suggesting that estradiol might inhibit spatial ability. Vaughan et al (260) found no effect of testosterone administration on cognition, while Cherrier et al (261-263) reported an effect on visuo-spatial cognition. Testosterone also enhanced verbal fluency. Hypogonadal men performed worse on tests of verbal fluency than eugonadal men, and showed improvement after testosterone replacement (264-266). In transsexual males (267), administration of anti-androgen and estrogen, prior to surgery for gender reassignment, decreased anger and aggression proneness, sexual arousability, and spatial skills, and increased verbal fluency ability. Conversely, testosterone administration to females decreased verbal fluency and increased spatial skills. Testosterone administration may also improve verbal memory in women (268).

Testosterone is aromatized to estrogen in the brain, and some effects of testosterone on cognition might be mediated through its conversion to estradiol. Androgen receptors are expressed in the brain (269), and androgen effects on brain organization during development (270-271) are mediated through androgen receptor. Androgens increase neurite arborization, facilitating intercellular communication (270-273). Testosterone also affects serotonin, dopamine, acetylcholine (272 ), and calcium signaling (273).

E.4.c. Synopsis of the Effects of Testosterone on Cognition. The literature on testosterone and cognition is highly equivocal; some, but not all studies, demonstrate improvements in tests of spatial cognition, verbal fluency and verbal memory. The inconsistency in findings cannot yet be interpreted as evidence that there is no effect. Rather methodological problems appear to limit the generalizability of results. Limitations of previous studies include limited sample sizes with heterogeneous, poorly defined samples; the use of a variety of neuropsychological tests, including some that lack psychometric validation; and the use of differing protocols in clinical trials. The effects of testosterone therapy on clinically important outcomes in men with cognitive impairment have not been studied.

E.5. Testosterone Effects on Mood, Energy, and Health-Related Quality of Life   Circulating concentrations of testosterone have not been consistently associated with mood indices and depressive symptoms in older men and in men with chronic illnesses (102-107). In intervention trials in eugonadal men, testosterone replacement did not have a significant effect on mood (274); in hypogonadal men, some studies have shown an effect whereas others have not. In an open-label trial, androgens improved positive aspects of mood and reduced negative aspects of mood in young, hypogonadal men (275).

In general, androgen deficiency does not appear to be an important factor in the pathophysiology of major depression. Placebo-controlled trials of testosterone in men with refractory depression have not consistently shown a beneficial effect of testosterone (275-278).  A meta-analysis of randomized clinical trials did not reveal a clinically meaningful effect of testosterone on depression (5).

In HIV-infected men with low testosterone levels, testosterone supplementation was more effective than placebo in restoring libido and energy, and alleviating depressed mood (279-280). The depression scores in HIV-infected men were increased in association with hypogonadism in men with AIDS wasting, and administration of testosterone resulted in a significant improvement in depression inventory score (279).

There is anecdotal evidence that androgens improved energy and reduced sense of fatigue (280). Testosterone administration increases hemoglobin and red cell mass, stimulates 2, 3 BPG concentrations thereby shifting the oxygen – hemoglobin dissociation curve favorably to improve greater oxygen delivery, and induces muscle capillarity (281-283). Additionally, testosterone stimulates mitochondrial biogenesis and mitochondrial quality (284). All of these adaptations would be expected to improve net oxygen delivery to the muscle, improve aerobic performance and reduce fatigability. However, the effects of testosterone on fatigue have not been studied in randomized trials.

Supraphysiologic doses of androgenic steroids such as those abused by athletes and recreational body builders have been associated with aggressive responses to provocative situations (285), increased scores on Young’s manic scale, and with affective and psychotic disorders in some individuals (286); these adverse effects have not been reported with physiologic testosterone replacement.

Physical, sexual, and cognitive functions are important determinants of the health-related quality of life . For instance, in HIV-infected individuals, health-related quality of life correlated significantly with lean body mass (287). Cognitive function is an important determinant of an individual’s ability to live independently. By improving some aspects of physical, sexual, and cognitive functions, testosterone supplementation might be expected to improve health-related quality of life. However, only a few small trials have evaluated the effects of testosterone on health-related quality of life. A systematic review of a small number of randomized trials has not revealed a significant improvement in composite health-related quality of life scores, but testosterone therapy has been shown to improve scores on the physical function domain of SF-36 (5, 137).

F.         Considerations in Testosterone Therapy of Older Men with Low Testosterone Levels

The risks and benefits of long-term testosterone therapy on health-related outcomes in older men with symptomatic conditions associated with low testosterone levels are unknown. Recognizing the lack of evidence of the safety and effectiveness of testosterone therapy in older men with symptomatic androgen deficiency, the expert panel of the Endocrine Society recommended against testosterone therapy of all older men with low testosterone levels (5). Instead the panel suggested that “clinicians consider offering testosterone therapy on an individualized basis to older with consistently low testosterone levels on more than one occasion and significant symptoms of androgen deficiency, after appropriate discussion of the uncertainties of the risks and benefits of testosterone therapy in older men” (5). The panel’s recommendations were guided by the recognition of the paucity and low quality of evidence, and by the sober realization that high quality evidence of the efficacy and safety will not be available for a very long time.

Although the prevalence of low testosterone levels in older men is arguably high, the usefulness of population screening cannot be evaluated for several reasons. Because of the lack of agreement on a case definition, the paucity of data on the performance characteristics of the screening instruments (e.g., the ADAM questionnaire (289), the Aging Male Symptoms questionnaire (290), and the MMAS questionnaire (291) and the lack of clarity on the public health impact of the androgen deficiency syndrome in the general population, screening of all older men for androgen deficiency is not justified.

Prior to prescribing testosterone therapy, a careful general health evaluation is necessary to identify any potential conditions that might increase the risk of testosterone therapy. The contraindications to testosterone therapy are listed in Table 1. Also, an explicit discussion of the uncertainties about the benefits and risks of testosterone therapy should precede prescription of testosterone therapy. Men receiving testosterone therapy should be monitored using a standardized monitoring plan to facilitate early detection of adverse events and to minimize the risk of unnecessary prostate biopsies (Table 2), as recommended by the Endocrine Society expert panel (Table 3).

The clinical pharmacology of the available testosterone formulations is summarized in Table 4.  Testosterone therapy can be instituted using any of the available approved formulations based on considerations of pharmacokinetics, patient convenience and preference, cost, and formulation-specific adverse effects. Suggestions for initial treatment regimens are provided in Table 5 with the caveat that dose and regimen should be adjusted based on measurement of serum testosterone levels after initiation of therapy. The aim should be to raise testosterone levels into the mid-normal range for healthy young men.

  1. G.           Risks of Testosterone Administration in Older Men

Short-term testosterone administration is associated with a low frequency of relatively mild adverse effects such as acne, oiliness of skin, and breast tenderness in healthy, young, androgen-deficient men with classical hypogonadism. However, the long term risks of testosterone supplementation in older men are largely unknown. There are several unique considerations in older men that may increase their risks of testosterone administration. Serum total and free testosterone concentrations are higher in older men than young men at any dose of testosterone therapy, presumably due to decreased testosterone clearance (40). Altered responsiveness of older men to testosterone administration might make them susceptible to a higher frequency of adverse events, such as erythrocytosis, or to unique adverse events not observed in young hypogonadal men. The baseline prevalence of disorders such as prostate cancer, benign prostatic hypertrophy, and cardiovascular disease that might be exacerbated by testosterone administration is high in older men; therefore, small changes in risk in either direction could have enormous public health impact. Furthermore, the clustering of co-morbid conditions in the frail elderly might render these men more susceptible to the adverse effects of testosterone therapy than healthy young hypogonadal men.

The contraindications for testosterone administration include history of prostate or breast cancer (Table 1). Benign prostatic hypertrophy is by itself not a contraindication, unless it is associated with severe symptoms, as indicated by IPSS symptom score of greater than 19. Testosterone should not be given without prior evaluation and treatment to men with baseline hematocrit greater than 50%, severe untreated sleep apnea, or congestive heart failure with Class III or IV symptoms (5).

The risks of testosterone administration include acne, oiliness of skin, erythrocytosis, induction or exacerbation of sleep apnea, leg edema, and breast tenderness or enlargement (5) (Table 2). Abnormalities of liver enzymes, hepatic neoplasms, and peliosis hepatis that have been reported previously with orally administered, 17-alpha alkylated androgens, have not been observed with replacement doses of parenterally administered testosterone formulations. The two major areas of concern and uncertainty are the effects of long-term testosterone administration on prostate cancer and cardiovascular events.

F.1. Testosterone Effects on the Risk of Atherosclerotic Heart Disease   The long-term consequences of testosterone supplementation on the risk of heart disease remain unknown and have been the subject of contentious debate. (123, 292-294).

F.1.a. Androgen Effects on Plasma Lipids.  Cross-sectional studies of middle-aged men found a positive relationship between serum testosterone levels and plasma HDL-cholesterol concentrations (294-296). Lower testosterone levels in men are associated with higher levels of dense LDL particles (295) and prothrombotic factors (297).

The effects of androgen supplementation on plasma lipids depend on the dose, the route of administration (oral or parenteral), the type of androgen (aromatizable or not) and the subject population (whether young or old, and hypogonadal or not). Supraphysiological doses of testosterone and non-aromatizable androgens frequently employed by body-builders undoubtedly decrease plasma HDL-cholesterol levels (298-301). However, administration of replacement doses of testosterone in older men has been associated with only a modest or no decrease in plasma HDL-cholesterol (5, 156-162).

It has been suggested that the decrease in HDL cholesterol with testosterone administration might be the result of increased cholesterol efflux from endothelial macrophages stimulating reverse cholesterol transport, and therefore, a beneficial effect, rather than the result of increased HDL catabolism (302).

F.1.c. Androgens and Other Cardiovascular Risk Factors.  Cross-sectional studies have found a positive association between circulating testosterone concentrations and tissue plasminogen activator activity (303), and a negative relationship between testosterone and plasminogen activator inhibitor-1 activity, fibrinogen, and some other prothrombotic factors (303), suggesting an antithrombotic effect of testosterone. However, intervention trials of testosterone or hCG administration generally have not found a significant effect of testosterone on inflammatory markers (304). Similarly, in another study, even supraphysiological doses of testosterone did not affect C-reactive protein (305).

F.1.d. Androgens and Coronary Artery Disease.  Whether variation of testosterone within the normal range is associated with risk of coronary artery disease remains controversial. Of the 30 cross-sectional studies reviewed by Alexandersen (123), 18 reported lower testosterone levels in men with coronary heart disease, 11 found similar testosterone levels in controls and men with coronary artery disease and 1 found higher levels of DHEAS. Prospective studies have failed to reveal an association of total testosterone levels and coronary artery disease (124-128, 306-308). The Rotterdam Study found that the common carotid artery intimal media thickness, a marker of generalized atherosclerosis, was the highest in older men in the lowest quartile of serum testosterone levels (128).

One interventional study (309), reported that testosterone undecanoate given orally improved angina pectoris in men with coronary heart disease. Testosterone infusion acutely improves coronary blood flow in a canine model and in men with coronary artery disease (310-316). Short-term administration of testosterone induces a beneficial effect on exercise-induced myocardial ischemia in men with coronary artery disease (315). This effect may be related to a direct coronary-relaxing effect. Testosterone replacement has been shown to increase the time to 1-mm ST-segment depression (313). However, in another study, there were no differences among the placebo or testosterone groups in peak heart rate, systolic blood pressure, maximal rate pressure product, perfusion imaging scores, or the onset of ST-segment depression (315). Studies by Yue et al (316) demonstrated testosterone-induced endothelium independent relaxation of rabbit coronary arteries via potassium conductance.

F.1.f. Effects of testosterone supplementation on atherosclerosis progression in animal models of atherogenesis. In a mouse model of atherosclerosis that is LDL-receptor deficient (317) surgical castration accelerated, and testosterone administration retarded the progression of atherosclerosis. The magnitude of testosterone effect on atherosclerosis progression is similar to that observed with estrogen administration. Favorable effects of testosterone on atherosclerosis in this mouse model are antagonized by concomitant administration of an aromatase inhibitor, suggesting that testosterone effects are possibly mediated through its conversion to estrogen in the vessel wall (317). Testosterone effects in retarding atherosclerosis progression were independent of plasma lipids (317). Many, though not all the studies in cholesterol-fed, castrated male rabbits are in agreement that testosterone does not promote atherogenesis (318). Taken together, these data provide evidence that testosterone, through its conversion to estradiol, can retard the progression of atherosclerosis in these animal models.

F.1.g The Effects of Testosterone on Cardiovascular Events. To-date, no randomized trials on the effects of testosterone on cardiovascular-related events have been published (167). Therefore, the published data have been derived necessarily from the analyses of the reported adverse events in randomized clinical trials. The number of cardiovascular-related events reported in randomized testosterone trials has been strikingly low—even lower than that expected for the age and comorbid conditions of the participants (167, 319-320).A randomized trial of testosterone in older men (The TOM Trial) with mobility limitation was stopped early due to a higher frequency of cardiovascular-related events in men assigned to testosterone than in those assigned to placebo (164), heightening concern about the cardiovascular safety of testosterone in frail older men.  In contrast to many other testosterone trials in older men, which recruited relatively healthy older men, the participants in the TOM trial had a high prevalence of chronic conditions, such as heart disease, diabetes mellitus, obesity, hypertension, and hyperlipidaemia (164). Men, 75 years of age or older, and men with higher on-treatment testosterone levels seemed to be at the greatest risk of cardiovascular-related events. The dose of testosterone used in the TOM trial was higher than that used in some previous trials, but not dissimilar from or lower than that used in some other trials.

Several meta-analyses of randomized testosterone trials have been conducted (292, 319-320). However, these meta-analyses are limited by the small size of most trials, heterogeneity of study populations, poor quality of adverse-event reporting, and short treatment duration in many trials. None of the testosterone trials to date was sufficiently powered to adequately assess safety outcomes. The rigor of adverse-event reporting varied greatly among studies. The most recent meta-analysis of randomized testosterone trials included 2,994 men from 27 eligible trials of 12 weeks or longer duration.  Randomization to testosterone was associated with an increased the risk of a cardiovascular-related event (odds ratio (OR) 1.54, 95% confidence interval (CI) 1.09 to 2.18) (Figure 9) (320). A remarkable finding of this meta-analysis was that the effect of testosterone therapy varied with the source of the trial’s funding (320). The risk of a cardiovascular-related event on testosterone therapy was even greater (OR 2.06, 95% CI 1.34 to 3.17) in trials that were not funded by the pharmaceutical industry; in contrast, the trials funded by the pharmaceutical industry did not reveal a significant increase in cardiovascular events. Vigen et al (321) assessed the association between testosterone therapy and all-cause mortality, myocardial infarction (MI), or stroke among male veterans with low testosterone levels (<300 ng/dL) who underwent coronary angiography in the Veterans Affairs (VA) system between 2005 and 2011. After adjusting for the presence of coronary artery disease, testosterone therapy was associated with increased risk of adverse outcomes (all-cause mortality, myocardial infarction or stroke) (hazard ratio, 1.29; 95% CI, 1.04 to 1.58). It should be noted that a separate retrospective analysis of men in the Veterans Affairs Health Care System reported reduced overall mortality in men receiving testosterone (322).

The Hormonal Regulators of Muscle and Metabolism in Aging (HORMA) trial reported a significantly greater increase in blood pressure in men treated with testosterone than in those treated with placebo (323). Testosterone administration causes salt and water retention, which can induce edema and worsen pre-existing heart failure. Thus, large prospective randomized trials are needed to determine the effects of testosterone therapy on cardiovascular health.

F.1.h Synopsis of the Effects of Testosterone on Cardiovascular risk.  The cohort and cross-sectional studies collectively suggest a neutral or favorable effect of testosterone on coronary heart disease in men, although the evidence is far from conclusive. It is possible that frail elderly men with high burden of chronic diseases and cardiovascular risk factors may be at increased risk of cardiovascular-related adverse events (167). Long term randomized trials of the effects of testosterone replacement on cardiovascular events are needed and are particularly important because even small changes in incidence rates could have significant public health impact.

F.1.i. Testosterone, Diabetes, and Metabolic Syndrome

Spontaneous (134) and experimentally induced (324) androgen deficiency is associated with increased fat mass, and testosterone replacement decreased fat mass in older men with low testosterone levels (5). In epidemiologic studies, low testosterone levels are associated with higher levels of abdominal adiposity (325-326). Testosterone administration promotes the mobilization of triglycerides from the abdominal adipose tissue in middle-aged men (327). Surgical castration in rats impairs insulin sensitivity; physiologic testosterone replacement reverses this metabolic derangement (328). However, high doses of testosterone impair insulin sensitivity in castrated rats (328), suggesting a biphasic relationship in which both low and high testosterone levels impair insulin resistance. Androgens increase insulin-independent glucose uptake (329) and modulate LPL activity in a region-specific manner (330).

Testosterone levels are lower in men with type 2 diabetes mellitus compared with controls (331-335). Low total testosterone levels have been associated consistently with increased risk of type 2 diabetes mellitus and metabolic syndrome in community dwelling men both cross-sectionally and longitudinally (336-343).  However, the association of free testosterone and type 2 diabetes mellitus has been inconsistent; some studies have reported a weak relationship (336-337, 341) while others have failed to find any relationship (338, 340).  Circulating sex hormone binding globulin (SHBG) and some SHBG polymorphisms also have been associated negatively with the risk of type 2 diabetes (336-346).  For instance, individuals with the rs6257,rs179994, and rs6259 variants alleles of the SHBG single nucleotide polymorphism (SNP) have lower plasma SHBG levels and a higher risk of type 2 diabetes (342, 344-346). As total testosterone and SHBG levels are highly correlated, we determined whether SHBG is an independent predictor of T2DM. Accordingly, we performed longitudinal analyses of men participating in the Massachusetts Male Aging Study (347), a population-based study of men aged 40-70 years (Figure 10). After adjustment for age, body mass index, hypertension, smoking, alcohol intake and physical activity, the hazard ratio (HR) for incident type 2 diabetes was 2.0 for each one SD decrease in SHBG and 1.29 for each one SD decrease in total testosterone (347). Free testosterone was not significantly associated with type 2 diabetes. The strong association of T2DM risk with SHBG persisted even after additional adjustment for free testosterone. Thus, SHBG, but not free testosterone, is an independent predictor of incident type 2 diabetes. Although it is possible that SHBG is a marker of insulin resistance, and low SHBG levels reflect the effects of hyperglycemia or insulin resistance, the association of SHBG polymorphisms with type 2 diabetes suggests an important mechanistic role of SHBG in the pathogenesis of type 2 diabetes.

Interventional trials have yielded inconsistent results. Acute and severe androgen deficiency induced by administration of a GnRH agonist or antagonist worsens measures of insulin sensitivity. Thus, acute withdrawal of testosterone therapy in men with idiopathic hypogonadotropic hypogonadism (348) and administration of androgen deprivation therapy in men with prostate cancer (349) is associated with the development of insulin resistance. The men with prostate cancer who are receiving androgen deprivation therapy are at increased risk of the type 2 diabetes (349). Although several randomized testosterone trials have been conducted in men with type 2 diabetes mellitus, only the results of one such trial have been published (350). In the TIMES2 study (350), the men with type 2 diabetes mellitus and/or metabolic syndrome were randomized to either 2% testosterone gel or placebo gel for 6 months. Randomization to testosterone arm was associated with greater improvements in sexual function and plasma lipid levels than placebo (350). However, the changes in HbA1c levels did not differ between groups (350). Homeostasis model assessment of insulin resistance (HOMA-IR), a marker of insulin resistance, improved modestly in men who were assigned to testosterone compared with placebo (350). Overall, this and other unpublished studies have failed to show improvements in diabetes outcomes or consistent changes in measures of insulin sensitivity (167, 350-353) even though interventional trials have found a consistent reduction in whole body fat as well as abdominal fat (167, 353, 354).

F.2. Testosterone and Prostate Cancer Risk   There is no evidence that testosterone administration causes prostate cancer. Also, there is no consistent relationship between endogenous serum testosterone levels and the risk of prostate cancer (5, 355-356). However, there are a number of areas of concern that are discussed below. Prostate cancer is a common, androgen–dependent tumor, and androgen administration may promote the growth of a pre-existing prostate cancer (356-357). Testosterone administration is absolutely contraindicated in men with history of prostate cancer (5, 355). The prevalence of subclinical, microscopic foci of prostate cancer in older men is high (358-366). There is concern that testosterone administration might make these subclinical foci of cancer grow and become clinically overt. In addition, older men with low testosterone levels may have prostate cancer (367-368). Morgentaler et al (367-368) reported a high prevalence of biopsy-detectable prostate cancer in men with low total or free testosterone levels despite normal PSA levels and normal digital rectal examinations. However, this study did not have a control group, and we do not know whether sextant biopsies of age-matched controls with normal testosterone levels would yield a similarly high incidence of biopsy-detectable cancer. Therefore, this study should not be interpreted to conclude that there is a higher prevalence of prostate cancer in older men with low testosterone levels, or that low testosterone levels are an indication for performing prostate biopsy.

F.2.a. Androgen Levels and Prostate Cancer Risk: Data from Cross-sectional Studies. Overall, in cross-sectional, epidemiological studies, there has not been a consistent association between circulating androgen levels and the occurrence of prostate cancer (369-388). While one meta-analysis found no association between serum testosterone levels and prostate cancer (374), another found a slightly increased risk of prostate cancer in men with the highest testosterone levels (382). A recent meta-analysis of epidemiologic studies concluded that there is no consistent relationship between endogenous testosterone levels and the risk of prostate cancer (356).

F.2.c. Effects of Testosterone Therapy on Prostate Events. None of the testosterone trials in middle-aged or older men has had sufficient power to detect meaningful differences in prostate event rates between testosterone and placebo-treated men. A systematic review of randomized testosterone trials in middle-aged and older men found higher rates of prostate events in testosterone-treated men than in placebo-treated men (Figure 11 (292). Men treated with testosterone in these trials were at significantly higher risk for undergoing prostate biopsy than placebo-treated men (292). Because of the high prevalence of subclinical prostate cancer in older men, the higher number of prostate biopsies in testosterone-treated men is likely to yield higher detection rates of prostate cancer in comparison with placebo-treated men. Thus, testosterone therapy of middle-aged and older men is associated with a higher risk of prostate biopsy and a bias towards detection of a higher number of prostate events (167, 292).

Administration of exogenous testosterone or suppression of circulating levels of testosterone by administration of a GnRH antagonist is not associated with proportionate changes in intra-prostatic testosterone or DHT concentrations. For instance, in arandomized controlled trial, Marks et al (389) measured intraprostatic testosterone and DHT levels in older men treated with placebo or testosterone. Surprisingly, intraprostatic DHT concentrations were not significantly higher in testosterone-treated men than in placebo-treated men (389). Similarly, the expression levels of androgen-dependent genes in the prostate were not significantly altered by testosterone administration (389). In separate studies, lowering of circulating testosterone levels by administration of a GnRH antagonist was not associated with changes in intraprostatic androgen concentrations (390-391).

F.2.b. Effects of Testosterone Replacement on Serum PSA Levels. Serum PSA levels are lower in androgen–deficient men and are restored to normal following testosterone replacement (5, 392-400). Lowering of serum testosterone concentrations by withdrawal of androgen therapy in young, hypogonadal men is associated with a decrease in serum PSA levels. Similarly, treatment of men with benign prostatic hyperplasia with a 5-alpha reductase inhibitor, finasteride, is associated with a significant lowering of serum and prostatic PSA levels (400-401). Conversely, testosterone supplementation increases PSA levels (393-400).  However, serum PSA levels do not increase progressively in healthy hypogonadal men with replacement doses of testosterone. Placebo–controlled trials of testosterone administration in older men have reported either minimal increase or no significant change in serum PSA levels in testosterone–treated men (157-158).  The increase in PSA levels during testosterone replacement might trigger evaluation and biopsy in some patients (5, 355).

More intensive PSA screening and follow-up of men receiving testosterone replacement might lead to an increased number of prostate biopsies and the detection of subclinical prostate cancers that would have otherwise remained undetected (5, 355).Serum PSA levels tend to fluctuate when measured repeatedly in the same individual over time (402-404). When serum PSA levels in androgen deficient men on testosterone replacement therapy show a change from a previously measured value, the clinician has to decide whether the change warrants detailed evaluation of the patient for prostate cancer, or whether it is simply due to test–to–test variability in PSA measurement. Therefore, it is important to set criteria for monitoring PSA changes during testosterone supplementation. Criteria that use very low thresholds for performing prostate biopsy relative to test-retest variability will likely result in an excessive number of biopsies with their associated costs, psychological trauma, and morbidity. On the other hand, criteria that use unreasonably high thresholds for performing prostate biopsies may fail to detect clinical prostate cancers at an early stage.

            There is considerable test-retest variability in PSA measurements (402-404). Some of this variability is due to the inherent assay variability, and a significant portion of this variability is due to unknown factors. Fluctuations are larger in men with high mean PSA levels. Variability can be even greater if measurements are performed in different laboratories that use dissimilar assay methodology (402-404).

From a clinical perspective, an important issue is what increment in PSA level should warrant a prostate biopsy in older men receiving testosterone replacement. To address this issue, we conducted a systematic review of published studies of testosterone replacement in hypogonadal men (355). This review indicated that the weighted effect size of the change in PSA after testosterone replacement in young, hypogonadal men is 0.68 standard deviation units (95% confidence interval 0.55 to 0.82). This means that the effect of testosterone replacement therapy is to increase PSA levels by an average 0.68 standard deviations over baseline. Because the average standard deviation was 0.47 in this systematic analysis, the standard deviation score of 0.68 translates into an average increase in serum PSA levels of about 0.30 ng/ml in young hypogonadal men (355). There is considerable variability in the magnitude of change in PSA after testosterone supplementation among these studies, in part due to heterogeneity of study populations, inclusion of older men in some studies but not others, and differences in PSA assays. In addition, many patients who were enrolled in these studies were likely receiving testosterone replacement therapy previously; we do not know whether the washout period was sufficient to return PSA levels to baseline. Therefore, it is possible that because of inadequate washout, the increments in serum PSA levels after testosterone administration might have been under-estimated.

We performed a separate systematic review of data from placebo-controlled trials of testosterone supplementation in older men with low or low normal testosterone concentrations (355). The weighted effect size in six studies of older men was 1.48 standard deviation units, with a 95% confidence interval of 1.21 to 1.75. Thus, on average, older men experience a greater increase in serum PSA concentrations than younger men. The average effect of testosterone replacement in older men is to increase PSA levels by almost 1.5 standard deviations over baseline. There is, however, significant variability in the results among these six studies (p<0.0001), and the average standard deviation was skewed by one study, which had a very high standard deviation (355). After excluding this study, the average change in serum PSA levels after testosterone replacement in studies of older men was 0.43 ng/mL.

The data from the Proscar Long-Term Efficacy and Safety Study (PLESS) demonstrated that the 90% confidence interval for the change in PSA values measured 3 to 6 months apart is 1.4 ng/mL (400). Therefore, a change in PSA of >1.4 ng/ml between any two values measured 3 to 6 month apart in the same patient should be verified by a repeat PSA measurement (5, 355). If the repeated measurement confirms a change of >1.4 ng/ml from the previous value, then that patient should be referred for Urologic evaluation.

Carter et al, based on the analysis of PSA data from the Baltimore Longitudinal Study of Aging, reported that PSA velocity, defined as the annual rate of change of PSA, is different in men who develop prostate cancer than in those who do not (405-407). Thus, PSA velocity greater than 0.7 ng/ml/year was unusual in men without prostate cancer whose baseline PSA was between 4 and 10 ng/ml (405-407). However, most men being considered for testosterone replacement will have baseline PSA less than 4 ng/ml. In a subsequent analysis, the same group reported that the PSA velocity in men with baseline PSA between 2 and 4 ng/ml was 0.2 ng/ml/year (407). Because test-to-retest variability in PSA measurement is far greater than this threshold, it is likely that the use of this threshold of 0.2 ng/ml/year to select men for prostate biopsy would lead to many unnecessary biopsies. These considerations of interassay variability and the longitudinal change in PSA led an Endocrine Society Expert Panel to suggest that in men receiving testosterone replacement, a PSA velocity of greater than 0.4 ng/ml/year in men whose baseline PSA is less than 4 ng/ml should lead to urological evaluation (5).  Carter et al have emphasized that PSA velocity should not be used for data of less than 2 years duration (405-408).

In eugonadal, young men, administration of supraphysiological doses of testosterone does not further increase serum PSA levels (140, 143, 409). These data are consistent with dose response studies in young men that demonstrate that maximal serum concentrations of PSA are achieved at testosterone levels that are at the lower end of the normal male range; higher testosterone concentrations are not associated with higher PSA levels (140, 143).

In summary, these data suggest that the administration of replacement doses of testosterone to androgen-deficient men can be expected to produce a modest increment in serum PSA levels. Increments in PSA levels after testosterone supplementation in androgen-deficient men are generally less than 0.5 ng/mL, and increments in excess of 1.0 ng/mL over a 3-6 month period are unusual. Nevertheless, administration of testosterone to men with baseline PSA levels between 2.5 and 4.0 ng/mL will cause PSA levels to exceed 4.0 ng/mL in some men. Increments in PSA levels above 4 ng/mL will trigger a urological consultation and many of these men will be asked to undergo prostate biopsies.

F.2.c. Monitoring PSA Levels in Older Men Receiving Testosterone Replacement (Tables 3 and 6) Older men considering testosterone supplementation should undergo digital examination of the prostate, evaluation of risk factors for prostate cancer, and a baseline PSA measurement (5). Men with history of prostate cancer should not be given androgen supplementation, and those with palpable abnormalities of the prostate or PSA levels greater than 3 ng/ml should undergo urological evaluation. After initiation of testosterone replacement therapy, PSA levels and digital examination of the prostate should be repeated at 3, 6, and 12 months, and annually thereafter (5). In patients in whom sequential PSA measurements are available for more than two years, PSA velocity criterion can be useful in evaluating change in PSA levels. A PSA velocity of greater than 0.4 ng/ml/year in men with baseline PSA less than 3 ng/ml should be evaluated (Table 6). Although measurements of free PSA and PSA density have been proposed to enhance the specificity of PSA measurement, long term data, especially from studies of testosterone replacement in older men, are lacking.

F.2.d. Testosterone and Benign Prostatic Hypertrophy   Testosterone replacement can be administered safely to men with benign prostatic hypertrophy who have mild to moderate symptom scores. The severity of symptoms associated with benign prostatic hypertrophy can be assessed by using either the International Prostate Symptom Score (IPSS) or the American Urological Association (AUA) Symptom questionnaires. Androgen deficiency is associated with decreased prostate volume and androgen replacement increases prostate volumes to those in age–matched controls (389, 392, 396-397). In patients with pre–existing, severe symptoms of benign prostatic hypertrophy, even small increases in prostate volume during testosterone administration may exacerbate obstructive symptoms. In these men, testosterone should either not be administered or administered with careful monitoring of obstructive symptoms.

F.3. Erythrocytosis   Testosterone replacement is associated with increased red cell mass and hemoglobin levels (Figure 9) (410-414). Therefore, testosterone replacement should not be administered to men with baseline hematocrit of 50% or greater without appropriate evaluation and treatment of erythrocytosis (5) (Table 3). Administration of testosterone to androgen–deficient young men is typically associated with a small increase in hemoglobin levels. Clinically significant erythrocytosis is uncommon in young hypogonadal men during testosterone replacement therapy, but can occur in men with sleep apnea, significant smoking history, or chronic obstructive lung disease. Testosterone administration in older men is associated with more variable and somewhat greater increments in hemoglobin than observed in young, hypogonadal men (415). The magnitude of hemoglobin increase during testosterone therapy appears related to the testosterone dose, the increase in testosterone concentrations during testosterone therapy, and age (415). Testosterone replacement by means of a transdermal system has been reported to produce a lesser increase in hemoglobin levels than that associated with testosterone esters (416).

Testosterone increases hemoglobin and hematocrit by multiple mechanisms (417, 281-283). Testosterone administration stimulates erythropoiesis, suppresses hepcidin transcription by blocking BMP signaling, and increases iron availability for erythropoiesis (417, 281-283). Additionally, testosterone appears to alter the set-point of the relationship between erythropoietin and hemoglobin (281). Testosterone supplementation can correct anemia in older men with anemia of aging and in older mice (281, 283).

F.4. Monitoring Hematocrit During Testosterone Replacement Therapy (Table 3) Hemoglobin levels should be measured at baseline, and 3 and 6 months after institution of testosterone replacement, and every 6 months thereafter. It is not clear what absolute hematocrit level or magnitude of change in hematocrit warrants discontinuation of testosterone administration. Plasma viscosity increases disproportionately as hematocrit rises above 50%. Hematocrit levels above 54% are also associated with increased risk of stroke. Therefore, testosterone dose should be withheld if hematocrit rises above 54%; once hematocrit falls to a safe level, testosterone therapy may be re-initiated  at a reduced dose (5). Consideration should also be given to switching to a transdermal system, if the men are receiving injectable esters. Periodic phlebotomy is also a reasonable option in men in whom hematocrit rises above this threshold during testosterone supplementation.

F.5. Sleep Apnea   Circulating testosterone concentrations are related to sleep rhythm and are generally higher during sleep than during waking hours (418-421). Testosterone secretory peaks coincide with the onset of rapid-eye movement sleep. Aging is associated with decreased sleep efficiency, reduced numbers of REM sleep episodes, and altered REM sleep latency, which may contribute to lower circulating testosterone concentrations (419-423).  The degree of sleep-disordered breathing increases with age, and is associated with reduced overnight plasma bioavailable testosterone.  Thus, changes in sleep efficiency and architecture are associated with alterations in testosterone levels in older men (419-423).

Testosterone can induce or exacerbate sleep apnea in some individuals, particularly those with obesity or chronic obstructive lung disease(418-424). This appears to be due to direct effects of testosterone on laryngeal muscles. However, the occurrence of sleep apnea, de novo, in healthy older men treated with physiologic testosterone replacement, is very infrequent.

In men with obesity and obstructive sleep apnoea, testosterone administration has been reported to worsen sleep-disordered breathing (421). Testosterone administration depresses hypercapnoeic ventilator drive and induces apnoea in primate infants (423). Short-term administration of high doses of testosterone shortens sleep duration and worsens sleep apnoea in older men (425) The frequency of sleep apnoea in randomized testosterone trials in older men has been very low (5, 319). Obstructive sleep apnoea is often associated with low testosterone levels (426).

Testosterone should not be given to men with severe obstructive sleep apnea without evaluation and treatment of sleep apnea. Several screening instruments can be used to detect sleep apnea. A history of loud snoring, and daytime somnolence, in an obese individual with hypertension increases the likelihood of sleep apnea.

F.6. Breast Enlargement  Testosterone administration can induce breast enlargement due to testosterone conversion to estradiol although this is an uncommon complication. Even with administration of supraphysiological doses of testosterone enanthate, less than 4% of men in a contraceptive trial developed detectable breast enlargement (410). Breast cancer is listed as a contraindication for testosterone replacement therapy primarily because of concern that increased estrogen levels during testosterone treatment might exacerbate breast cancer growth. There are, however, few case reports of breast cancer occurring as a complication of testosterone treatment. Men with Klinefelter’s syndrome have a higher risk of breast cancer than general population ().

 

CHANGES IN THE GAMETOGENIC COMPARTMENT OF THE TESTIS

Women are more fertile below the age of 40, and fertility ceases at the inception of menopause, around age 50.  Increasing age in women confers greater risk for infertility, spontaneous abortion, and genetic and chromosomal defects among offspring. In contrast, there is no critical age at which sperm production or function, and fertility cease in men (427-434).  Although serum testosterone levels decrease below the normal range in a significant minority of older men, men over the age of 60 years commonly father children; the oldest father on record was 94-years old (427,429). Even though many older men are fertile, the overall fertility and fecundity declines with aging. The interpretability of data on the effects of aging on male fertility is limited by the small size of the studies and the low overall event rates.

Although there is a positive association between paternal age and incidence of aneuploidy, it has been difficult to dissociate the effect of paternal age from the confounding influence of the advanced maternal age. After accounting for various confounders, there does not appear to be a major independent effect of increased paternal age on the incidence of autosomal aneuploidies (428-429, 436-437, 439-440).The existence of a paternal age effect on Down syndrome is controversial.  Earlier studies from the 1960s and 1970s found no correlation between Down syndrome and paternal age (e.g., 441). However, a study in New York from 1983 to 1997 found a significant greater numbers of mothers and fathers 35 years of age and older, respectively, among parents of patients with Down’s syndrome (442). Among the cases of Down syndrome evaluated, paternal age had a significant effect only in mothers 35 years of age or older, and was the greatest in couples greater than 40 years of age where the risk was 6 times the rate of couples younger than 35 years of age (442).

Approximately one third of babies with diseases due to new autosomal dominant mutations are fathered by men aged 40 years or older (443).Paternal age has been associated with a significant increase in the risk of germ line mutations in FGFR2, FGFR3, and RET genes and inherited autosomal dominant diseases, such as Apert's syndrome, achondroplasia, and Costello Syndrome, respectively, in the offspring of older men (434, 439-440, 444-450). These monogenic disorders have been referred to as paternal age effect disorders (PAE).  Some other disorders such as schizophrenia and autism have also been linked to paternal age (439-440, 449-450). The rate of de novo mutations increases with paternal age (449), which may contribute to the increase risk of neurodevelopmental diseases such as schizophrenia and autism (449). The accumulation of these de novo germ line mutations with increasing  paternal age has been explained by the “selfish spermatogonial selection" hypothesis (445-446).  According to this hypothesis, the somatic mutations in male germ cells that enhance the proliferation of germ cells could lead to within-testis expansion of mutant clonal lines (447-448), thus favoring the propagation of germ cells carrying these pathogenic mutations, and increasing the risk of mutations in the offspring of older fathers (447-448).Interestingly, the risk of autism has also been associated with the age of the father as well as the grandparent (450). These concerns have prompted the American Society of Reproductive Medicine to state in their guidelines that semen donors should be younger than 50 years of age so that potential hazards related to aging are diminished (430).  More recently, the guidelines lowered the age limit of semen donors to 40 years.

Some cardiac defects have also been attributed to aberrant genetic input from older men.  For instance, a case-control study of 4,110 individuals with congenital heart defects born between 1952 and 1973 in British Columbia, found a general pattern of increasing risk with increasing paternal age among cases relative to controls for ventricular septal defects, atrial septal defects and patent ductus arteriosus(443-444).  The risk of schizophrenia has also been reported to increase with paternal age (445) and possible loci affecting this risk have been identified (451). In addition, a modest proportion of preeclampsia, normally associated with increased maternal risk factors including age, might be attributable to an increase in paternal age although no gene loci have been identified (452). These observations need further corroboration.

There are no longitudinal studies in men of any age demonstrating defined changes in the reproductive tract that would explain a decline in fertility. Problems might occur at many levels.  Thus aging might affect fertility at the level of the 1) germ cell, decreasing sperm number, sperm DNA integrity, and chromosomal quality; 2) supportive cells, affecting sperm quality and number; 3) accessory glands, affecting sperm motility and function; and 4) deposition of sperm into the vagina, decreasing erectile function, ejaculation and frequency of intercourse.

  1. A.   Changes in fertility of older men     

A review of studies examining fertility at different ages demonstrated significant age-related differences in fertility rates men, including lower pregnancy rates, increased time to pregnancy, and subfecundity in men older than 50 years (428-429, 438-439, 453). Some changes in fertility rates might be related to age-related decrease in sexual activity.  A literature review found no significant change in sperm concentration with aging when comparing men under the age of 30 to those greater than 50 years (433).However, in general, semen volume, sperm motility, and the number of morphologically normal sperm decrease with advancing age (Table 7; 428-438, 449, 454). A number of these studies, however, did not control for important confounding variables.  Of the 21 studies in which sperm densities were compared among men of different age groups (433), only four studies adjusted for the duration of abstinence, well known to affect sperm concentration. In addition, there is significant heterogeneity in the populations studied; most of the studies examined data from semen of sperm donors while others examined men from infertility clinics.  Sperm donors might represent a healthier group of men than the general population; conversely men in infertility clinics might be more likely to have abnormalities of sperm number or function. Even studies that have controlled for abstinence as well as alcohol and tobacco use have shown an age-related decrease in semen volume. In one study of men whose partners had bilateral tubal obstruction or absence of both tubes and who were treated by conventional IVF, the odds ratio of failure to conceive was higher for men 40 years of age or older (455).

  1. B.   Changes in the Germ Cell Compartment

In a comparison of younger men (21-25 years)with older men (>50) referred for andrological evaluation, the ejaculate volume, progressive sperm motility, and sperm morphology were lower in older men than younger men after adjustment for duration of sexual abstinence, (456). The older men also had a higher frequency of sperm tail defects, suggesting epididymal dysfunction (457). In addition, the fructose content was significantly lower in older men suggesting a defect in the seminal vesicle contribution to semen (457). There were no significant differences in sperm concentration and testicular size between the young and older men in this study.

Necropsies on adult men of different ages have revealed that the testicular volume was lower only in men in the 8th decade of life (458). A recent study examined testicular germ cells obtained by orchidectomy from 36 older men with advanced prostate cancer and by testicular biopsy from 21 younger men with obstructive azoospermia, as controls (459). The ratios of primary spermatocytes, round spermatids, and elongated spermatids to Sertoli cells were significantly decreased in the testes of older men, but the ratio of spermatogonia to Sertoli cell number remained unchanged (459-460). Older men are characterized by lower rates of germ cell apoptosis and cell proliferation compared with younger men, suggesting that germ cell proliferation and apoptosis diminish with aging (460).

Other studies evaluating the fidelity of the germ cell compartment are cross-sectional and depend on analyses of sperm number and semen quality; large-scale chromosomal analyses in healthy community dwelling men are scarce as most data are derived from fertility clinics.  A review of studies examining semen quality at different ages demonstrated significant age-related decrease in semen volume and sperm morphology.  The change in sperm morphology has been hypothesized to be due to an increase in aneuploidy with age.  Härkönen et al (437) found that sperm morphology was directly associated with the number of chromosomes in sperm and that men with higher aneuploidy rates for chromosomes 13, 18, 21, X and Y had lower sperm motility and sperm concentrations. In spite of the changes in sperm morphology and motility from older men, in vitro fertilizing capacity of the sperm is well preserved (454-455). In some older men, degenerating germ cells can be observed suggesting loss of germ cells with age.

There are several difficulties in interpreting these data on age-related changes in sperm density and function. The normal range for sperm concentration in men is wide where sperm concentration above 15 million/ml (total sperm per ejaculate > 39 million) is considered normal. Thus, even though average sperm concentrations might decline with aging, they might still be in the normal range (453-454, 459).  Furthermore, normal sperm counts might not always correlate with normal sperm function.

Studies in flies demonstrate more germ cells during larval than adult stages suggesting age-related quiescence of the germ line (461). Significant age-related decreases in germ cells and spermatogenesis also have been reported in rodents and primates (462-466).  The Brown Norway rat has been studied as a model of aging of the human male reproductive system because in this rodent model, serum testosterone levels decrease with aging, as they do in humans (463-465). Along with changes in hypothalamic-pituitary hormones, alterations in sperm counts, sperm maturation, Sertoli cell number, and progeny outcomes have been observed in this rodent model (Table 8; 452, 463-472).  Analysis of ribosomal DNA from germ cells of the male brown Norway rat has revealed hypermethylation of ribosomal DNA(466, 473).  Alterations in ribosomes have been theorized to promote aging of cells by multiplying errors in protein synthesis which initially might elude gross morphological analysis but eventually might lead to germ cell degeneration (473).  Further assessment of spermatogonial stem cell populations is needed.  In many animal models of life span extension, there is a trade-off between longer life and fecundity, although there are some exceptions  (474).

  1. Changes in Supportive Cells and Accessory Glands

Since Sertoli and Leydig cells are crucial to spermatogenesis, changes in these cells could affect sperm number and function.   Age-related changes in the supporting structures for sperm maturation have been described in the Brown Norway rat.  These changes include reductions in the numbers of Leydig and Sertoli cells, seminiferous tubules, and in epididymal cell number and function (464-466). Changes in the supporting cells and structures for sperm maturation have been invoked to explain the age-related decrease in sperm number and fecundity in rats. In stallions, the numbers of Sertoli cells decreases with aging but individual Sertoli cells display a remarkable capacity to accommodate greater numbers of developing germ cells(475).

In men, Sertoli cell number has been reported to be lower in men aged 50 to 85 years than in men aged 20 to 48 years (476). The apoptotic rate of primary spermatocytes in aged men was also significantly elevated compared with that of younger men, resulting in a decrease of the number of primary spermatocytes per Sertoli cell (460), leading the authors to suggest that there might be a failure of the Sertoli cells to support spermatogenesis in older men.

Sertoli cells produce inhibin, which regulates gonadotropin expression from the pituitary. Inhibin B has been identified as the physiologically important form of inhibin in men and as a valuable serum marker of Sertoli cell function and spermatogenesis.  Higher gonadotropins and lower inhibin levels in older men suggest a decline in Sertoli cell function(476); however changes in circulating inhibin B levels with advancing age have been inconsistent(476-479).Overall, these data suggest a possible decline in Sertoli cell number and function in older men with little affect on spermatogenesis.

Aging is accompanied by a progressive, albeit variable, decline of Leydig cell function with a decrease of mean serum free (or bioavailable) testosterone levels in the population between age 25 and 75 years(480). Total Leydig cell volume and the absolute number of Leydig cells decline with advancing age, although total testis weight does not change substantially with age (480-484). In one study, age accounted for more than a third of the variation in Leydig cell number, and explained more than half the variation in daily sperm production (483). This might in part be explained by a fusion of Leydig cells resulting in fewer but multinucleated Leydig cells with age (484).  The functionality of the multinuicleated cells is not known.

  1. Conclusion             

In male mammals, changes at all levels of the hypothalamic-pituitary-testicular axis, including alterations in the GnRH pulse generator, gonadotropin secretion, and testicular steroidogenesis, in addition to alterations of feed-forward and feed-backrelationships contribute to an age-related decline in circulating testosterone concentrations.  The rate of age-related decline in testosterone levels is affected by the presence of chronic illness, adiposity, medication, sampling time, and the methods of testosterone measurement.  Epidemiologic surveys reveal an association of low testosterone levels with changes in body composition, physical function and mobility, risk of diabetes, metabolic syndrome, coronary artery disease, and fracture risk. Testicular morphology, semen production, and fertility are maintained up to a very old age in men. There is evidence of a small increase in the risk of specific genetic disorders among the offspring of older men.

Age-related decline in testosterone should be distinguished from classical hypogonadism due to known diseases of the hypothalamus, pituitary and the testis. In young hypogonadal men who have a known disease of the hypothalamus, pituitary and testis, testosterone therapy is generally beneficial and has been associated with a low frequency of adverse events. However, neither the benefits in improved health outcomes nor the long term risks of testosterone therapy are known in older men with age-related decline in testosterone levels.  The clinical consequences of age-related changes in circulating testosterone concentrations and epigenetic changes in sperm DNA are poorly understood.

REFERENCES
References

1.         Rossouw JE, Anderson GL, Prentice RL, LaCroix AZ, Kooperberg C, Stefanick ML, Jackson RD, Beresford SA, Howard BV, Johnson KC, Kotchen JM, Ockene J; Writing Group for the Women's Health Initiative Investigators. 2002 Risks and benefits of estrogen plus progestin in healthy postmenopausal women: principal results From the Women's Health Initiative randomized controlled trial. JAMA 288: 321-333.

  1. Hulley, S., D. Grady, T. Bush, C. Furberg, D. Herrington, B. Riggs, E. Vittinghoff. (1998). Randomized trial of estrogen plus progestin for secondary prevention of coronary heart disease in postmenopausal women. Heart and Estrogen/progestin Replacement Study (HERS) Research Group. JAMA 280(7): 605-13.
  2. Harman SM, Naftolin F, Brinton EA, Judelson DR. Is the estrogen controversy over? Deconstructing the Women's Health Initiative study: a critical evaluation of the evidence. Ann N Y Acad Sci. 2005 Jun;1052:43-56.
  3. Barrett-Connor E. Hormones and heart disease in women: the timing hypothesis. Am J Epidemiol. 2007 Sep 1;166(5):506-10.
  4. Bhasin S, Cunningham GR, Hayes FJ, Matsumoto AM, Snyder PJ, Swerdloff RS, Montori VM. Testosterone therapy in adult men with androgen deficiency syndromes: an endocrine society clinical practice guideline. J Clin Endocrinol Metab. 2010;95:2536-2544.
  5. (2002) AACE Guidelines for evaluation and treatment of hypogonadism in adult male patients - 2002 update. Endocrine Practice 8:439-456.
  6. BhasinS, Fisher CE. Male Hypogonadism. http://pier.acponline.org/physicians/diseases/d169. [Date accessed: 2003 July 18] In: PIER [online database]. Philadelphia, American College of Physicians, 2003.
  7. Wang, C. et al. Investigation, treatment, and monitoring of late-onset hypogonadism in males: ISA, ISSAM, EAU, EAA, and ASA recommendations. J Androl. 2009;30:1-9.
  8. Matsumoto, A. M. (2002). Andropause: clinical implications of the decline in serum testosterone levels with aging in men. J Gerontol A Biol Sci Med Sci 57(2): M76-99.
  9. Bhasin S, Buckwalter JG. Testosterone supplementation in older men: a rational idea whose time has not yet come. J Androl. 2001 22:718-731.
  10. Harman SM, Metter EJ, Tobin JD, Pearson J, Blackman MR; Baltimore Longitudinal Study of Aging. Longitudinal effects of aging on serum total and free testosterone levels in healthy men. Baltimore Longitudinal Study of Aging. J Clin Endocrinol Metab. 2001 86:724-731.
  11. Gray A, Feldman HA, McKinlay JB, Longcope C. Age, disease, and changing sex hormone levels in middle-aged men: results of the Massachusetts Male Aging Study. J Clin Endocrinol Metab 1991;73:1016-1025.
  12. Feldman HA, Longcope C, Derby CA, Johannes CB, Araujo AB, Coviello AD, Bremner WJ, McKinlay JB. Age trends in the level of serum testosterone and other hormones in middle-aged men: longitudinal results from the Massachusetts male aging study. J Clin Endocrinol Metab. 2002 Feb;87(2):589-98.
  13. Morley JE, Kaiser FE, Perry HM, 3rd, et al. Longitudinal changes in testosterone, luteinizing hormone, and follicle- stimulating hormone in healthy older men. Metabolism 1997; 46:410-3.
  14. Simon D, Preziosi P, Barrett-Connor E, Roger M, others a. The influence of aging on plasma sex hormones in men: the Telecom Study. Am J Epidemiol 1992;135:783-791.
  15. Zmuda JM, Cauley JA, Kriska A, Glynn NW, Gutai JP, Kuller LH. Longitudinal relation between endogenous testosterone and cardiovascular disease risk factors in middle-aged men. A 13-year follow-up of former Multiple Risk Factor Intervention Trial participants. Am J Epidemiol 1997;146:609-17.
  16. Feldman HA, Longcope C, Derby CA, Johannes CB, Araujo AB, Coviello AD, Bremner WJ, McKinlay JB. Age trends in the level of serum testosterone and other hormones in middle-aged men: longitudinal results from the Massachusetts male aging study. J Clin Endocrinol Metab. 2002 87:589-598.
  17. Ferrini RL, Barrett-Connor E. Sex hormones and age: a cross-sectional study of testosterone and estradiol and their bioavailable fractions in community-dwelling men. Am J Epidemiol. 1998 Apr 15;147(8):750-4.
  18. Dai WS, Kuller LH, LaPorte RE, Gutai JP, Falvo-Gerard L, Caggiula A. The epidemiology of plasma testosterone levels in middle-aged men. Am J Epidemiol. 1981 Dec;114(6):804-16.
  19. Pollanen P, Harkonen K, Makinen J, Irjala K, Saad F, Hubler D, Oettel M, Koskenuo M, Huhtaniemi I. Facts and fictions of andropause: lessons from the Turku male Ageing Study. 1st International Congress on Male Health in Vienna, November 2001
  20. Wu FCW. Subclinical hypogonadism. Abstract S22-3. Endocrine Scoeity Meetings, Philadelphia, PA, June 18-22, 2003.
  21. Bremner WJ, Prinz PN. A loss of circadian rhythmicity in blood testosterone levels with aging in normal men. J Clin Endocrinol Metab 1983;56:1278-1281.
  22. Harman SM, Tsitouras PD. Reproductive hormones in aging men I.  Measurement of sex steroids, basal luteinizing hormone and Leydig cell  response to human chorionic gonadotropin. J Clin Endocrinol Metab 1980;51:35-41.
  23. Harman SM, Tsitouras PD, Costa PT, Blackman MR. Reproductive hormones in aging men.  II.  Basal pituitary gonadotropins and gonadotropin responses to luteinizing hormone-releasing hormone. J Clin Endocrinol Metab 1982;54:547-551.
  24. Murono EP, Nankin HR, Lin T, Osterman J. The aging Leydig cell: VI. Response of testosterone precursors to gonadotropin in men. Acta Endocrinologica (Copenhagen) 1982; 100:455-461.
  25. Nieschlag E, Lammer U, Freischem CW, Langer K, Wickings EJ. Reproductive function in young fathers and grandfathers. J Clin Endocrinol Metab 1982;51:675-681.
  26. Tenover JS, Matsumoto AM, Plymate SR, Bremner WJ. The effects of aging in normal men on bioavailable testosterone and luteinizing hormone secretion: Response to clomiphene citrate. J Clin Endocrinol Metab 1987;65:1118-1126.
  27. Wu, F.C. et al. Identification of late-onset hypogonadism in middle-aged and elderly men. N Engl J Med 363, 123-35 (2010).
  28. Bhasin, S. et al. Reference ranges for testosterone in men generated using liquid chromatography tandem mass spectrometry in a community-based sample of healthy nonobese young men in the Framingham Heart Study and applied to three geographically distinct cohorts. J Clin Endocrinol Metab 96, 2430-9 (2011).
  29. Liverman, C.T. & Blazer, D.G. Testosterone and aging: clinical research directions (Joseph Henry Press, 2004).
  30. Orwoll E, Lambert LC, Marshall LM, Phipps K, Blank J, Barrett-Connor E, Cauley J, Ensrud K, Cummings S. Testosterone and estradiol among older men. J Clin Endocrinol Metab. 2006;91:1336-44.
  31. Zumoff B, Strain GW, Kream J, et al. Age variation of the 24 hour mean plasma concentration of androgens, estrogens, and gonadotropins in normal adult men. J Clin Endocrinol Metab 1982; 54:534-538.
  32. Pirke KM, Doerr P. Age related changes and interrelationships between plasma testosterone, oestradiol and testosterone-binding globulin in normal adult males. Acta Endocrinol (Copenh) 1973; 74:792-800.
  33. McKinlay JB, Longcope C, Gray A. The questionable physiologic and epidemiologic basis for a male climacteric syndrome: preliminary results from the Massachusetts Male Aging Study. Maturitas. 1989 Jun;11(2):103-15.
  34. Tajar, A. et al. Characteristics of androgen deficiency in late-onset hypogonadism: results from the European Male Aging Study (EMAS). J Clin Endocrinol Metab 97, 1508-16 (2012).
  35. Mohr BA, Bhasin S, Link CL, O'Donnell AB, McKinlay JB. The effect of changes in adiposity on testosterone levels in older men: longitudinal results from the Massachusetts Male Aging Study. Eur J Endocrinol. 2006 Sep;155(3):443-52.
  36. Wu, F.C. et al. Hypothalamic-pituitary-testicular axis disruptions in older men are differentially linked to age and modifiable risk factors: the European Male Aging Study. J Clin Endocrinol Metab 93, 2737-45 (2008).
  37. Sartorius G, Spasevska S, Idan A, Turner L, Forbes E, Zamojska A, Allan CA, Ly LP, Conway AJ, McLachlan RI, Handelsman DJ. Serum testosterone, dihydrotestosterone and estradiol concentrations in older men self-reporting very good health: the healthy man study. Clin Endocrinol (Oxf). 2012 Nov;77(5):755-
  38. Ohlsson C, Wallaschofski H, Lunetta KL, Stolk L, Perry JR, Koster A, Petersen AK, Eriksson J, Lehtimäki T, Huhtaniemi IT, Hammond GL, Maggio M, Coviello AD; EMAS Study Group, Ferrucci L, Heier M, Hofman A, Holliday KL, Jansson JO, Kähönen M, Karasik D, Karlsson MK, Kiel DP, Liu Y, Ljunggren O, Lorentzon M, Lyytikäinen LP, Meitinger T, Mellström D, Melzer D, Miljkovic I, Nauck M, Nilsson M, Penninx B, Pye SR, Vasan RS, Reincke M, Rivadeneira F, Tajar A, Teumer A, Uitterlinden AG, Ulloor J, Viikari J, Völker U, Völzke H, Wichmann HE, Wu TS, Zhuang WV, Ziv E, Wu FC, Raitakari O, Eriksson A, Bidlingmaier M, Harris TB, Murray A, de Jong FH, Murabito JM, Bhasin S, Vandenput L, Haring R. Genetic determinants of serum testosterone concentrations in men. PLoS Genet. 2011 Oct;7(10):e1002313. doi: 10.1371/journal.pgen.1002313.
  39. Coviello AD, Lakshman K, Mazer NA, Bhasin S. Differences in the apparent metabolic clearance rate of testosterone in young and older men with gonadotropin suppression receiving graded doses of testosterone. J Clin Endocrinol Metab 2006;91:4669-75.
  40. Ishimaru T, Pages L, Horton R. Altered metabolism of androgens in elderly men with benign prostatic hyperplasia. J Clin Endocrinol Metab. 1977 Oct;45(4):695-701.
  41. Mulligan T, Iranmanesh A, Johnson ML, Straume M, Veldhuis JD. Aging alters feed-forward and feedback linkages between LH and testosterone in healthy men. Am J Physiol 1997 Oct;273(4 Pt 2):R1407-13.
  42. Winters SJ, Sherins RJ, Troen P. The gonadotropin-suppressive activity of androgen is increased in elderly men. Metabolism. 1984 33:1052-1059.
  43. Urban RJ, Veldhuis JD, Blizzard RM, Dufau ML. Attenuated release of biologically active luteinizing hormone in healthy aging men. J Clin Invest. 1988 Apr;81(4):1020-9.
  44. Zwart AD, Urban RJ, Odell WD, Veldhuis JD. Contrasts in the gonadotropin-releasing hormone dose-response relationships for luteinizing hormone, follicle-stimulating hormone and alpha-subunit release in young versus older men: appraisal with high-specificity immunoradiometric assay and deconvolution analysis. Eur J Endocrinol. 1996 135:399-406.
  45. Gruenewald DA, Naai MA, Marck BT, Matsumoto AM. Age-related decrease in hypothalamic gonadotropin-releasing hormone (GnRH) gene expression, but not pituitary responsiveness to GnRH, in the male Brown Norway rat. J Androl. 2000 21:72-84.
  46. Bonavera JJ, Swerdloff RS, Sinha Hakim AP, Lue YH, Wang C. Aging results in attenuated gonadotropin releasing hormone-luteinizing hormone axis responsiveness to glutamate receptor agonist N-methyl-D-aspartate. J Neuroendocrinol. 1998 10:93-99.
  47. Deslypere JP, Kaufman JM, Vermeulen T, Vogelaers D, Vandalem JL, Vermeulen A. Influence of age on pulsatile luteinizing hormone release and responsiveness of the gonadotrophs to sex hormone feedback in men. J Clin Endocrinol Metab. 1987 64:68-73.
  48. Mahmoud AM, Goemaere S, De Bacquer D, Comhaire FH, Kaufman JM. Serum inhibin B levels in community-dwelling elderly men. Clin Endocrinol (Oxf). 2000 53:141-147.
  49. Pincus SM, Mulligan T, Iranmanesh A, Gheorghiu S, Godschalk M, Veldhuis JD. Older males secrete luteinizing hormone and testosterone more irregularly, and jointly more asynchronously, than younger males. Proc Natl Acad Sci U S A. 1996 Nov 26;93(24):14100-5.
  50. Keenan DM, Veldhuis JD. Disruption of the hypothalamic luteinizing hormone pulsing mechanism in aging men. Am J Physiol Regul Integr Comp Physiol. 2001 Dec;281(6):R1917-24.
  51. Morales A, Lunenfeld B; International Society for the Study of the Aging Male. Investigation, treatment and monitoring of late-onset hypogonadism in males. Official recommendations of ISSAM. International Society for the Study of the Aging Male. Aging Male. 2002 Jun;5(2):74-86
  52. Andropause, Consensus Panel. (2002). Proceedings of the Andropause Consensus Panel, The Endocrine Society.
  53. Bhasin S, Tenover JS. Age-associated sarcopenia--issues in the use of testosterone as an anabolic agent in older men. J Clin Endocrinol Metab 1997(b);82:1659-1660.
  54. Perry HM 3rd, Miller DK, Patrick P, Morley JE. Testosterone and leptin in older African-American men: relationship to age, strength, function, and season. Metabolism 2000;49:1085-91.
  55. Melton LJ, 3rd, Khosla S, Riggs BL. Epidemiology of sarcopenia. Mayo Clin Proc 2000; 75 Suppl:S10-2; discussion S12-3.
  56. Baumgartner RN, Koehler KM, Gallagher D, et al. Epidemiology of sarcopenia among the elderly in New Mexico [published erratum appears in Am J Epidemiol 1999 Jun 15;149(12):1161]. Am J Epidemiol 1998; 147:755-63.
  57. Baumgartner RN, Waters DL, Gallagher D, Morley JE, Garry PJ. Predictors of skeletal muscle mass in elderly men and women. Mech Ageing Dev. 1999;107:123-36.
  58. Holloszy JO. Workshop on sarcopenia: muscle atrophy in old age. J Gerontol 1995; 50A Spec No:1-161.
  59. Dutta C, Hadley EC. The significance of sarcopenia in old age. J Gerontol A Biol Sci Med Sci 1995; 50 Spec No:1-4.
  60. Kohrt WM, Malley MT, Dalsky GP, Holloszy JO. Body composition of healthy sedentary and trained, young and older men and women. Med Sci Sports Exerc 1992; 24:832-7.
  61. American Medical Association white paper on elderly health. Report of the Council on Scientific Affairs [published erratum appears in Arch Intern Med 1991;151(2):265]. Arch Intern Med 1990; 150:2459-72.
  62. Borkan GA, Hults DE, Gerzof SG, Robbins AH, Silbert CK. Age changes in body composition revealed by computed tomography. J Gerontol 1983;38:673-7.
  63. Borkan GA, Norris AH. Fat redistribution and the changing body dimensions of the adult male. Hum Biol 1977; 49:495-513.
  64. Cohn, SH, Vartsky D, Yasumura S, Sawitsky A, Zanzi I, Vaswani A, Ellis KJ. Compartmental body composition based on total-body nitrogen, potassium, and calcium. Am J Physiol 1980;239:E524-E530.
  65. Novak, LP. 1972 Aging, total body potassium, fat free-mass, and cell mass in males and females between ages 18-85 years. J Gerontol 1972;27:438-443.
  66. Schoeller, DA. Changes in total body water with age. Am J Clin Nutr 1989;50:1176-1181
  67. Evans, WJ. 1995 Effects of exercise on body composition and functional capacity of the elderly. J Gerontol 50A:147-150
  68. Flegg, JL and Lakatta, ED. Role of muscle loss in the age-associated reductions in VO2max. J Appl Physiol 1988;65:1147-1151
  69. Proctor, DN, Balagopal, P and Nair, KS. Age-related sarcopenia in humans is associated with reduced synthetic rates of specific muscle proteins. J Nutr 1998;128:351S-355S.
  70. Lexell, J, Henriksson-Larsen, K, Wimblod, B and Sjostrom, M. Distribution of different fiber types in human skeletal muscles: effects of aging studied in whole muscle cross section. Muscle Nerve 1983;6:588-595
  71. Lexell, J. Human aging, muscle mass, and fiber type composition. J Gerontol. 1995;50A:11-16
  72. Frontera, WR, Hughes, VR, Lutz, KJ and Evans, WJ. A cross-sectional study of muscle strength and mass in 45- to 75- yr-old men and women. J Appl Physiol 1991;71:644-650.
  73. Harris, T. Muscle mass and strength:  relation to function in population studies. J Nutr. 1997;27:1004S-1006S
  74. Skelton, DA, Greig, CA, Davies, JM and Young, A. Strength, power, and related functional ability of healthy people aged 65-89 years. Age Aging 1995;23:371-377.
  75. Roy TA, Blackman MR, Harman SM, Tobin JD, Schrager M, Metter EJ. Interrelationships of serum testosterone and free testosterone index with FFM and strength in aging men. Am J Physiol Endocrinol Metab. 2002 Aug;283(2):E284-94.
  76. O'Donnell AB, Travison TG, Harris SS, Tenover JL, McKinlay JB. Testosterone, dehydroepiandrosterone, and physical performance in older men: results from the Massachusetts Male Aging Study. J Clin Endocrinol Metab. 2006 Feb;91(2):425-31.
  77. Mohr BA, Bhasin S, Kupelian V, Araujo AB, O'Donnell AB, McKinlay JB. Testosterone, sex hormone-binding globulin, and frailty in older men. J Am Geriatr Soc. 2007 Apr;55(4):548-55.
  78. van den Beld AW, de Jong FH, Grobbee DE, Pols HA, Lamberts SW. Measures of bioavailable serum testosterone and estradiol and their relationships with muscle strength, bone density, and body composition in elderly men. J Clin Endocrinol Metab. 2000 Sep;85(9):3276-82.
  79. Schaap LA, Pluijm SM, Smit JH, van Schoor NM, Visser M, Gooren LJ, Lips P. The association of sex hormone levels with poor mobility, low muscle strength and incidence of falls among older men and women. Clin Endocrinol (Oxf). 2005 Aug;63(2):152-60.
  80. Orwoll E, Lambert LC, Marshall LM, Blank J, Barrett-Connor E, Cauley J, Ensrud K, Cummings SR; Osteoporotic Fractures in Men Study Group. Endogenous testosterone levels, physical performance, and fall risk in older men. Arch Intern Med. 2006 Oct 23;166(19):2124-31.
  81. Krasnoff JB, Basaria S, Pencina MJ, Jasuja GK, Vasan RS, Ulloor J, Zhang A, Coviello A, Kelly-Hayes M, D'Agostino RB, Wolf PA, Bhasin S, Murabito JM. Free testosterone levels are associated with mobility limitation and physical performance in community-dwelling men: the Framingham Offspring Study. J Clin Endocrinol Metab. 2010 Jun;95(6):2790-9.
  82. Hyde Z, Flicker L, Almeida OP, Hankey GJ, McCaul KA, Chubb SA, Yeap BB. Low free testosterone predicts frailty in older men: the health in men study. J Clin Endocrinol Metab. 2010 Jul;95(7):3165-72.
  83. Travison TG, Nguyen AH, Naganathan V, Stanaway FF, Blyth FM, Cumming RG, Le Couteur DG, Sambrook PN, Handelsman DJ. Changes in reproductive hormone concentrations predict the prevalence and progression of the frailty syndrome in older men: the concord health and ageing in men project. J Clin Endocrinol Metab. 2011 Aug;96(8):2464-74.
  84. Cawthon PM, Ensrud KE, Laughlin GA, Cauley JA, Dam TT, Barrett-Connor E, Fink HA, Hoffman AR, Lau E, Lane NE, Stefanick ML, Cummings SR, Orwoll ES; Osteoporotic Fractures in Men (MrOS) Research Group. Sex hormones and frailty in older men: the osteoporotic fractures in men (MrOS) study. J Clin Endocrinol Metab. 2009 Oct;94(10):3806-15.
  85. Travison TG, Morley JE, Araujo AB, O'Donnell AB, McKinlay JB. The relationship between libido and testosterone levels in aging men. J Clin Endocrinol Metab. 2006 Jul;91(7):2509-13.
  86. Zitzmann M, Faber S, Nieschlag E. Association of specific symptoms and metabolic risks with serum testosterone in older men. J Clin Endocrinol Metab. 2006 Nov;91(11):4335-43.
  87. Marberger M, Roehrborn CG, Marks LS, Wilson T, Rittmaster RS. Relationship among serum testosterone, sexual function, and response to treatment in men receiving dutasteride for benign prostatic hyperplasia. J Clin Endocrinol Metab. 2006 Apr;91(4):1323-8.
  88. Araujo AB, Esche GR, Kupelian V, O'donnell AB, Travison TG, Williams RE, Clark RV, McKinlay JB. Prevalence of Symptomatic Androgen Deficiency in Men. J Clin Endocrinol Metab. 2007 Aug 14; [Epub ahead of print]
  89. Kaiser FE, Viosca SP, Morley JE, Mooradian AD, Davis SS, Korenman SG. Impotence and aging: clinical and hormonal factors. J Am Geriatr Soc 1988;36:511-9.
  90. Korenman SG, Morley JE, Mooradian AD, et al. Secondary hypogonadism in older men: its relation to impotence. J Clin Endocrinol Metab 1990; 71:963-9.
  91. O'Connor DB, Lee DM, Corona G, Forti G, Tajar A, O'Neill TW, Pendleton N, Bartfai G, Boonen S, Casanueva FF, Finn JD, Giwercman A, Han TS, Huhtaniemi IT, Kula K, Labrie F, Lean ME, Punab M, Silman AJ, Vanderschueren D, Wu FC; European Male Ageing Study Group. The relationships between sex hormones and sexual function in middle-aged and older European men. J Clin Endocrinol Metab. 2011 Oct;96(10):E1577-87.
  92. Dodge HH, Du Y, Saxton JA, Ganguli M. Cognitive domains and trajectories of functional independence in nondemented elderly persons. J Gerontol A Biol Sci Med Sci. 2006 Dec;61(12):1330-7.
  93. Liu-Ambrose T, Pang MY, Eng JJ. Executive function is independently associated with performances of balance and mobility in community-dwelling older adults after mild stroke: implications for falls prevention. Cerebrovasc Dis. 2007;23(2-3):203-10.
  94. Janowsky JS, Oviatt SK, Orwoll ES. Testosterone influences spatial cognition in older men. Behav Neurosci. 1994 Apr;108(2):325-32.
  95. Janowsky JS. The role of androgens in cognition and brain aging in men. Neuroscience. 2006;138(3):1015-20.
  96. Fales CL, Knowlton BJ, Holyoak KJ, Geschwind DH, Swerdloff RS, Gonzalo IG. Working memory and relational reasoning in Klinefelter syndrome. J Int Neuropsychol Soc. 2003 Sep;9(6):839-46.
  97. Alexander GM, Swerdloff RS, Wang C, Davidson T, McDonald V, Steiner B, Hines M. Androgen-behavior correlations in hypogonadal men and eugonadal men. II. Cognitive abilities. Horm Behav. 1998 Apr;33(2):85-94.
  98. Barrett-Connor E, Goodman-Gruen D, Patay B. Endogenous sex hormones and cognitive function in older men. J Clin Endocrinol Metab 1999(a);84:3681-5.
  99. Gouchie C, Kimura D. The relationship between testosterone levels and cognitive ability patterns. Psychoneuroendocrinology 1991; 16:323-34.
  100. Shute VJ PJ, Hubert L, Reynolds RW. The relationship between androgen levels and human spatial abilities. Bull Psychonomic Soc 1983; 21:465-468.
  101. Margolese HC. The male menopause and mood: testosterone decline and depression in the aging male--is there a link? J Geriatr Psychiatry Neurol 2000;13(2):93-101
  102. Barrett-Conner E, Von Muhlen DG, Kritz-Silverstein D. Bioavailable testosterone and depressed mood in older men: the Rancho Bernardo study. J Clin Endocrinol Metab 1999(b);84:573-577.
  103. Van Honk J, Tuiten A, Verbaten R, van den Hout M, Koppeschaar H, Thijssen J, de Haan E. Correlations among salivary testosterone, mood, and selective attention to threat in humans. Hormones Behav 1999; 36:17-24.
  104. Seidman SN, Araujo AB, Roose SP, McKinlay JB. Testosterone level, androgen receptor polymorphism, and depressive symptoms in middle-aged men. Biol Psychiatry. 2001 Sep 1;50(5):371-6.
  105. Seidman SN, Araujo AB, Roose SP, Devanand DP, Xie S, Cooper TB, McKinlay JB. Low testosterone levels in elderly men with dysthymic disorder. Am J Psychiatry. 2002 Mar;159(3):456-9.
  106. Delhez M, Hansenne M, Legros JJ. Andropause and psychopathology: minor symptoms rather than pathological ones. Psychoneuroendocrinology. 2003 Oct;28(7):863-74.
  107. Fink HA, Ewing SK, Ensrud KE, Barrett-Connor E, Taylor BC, Cauley JA, Orwoll ES. Association of testosterone and estradiol deficiency with osteoporosis and rapid bone loss in older men.
    J Clin Endocrinol Metab. 2006 Oct;91(10):3908-15.
  108. Greendale GA, Edelstein S, Barrett-Connor E. Endogenous sex steroids and bone mineral density in older women and men: the Rancho Bernardo Study. J Bone Miner Res 1997;12:1833-43.
  109. Amin S, Zhang Y, Sawin CT, Evans SR, Hannan MT, Kiel DP, Wilson PW, Felson DT. Association of hypogonadism and estradiol levels with bone mineral density in elderly men from the Framingham study. Ann Intern Med. 2000 Dec 19;133(12):951-63.
  110. Khosla S, Melton LJ III, Atkinson EJ, et al. Relationship of serum sex steroid levels and bone turnover markers with bone mineral density in men and women: a key role for bioavailable estrogen. J Clin Endocrinol Metab 1998;83:2266-2274.
  111. Center JR, Nguyen TV, Sambrook PN, Eisman JA. Hormonal and biochemical parameters in the determination of osteoporosis in elderly men. J Clin Endocrinol Metab 1999;84:3626-3635.
  112. Boonen S, Vanderscheren D, Geusens P, Bouillon R. Age-associated endocrine deficiencies as potential determinants of femoral neck (type II) osteoporotic fracture occurrence in elderly men. Int J Androl 1997;20:134-143.
  113. Khosla S, Melton LJ 3rd, Robb RA, Camp JJ, Atkinson EJ, Oberg AL, Rouleau PA, Riggs BL. Relationship of volumetric BMD and structural parameters at different skeletal sites to sex steroid levels in men. J Bone Miner Res. 2005 May;20(5):730-40.
  114. Mellstrom D, Johnell O, Ljunggren O, Eriksson AL, Lorentzon M, Mallmin H, Holmberg A, Redlund-Johnell I, Orwoll E, Ohlsson C. Free testosterone is an independent predictor of BMD and prevalent fractures in elderly men: MrOS Sweden. J Bone Miner Res. 2006 Apr;21(4):529-35.
  115. Shores MM, Moceri VM, Gruenewald DA, Brodkin KI, Matsumoto AM, Kivlahan DR. Low testosterone is associated with decreased function and increased mortality risk: a preliminary study of men in a geriatric rehabilitation unit. J Am Geriatr Soc. 2004 Dec;52(12):2077-81.
  116. Shores MM, Matsumoto AM, Sloan KL, Kivlahan DR. Low serum testosterone and mortality in male veterans. Arch Intern Med. 2006 Aug 14-28;166(15):1660-5.
  117. Araujo AB, Kupelian V, Page ST, Handelsman DJ, Bremner WJ, McKinlay JB. Sex steroids and all-cause and cause-specific mortality in men. Arch Intern Med. 2007 Jun 25;167(12):1252-60.
  118. Laughlin GA, Barrett-Connor E, Bergstrom J. Low Serum Testosterone and Mortality in Older Men. J Clin Endocrinol Metab. 2007 Oct 2; [Epub ahead of print]
  119. Araujo AB, Dixon JM, Suarez EA, Murad MH, Guey LT, Wittert GA. Clinical review: Endogenous testosterone and mortality in men: a systematic review and meta-analysis. J Clin Endocrinol Metab. 2011 Oct;96(10):3007-19.
  120. Corona G, Rastrelli G, Monami M, Guay A, Buvat J, Sforza A, Forti G, Mannucci E, Maggi M. Hypogonadism as a risk factor for cardiovascular mortality in men: a meta-analytic study. Eur J Endocrinol. 2011 Nov;165(5):687-701.
  121. Litman HJ, Bhasin S, O'Leary MP, Link CL, McKinlay JB; BACH Survey Investigators. An investigation of the relationship between sex-steroid levels and urological symptoms: results from the Boston Area Community Health survey. BJU Int. 2007 Aug;100(2):321-6. Epub 2007 May 17.
  122. Alexandersen P, Haarbo J, Christiansen C. The relationship of natural androgens to coronary heart disease in males: a review. Atherosclerosis 1996; 125:1-13.
  123. Barrett-Connor E, Khaw KT. Endogenous sex hormones and cardiovascular disease in men. A prospective population-based study. Circulation 1988;78:539-45.
  124. Cauley JA, Gutai JP, Kuller LH, Dai WS. Usefulness of sex steroid hormone levels in predicting coronary artery disease in men. Am J Cardiol. 1987 Oct 1;60(10):771-7.
  125. Contoreggi CS, Blackman MR, Andres R, Muller DC, Lakatta EG, Fleg JL, Harman SM. Plasma levels of estradiol, testosterone, and DHEAS do not predict risk of coronary artery disease in men. J Androl. 1990 Sep-Oct;11(5):460-70.
  126. Yarnell JW, Beswick AD, Sweetnam PM, Riad-Fahmy D. Endogenous sex hormones and ischemic heart disease in men. The Caerphilly prospective study. Arterioscler Thromb. 1993 Apr;13(4):517-20.
  127. Hak AE, Witteman JC, de Jong FH, Geerlings MI, Hofman A, Pols HA. Low levels of endogenous androgens increase the risk of atherosclerosis in elderly men: the Rotterdam study. J Clin Endocrinol Metab. 2002 Aug;87(8):3632-9.
  128. Bassey, E.J, Fiatarone, MA, O’Neill, EF, Kelly, M, Evans, WJ. 1992  Leg extensor power and functional performance in very old men and women. Clin Sci 82:321-327
  129. Rantanen, T and Avela, J. Leg extension power and walking speed in very old people living independently. J Gerontol 1997;52A:M225-M23
  130. Whipple, RH, Wolfson, LI and Amerman, M. 1987  The relationship of knee and ankle weakness to falls in nursing home residents:  an isokinetic study. J Am Geriatr Soc 1987;35: 13-20.
  131. Wolfson, L, Judge, J, Whipple, R and King, M. Strength is a major factor in balance, gait, and the occurrence of falls. J Gerontol 1995;50A:64-67
  132. Bhasin S, Calof OM, Storer TW, Lee ML, Mazer NA, Jasuja R, Montori VM, Gao W, Dalton JT. Drug insight: Testosterone and selective androgen receptor modulators as anabolic therapies for chronic illness and aging. Nat Clin Pract Endocrinol Metab. 2006 Mar;2(3):146-59.
  133. Katznelson L, Finkelstein JS, Schonenfeld DA, Rosenthal DI, Anderson EJ, Klibanski A. Increase in bone density and lean body mass during testosterone administration in men with acquired hypogonadism. J Clin Endocrinol Metab 1996;81:4358-4365.
  134. Bhasin S, Storer TW, Berman N, Yarasheski KE, Clevenger B, Phillips J et al. Testosterone replacement increases fat-free mass and muscle size in hypogonadal men. J Clin Endocrinol Metab 1997(a);82:407-413.
  135. Brodsky IG, Balagopal P, Nair KS. Effects of testosterone replacement on muscle mass and muscle protein synthesis in hypogonadal men--a clinical research center study. J Clin Endocrinol Metab 1992;75:1092-1098.
  136. Snyder PJ, Peachey H, Berlin JA, Hannoush P, Haddad G, Dlewati A, Santanna J, Loh L, Lenrow DA, Holmes JH, Kapoor SC, Atkinson LE, Strom BL. Effects of testosterone replacement in hypogonadal men. J Clin Endocrinol Metab 2000;85(8):2670-7.
  137. Wang C, Swedloff RS, Iranmanesh A, Dobs A, Snyder PJ, Cunningham G, Matsumoto AM, Weber T, Berman. Transdermal testosterone gel improves sexual function, mood, muscle strength, and body composition parameters in hypogonadal men. Testosterone Gel Study Group. J Clin Endocrinol Metab 2000;85(8):2839-53.
  138. Steidle C, Schwartz S, Jacoby K, Sebree T, Smith T, Bachand R; North American AA2500 T Gel Study Group. AA2500 testosterone gel normalizes androgen levels in aging males with improvements in body composition and sexual function. J Clin Endocrinol Metab. 2003 Jun;88(6):2673-81.
  139. Bhasin S, Woodhouse L, Casaburi R, Singh AB, Bhasin D, Berman N, Chen X, Yarasheski KE, Magliano L, Dzekov C, Dzekov J, Bross R, Phillips J, Sinha-Hikim I, Shen R, Storer TW. Testosterone dose-response relationships in healthy young men. Am J Physiol Endocrinol Metab. 2001 281:E1172-81.
  140. Storer TW, Magliano L, Woodhouse L, Lee ML, Dzekov C, Dzekov J, Casaburi R, Bhasin S. Testosterone dose-dependently increases maximal voluntary strength and leg power, but does not affect fatigability or specific tension. J Clin Endocrinol Metab. 2003 Apr;88(4):1478-85.
  141. Woodhouse LJ, Reisz-Porszasz S, Javanbakht M, Storer TW, Lee M, Zerounian H, Bhasin S. Development of models to predict anabolic response to testosterone administration in healthy young men. Am J Physiol Endocrinol Metab. 2003 May;284(5):E1009-17.
  142. Bhasin S, Storer TW, Berman N, et al. The effects of supraphysiologic doses of testosterone on muscle size and strength in normal men [see comments]. N Engl J Med 1996; 335:1-7.
  143. Urban RJ, Bodenburg YH, Gilkison C, Foxworth J, Coggan AR, Wolfe RR et al. Testosterone administration to elderly men increases skeletal muscle strength and protein synthesis. Am J Physiol 1995;269(5 Pt 1):E820-E826.
  144. Ferrando AA, Sheffield-Moore M, Yeckel CW, Gilkison C, Jiang J, Achacosa A, Lieberman SA, Tipton K, Wolfe RR, Urban RJ. 2002 Testosterone administration to older men improves muscle function: molecular and physiological mechanisms. Am J Physiol Endocrinol Metab. 282:E601-E607.
  145. Kong A, Edmonds P. Testosterone therapy in HIV wasting syndrome: systematic review and meta-analysis. Lancet Infect Dis. 2002 Nov;2(11):692-9.
  146. Johns K, Beddall MJ, Corrin RC. Anabolic steroids for the treatment of weight loss in HIV-infected individuals. Cochrane Database Syst Rev. 2005 Oct 19;(4):CD005483.
  147. Bhasin S, Storer TW, Javanbakht M, et al. Testosterone replacement and resistance exercise in HIV-infected men with weight loss and low testosterone levels [see comments]. JAMA 2000; 283:763-70.
  148. Grinspoon S, Corcoran C, Askari H, Schoenfeld D, Wolf L, Burrows B, Walsh M, Hayden D, Parlman K, Anderson E, Basgoz N, Klibanski A. Effects of androgen administration in men with the AIDS wasting syndrome. A randomized, double-blind, placebo-controlled trial. Ann Intern Med. 1998 Jul 1;129(1):18-26.
  149. Storer TW, Woodhouse LJ, Sattler F, Singh AB, Schroeder ET, Beck K, Padero M, Mac P, Yarasheski KE, Geurts P, Willemsen A, Harms MK, Bhasin S. A randomized, placebo-controlled trial of nandrolone decanoate in human immunodeficiency virus-infected men with mild to moderate weight loss with recombinant human growth hormone as active reference treatment. J Clin Endocrinol Metab. 2005 Aug;90(8):4474-82.
  150. Grinspoon S, Corcoran C, Parlman K, Costello M, Rosenthal D, Anderson E, Stanley T, Schoenfeld D, Burrows B, Hayden D, Basgoz N, Klibanski A. Effects of testosterone and progressive resistance training in eugonadal men with AIDS wasting. A randomized, controlled trial. Ann Intern Med. 2000 Sep 5;133(5):348-55.
  151. Bhasin S, Storer TW, Asbel-Sethi N, Kilbourne A, Hays R, Sinha-Hikim I, Shen R, Arver S, Beall G. Effects of testosterone replacement with a nongenital, transdermal system, Androderm, in human immunodeficiency virus-infected men with low testosterone levels. J Clin Endocrinol Metab. 1998 Sep;83(9):3155-62.
  152. Grinspoon S, Corcoran C, Stanley T, Baaj A, Basgoz N, Klibanski A. Effects of hypogonadism and testosterone administration on depression indices in HIV-infected men. J Clin Endocrinol Metab. 2000 Jan;85(1):60-5.
  153. Knapp PE, Storer TW, Herbst KL, Singh AB, Dzekov C, Dzekov J, LaValley M, Zhang A, Ulloor J, Bhasin S. Effects of a supraphysiological dose of testosterone on physical function, muscle performance, mood, and fatigue in men with HIV-associated weight loss. Am J Physiol Endocrinol Metab. 2008 Jun;294(6):E1135-43.
  154. Isidori AM, Giannetta E, Greco EA, Gianfrilli D, Bonifacio V, Isidori A, Lenzi A, Fabbri A. Effects of testosterone on body composition, bone metabolism and serum lipid profile in middle-aged men: a meta-analysis. Clin Endocrinol (Oxf). 2005 Sep;63(3):280-93.
  155. Sih R, Morley JE, Kaiser FE, Perry HM, III, Patrick P, Ross C. Testosterone replacement in older hypogonadal men: a 12-month randomized controlled trial [see comments]. J Clin Endocrinol Metab 1997;82(6):1661-1667.
  156. Snyder PJ, Peachey H, Hannoush P, Berlin JA, Loh L, Lenrow DA. Effect of testosterone treatment on body composition and muscle strength in men over 65 years of age. J Clin Endocrinol Metab 1999; 84(8):2647-2653.
  157. Page ST, Amory JK, Bowman FD, Anawalt BD, Matsumoto AM, Bremner WJ, Tenover JL. Exogenous testosterone (T) alone or with finasteride increases physical performance, grip strength, and lean body mass in older men with low serum T. J Clin Endocrinol Metab. 2005 Mar;90(3):1502-10.
  158. Schroeder ET, Singh AB, Bhasin S, Storer TW, Azen C, Davidson T, Martinez C, Sinha-Hikim I, Jaque V, Terk M, Sattler FR. The effects of an oral androgen on muscle and metabolism in older, community, dwelling men. Am J Physiol Endocrinol Metab. 2003;284(1):E120-8.
  159. Kenny AM, Prestwood KM, Gruman CA, Marcello KM, Raisz LG. Effects of transdermal testosterone on bone and muscle in older men with low bioavailable testosterone levels. J Gerontol A Biol Sci Med Sci. 2001 May;56(5):M266-72.
  160. Emmelot-Vonk MH, Verhaar HJ, Nakhai Pour HR, Aleman A, Lock TM, Bosch JL, Grobbee DE, van der Schouw YT. Effect of testosterone supplementation on functional mobility, cognition, and other parameters in older men: a randomized controlled trial. JAMA. 2008 Jan 2;299(1):39-52.
  161. Nair KS, Rizza RA, O'Brien P, Dhatariya K, Short KR, Nehra A, Vittone JL, Klee GG, Basu A, Basu R, Cobelli C, Toffolo G, Dalla Man C, Tindall DJ, Melton LJ 3rd, Smith GE, Khosla S, Jensen MD. DHEA in elderly women and DHEA or testosterone in elderly men. N Engl J Med. 2006 Oct 19;355(16):1647-59.
  162. Travison, T.G. et al. Clinical meaningfulness of the changes in muscle performance and physical function associated with testosterone administration in older men with mobility limitation. J Gerontol A Biol Sci Med Sci 66, 1090-9 (2011).
  163. Basaria, S. et al. Adverse events associated with testosterone administration. N Engl J Med 363, 109-22 (2010)
  164. Kenny, A.M. et al. Effects of transdermal testosterone on bone and muscle in older men with low bioavailable testosterone levels, low bone mass, and physical frailty. J Am Geriatr Soc 58, 1134-43 (2010).
  165. Srinivas-Shankar, U. et al. Effects of testosterone on muscle strength, physical function, body composition, and quality of life in intermediate-frail and frail elderly men: a randomized, double-blind, placebo-controlled study. J Clin Endocrinol Metab 95, 639-50 (2010).
  166. Spitzer M, Huang G, Basaria S, Travison TG, Bhasin S. Risks and benefits of testosterone therapy in older men. Nat Rev Endocrinol. 2013 Jul;9(7):414-24.
  167. Sinha-Hikim I, Artaza J, Woodhouse L, Gonzalez-Cadavid N, Singh AB, Lee MI, Storer TW, Casaburi R, Shen R, Bhasin S. Testosterone-induced increase in muscle size in healthy young men is associated with muscle fiber hypertrophy. Am J Physiol Endocrinol Metab. 2002 2831:E154-E164.
  168. Sinha-Hikim I, Roth SM, Lee MI, Bhasin S. Testosterone-induced muscle hypertrophy is associated with an increase in satellite cell number in healthy, young men. Am J Physiol Endocrinol Metab. 2003 Jul;285(1):E197-205.
  169. Bhasin S, Taylor WE, Singh R, Artaza J, Choi H, Sinha-Hikim I, Jasuja R, Gonzalez-Cadavid N. Mechanisms of anabolic effects of testosterone on the muscle: mesenchymal, pluripotent stem cell as the target of androgen action. J Gerontol Med Sci 2003 (in press).
  170. Singh R, Artaza J, Taylor WE, Gonzalez-Cadavid N, Bhasin S. Androgens stimulate myogenesis and inhibit adipogenic differentiation in pluripotent C3H10T1/2 cells. Endocrinology 2003 (in press).
  171. Singh R, Artaza JN, Taylor WE, Braga M, Yuan X, Gonzalez-Cadavid NF, Bhasin S. Testosterone inhibits adipogenic differentiation in 3T3-L1 cells: nuclear translocation of androgen receptor complex with beta-catenin and T-cell factor 4 may bypass canonical Wnt signaling to down-regulate adipogenic transcription factors. Endocrinology. 2006 Jan;147(1):141-54.
  172. Gupta V, Bhasin S, Guo W, Singh R, Miki R, Chauhan P, Choong K, Tchkonia T, Lebrasseur NK, Flanagan JN, Hamilton JA, Viereck JC, Narula NS, Kirkland JL, Jasuja R. Effects of dihydrotestosterone on differentiation and proliferation of human mesenchymal stem cells and preadipocytes. Mol Cell Endocrinol. 2008 Dec 16;296(1-2):32-40.
  173. Singh R, Bhasin S, Braga M, Artaza JN, Pervin S, Taylor WE, Krishnan V, Sinha SK, Rajavashisth TB, Jasuja R. Regulation of myogenic differentiation by androgens: cross talk between androgen receptor/ beta-catenin and follistatin/transforming growth factor-beta signaling pathways. Endocrinology. 2009 Mar;150(3):1259-68.
  174. Jasuja R, Costello JC, Singh R, Gupta V, Spina CS, Toraldo G, Jang H, Li H, Serra C, Guo W, Chauhan P, Narula NS, Guarneri T, Ergun A, Travison TG, Collins JJ, Bhasin S. Combined administration of testosterone plus an ornithine decarboxylase inhibitor as a selective prostate-sparing anabolic therapy. Aging Cell. 2013 Dec 4. doi: 10.1111/acel.12174.
  175. Bartsch W, Krieg M, Voigt KD. Quantification of endogenous testosterone, 5 alpha-dihydrotestosterone and 5 alpha-androstane-3 alpha, 17 beta-diol in subcellular fractions of the prostate, bulbocavernosus/levator ani muscle, skeletal muscle and heart muscle of the rat. J Steroid Biochem. 1980 13:259-264.
  176. Wilson JD.  The role of 5alpha-reduction in steroid hormone physiology.
    Reprod Fertil Dev. 2001;13(7-8):673-8.
  177. Andriole GL, Kirby R. Safety and tolerability of the dual 5alpha-reductase inhibitor dutasteride in the treatment of benign prostatic hyperplasia. Eur Urol 2003;44:82-8.
  178. Bhasin S, Travison TG, Storer TW, Lakshman K, Kaushik M, Mazer NA, Ngyuen AH, Davda MN, Jara H, Aakil A, Anderson S, Knapp PE, Hanka S, Mohammed N, Daou P, Miciek R, Ulloor J, Zhang A, Brooks B, Orwoll K, Hede-Brierley L, Eder R, Elmi A, Bhasin G, Collins L, Singh R, Basaria S. Effect of testosterone supplementation with and without a dual 5α-reductase inhibitor on fat-free mass in men with suppressed testosterone production: a randomized controlled trial. JAMA. 2012 Mar 7;307(9):931-9.
  179. Borst SE, Conover CF, Carter CS, et al. Anabolic effects of testosterone are preserved during inhibition of 5alpha-reductase. Am J Physiol Endocrinol Metab. 2007;293:E507-14.
  180. Matsumoto AM, Tenover L, McClung M, et al. The long-term effect of specific type II 5alpha-reductase inhibition with finasteride on bone mineral density in men: results of a 4-year placebo controlled trial. J Urol. 2002;167:2105-8.
  181. Amory JK, Anawalt BD, Matsumoto AM, et al. The effect of 5alpha-reductase inhibition with dutasteride and finasteride on bone mineral density, serum lipoproteins, hemoglobin, prostate specific antigen and sexual function in healthy young men. J Urol. 2008;179:2333-8
  182. Jones ME, Thorburn AW, Britt KL, Hewitt KN, Wreford NG, Proietto J, Oz OK, Leury BJ, Robertson KM, Yao S, Simpson ER. Aromatase-deficient (ArKO) mice have a phenotype of increased adiposity. Proc Natl Acad Sci U S A 2000;97(23):12735-40.
  183. Carani C, Qin K, Simoni M, Faustini-Fustini M, Serpente S, Boyd J, Korach KS, Simpson ER. Effect of testosterone and estradiol in a man with aromatase deficiency. N Engl J Med. 1997 Jul 10;337(2):91-5.
  184. Finkelstein JS, Lee H, Burnett-Bowie SA, Pallais JC, Yu EW, Borges LF, Jones BF, Barry CV, Wulczyn KE, Thomas BJ, Leder BZ. Gonadal steroids and body composition, strength, and sexual function in men. N Engl J Med. 2013 Sep 12;369(11):1011-22.
  185. Bhasin S, Enzlin P, Coviello A, Basson R. Sexual dysfunction in men and women with endocrine disorders. Lancet. 2007 Feb 17;369(9561):597-611.
  186. Arver S, Dobs AS, Meikle AW, Allen RP, Sanders SW, Mazer NA. Improvement of sexual function in testosterone deficient men treated for 1 year with a permeation enhanced testosterone transdermal system. J Urol 1996; 155:1604-8.
  187. Bancroft J, Wu FC. Changes in erectile responsiveness during androgen replacement therapy. Arch Sex Behav 1983;12:59-66.
  188. Davidson JM, Camargo CA, Smith ER. Effects of androgen on sexual behavior in hypogonadal men. J Clin Endocrinol Metab 1979; 48:955-8.
  189. Kwan M, Greenleaf WJ, Mann J, Crapo L, Davidson JM. The nature of androgen action on male sexuality: a combined laboratory- self-report study on hypogonadal men. J Clin Endocrinol Metab 1983; 57:557-62.
  190. Bagatell CJ, Heiman JR, Rivier JE, Bremner WJ. Effects of endogenous testosterone and estradiol on sexual behavior in normal young men [published erratum appears in J Clin Endocrinol Metab 1994 Jun;78(6):1520]. J Clin Endocrinol Metab 1994; 78:711-6.
  191. Carani C, Granata AR, Bancroft J, Marrama P. The effects of testosterone replacement on nocturnal penile tumescence and rigidity and erectile response to visual erotic stimuli in hypogonadal men. Psychoneuroendocrinology 1995;20:743-53.
  192. Cunningham GR, Hirshkowitz M, Korenman SG, Karacan I. Testosterone replacement therapy and sleep-related erections in hypogonadal men. J Clin Endocrinol Metab. 1990 Mar;70(3):792-7.
  193. Carani C, Bancroft J, Granata A, Del Rio G, Marrama P. Testosterone and erectile function, nocturnal penile tumescence and rigidity, and erectile response to visual erotic stimuli in hypogonadal and eugonadal men. Psychoneuroendocrinology. 1992 Nov;17(6):647-54.
  194. Lugg J, Ng C, Rajfer J, Gonzalez-Cadavid N. Cavernosal nerve stimulation in the rat reverses castration-induced decrease in penile NOS activity. Am J Physiol 1996; 271:E354-61.
  195. Shabsigh R. The effects of testosterone on the cavernous tissue and erectile function. World J Urol. 1997;15(1):21-6. Review.
  196. Mills TM, Lewis RW, Stopper VS. Androgenic maintenance of inflow and veno-occlusion during erection in the rat. Biol Reprod. 1998 Dec;59(6):1413-8.
  197. Mills TM, Dai Y, Stopper VS, Lewis RW. Androgenic maintenance of the erectile response in the rat. Steroids. 1999 Sep;64(9):605-9.
  198. Bhasin S, Fielder T, Peacock N, Sod-Moriah UA, Swerdloff RS. Dissociating antifertility effects of GnRH-antagonist from its adverse effects on mating behavior in male rats. Am J Physiol 1988; 254:E84-91.
  199. Fielder TJ, Peacock NR, McGivern RF, Swerdloff RS, Bhasin S. Testosterone dose-dependency of sexual and nonsexual behaviors in the gonadotropin-releasing hormone antagonist-treated male rat. J Androl 1989; 10:167-73.
  200. Buena F, Swerdloff RS, Steiner BS, et al. Sexual function does not change when serum testosterone levels are pharmacologically varied within the normal male range. Fertil Steril 1993; 59:1118-23.
  201. Kaiser FE, Viosca SP, Morley JE, Mooradian AD, Davis SS, Korenman SG. Impotence and aging: clinical and hormonal factors. J Am Geriatr Soc 1988;36:511-9.
  202. Korenman SG, Morley JE, Mooradian AD, et al. Secondary hypogonadism in older men: its relation to impotence. J Clin Endocrinol Metab 1990; 71:963-9.
  203. Morley JE, Korenman SG, Mooradian AD, Kaiser FE. Sexual dysfunction in the elderly male [clinical conference]. J Am Geriatr Soc 1987; 35:1014-22.
  204. Buvat J, Lemaire A. Endocrine screening in 1,022 men with erectile dysfunction: clinical significance and cost-effective strategy [see comments]. J Urol 1997; 158:1764-7.
  205. Citron JT, Ettinger B, Rubinoff H, et al. Prevalence of hypothalamic-pituitary imaging abnormalities in impotent men with secondary hypogonadism. J Urol 1996;155:529-33.
  206. Hajjar RR, Kaiser FE, Morley JE. Outcomes of long-term testosterone replacement in older hypogonadal males: a retrospective analysis. J Clin Endocrinol Metab 1997;82:3793-6.
  207. Jain P, Rademaker AW, McVary KT. Testosterone supplementation for erectile dysfunction: results of a meta-analysis. J Urol 2000;164:371-5.
  208. Bolona ER, Uraga MV, Haddad RM, Tracz MJ, Sideras K, Kennedy CC, Caples SM, Erwin PJ, Montori VM. Testosterone use in men with sexual dysfunction: a systematic review and meta-analysis of randomized placebo-controlled trials. Mayo Clin Proc. 2007 Jan;82(1):20-8.
  209. Isidori AM, Giannetta E, Gianfrilli D, Greco EA, Bonifacio V, Aversa A, Isidori A, Fabbri A, Lenzi A. Effects of testosterone on sexual function in men: results of a meta-analysis. Clin Endocrinol (Oxf). 2005 Oct;63(4):381-94.
  210. Shabsigh R, Kaufman JM, Steidle C, Padma-Nathan H. Randomized study of testosterone gel as adjunctive therapy to sildenafil in hypogonadal men with erectile dysfunction who do not respond to sildenafil alone. J Urol. 2004 Aug;172(2):658-63.
  211. Aversa A, Isidori AM, Spera G, Lenzi A, Fabbri A. Androgens improve cavernous vasodilation and response to sildenafil in patients with erectile dysfunction. Clin Endocrinol (Oxf). 2003 May;58(5):632-8.
  212. Buvat J, Montorsi F, Maggi M, Porst H, Kaipia A, Colson MH, Cuzin B, Moncada I, Martin-Morales A, Yassin A, Meuleman E, Eardley I, Dean JD, Shabsigh R. Hypogonadal men nonresponders to the PDE5 inhibitor tadalafil benefit from normalization of testosterone levels with a 1% hydroalcoholic testosterone gel in the treatment of erectile dysfunction (TADTEST study). J Sex Med. 2011 Jan;8(1):284-93. doi: 10.1111/j.
  213. Spitzer M, Basaria S, Travison TG, Davda MN, Paley A, Cohen B, Mazer NA, Knapp PE, Hanka S, Lakshman KM, Ulloor J, Zhang A, Orwoll K, Eder R, Collins L, Mohammed N, Rosen RC, DeRogatis L, Bhasin S. Effect of testosterone replacement on response to sildenafil citrate in men with erectile dysfunction: a parallel, randomized trial. Ann Intern Med. 2012 Nov 20;157(10):681-91.
  214. Stepan JJ, Lachman M, Zverina J, Pacovsky V, Baylink DJ. Castrated men exhibit bone loss: effect of calcitonin treatment on biochemical indices of bone remodeling. J Clin Endocrinol Metab 1989;69:523-7.
  215. Goldray D, Weisman Y, Jaccard N, Merdler C, Chen J, Matzkin H. Decreased bone mineral density in elderly men treated with gonadortropin-releaasing hormone agonist decapeptyl (D-Trp6-GnRH). J Clin Endocrinol Metab 1993;76:288-290.
  216. Daniell HW. Osteoporosis after orchiectomy for prostate cancer. J Urol 1997;157:439-444.
  217. Eriksson S, Eriksson A, Stege R, Carlstrom K. Bone mineral density in patients with prostate cancer treated with orchiectomy and with estrogens. Calcif Tissue Int 1995;57:97-99.
    1. Vanderschueren D, Van Herck E, Schot P, Rush E, Einhorn T, Geusens P, Bouillon R. The aged male rat as a model for human osteoporosis: evaluation by nondestructive measurements and biomechanical testing. Calcif Tissue Int. 1993 Nov;53(5):342-7.
  218. Finkelstein JS, Klibanski A, Neer RM, Grrenspan SL, Rosenthal DI, Crowley Jr WF. Osteoporosis in men with idiopathic hypogonadotropic hypogonadism. Ann Intern Med 1987;106:354-361.
  219. Bonjour J-P, Theintz G, Buchs B, Slosman D, Rizzoli R. Critical years and stages of puberty for spinal and femoral bone mass and accumulation during adolescence. J Clin Endocrinol Metab 1991;73:555-563.
  220. Behre HM, Kliesch S, Leifke E, Link TM, Nieschlag E. Long-term effect of testosterone therapy on bone mineral density in hypogonadal men. J Clin Endocrinol Metab 1997;82:2386-90.
  221. Leifke E, Korner HC, Link TM, Behre HM, Peters PE, Nieschlag E. Effects of testosterone replacement therapy on cortical and trabecular bone mineral density, vertebral body area and paraspinal muscle area in hypogonadal men. Eur J Endocrinol 1998;138:51-58.
  222. Guo CY, Jones TH, Eastell R. Treatment of isolated hypogonadotropic hypogonadism effect on bone mineral density and bone turnover. J Clin Endocrinol Metab. 1997 Feb;82(2):658-65.
  223. Aminorroaya A, Kelleher S, Conway AJ, Ly LP, Handelsman DJ. Adequacy of androgen replacement influences bone density response to testosterone in androgen-deficient men. Eur J Endocrinol. 2005 Jun;152(6):881-6.
  224. Wang C, Swerdloff RS, Iranmanesh A, Dobs A, Snyder PJ, Cunningham G, Matsumoto AM, Weber T, Berman N. Effects of transdermal testosterone gel on bone turnover markers and bone mineral density in hypogonadal men. Clin Endocrinol (Oxf). 2001 Jun;54(6):739-50.
  225. Stanley HL, Schmitt BP, Poses RM, Deiss WP. Does hypogonadism contribute to the occurrence of a minimal trauma hip fracture in elderly men? J Am Geriatr Soc 1991;39:766-71.
  226. Kenny AM, Prestwood KM, Gruman CA, Marcello KM, Raisz LG. Effects of transdermal testosterone on bone and muscle in older men with low bioavailable testosterone levels. J Gerontol A Biol Sci Med Sci. 2001 May;56(5):M266-72.
  227. Amory JK, Watts NB, Easley KA, Sutton PR, Anawalt BD, Matsumoto AM, Bremner WJ, Tenover JL. Exogenous testosterone or testosterone with finasteride increases bone mineral density in older men with low serum testosterone. J Clin Endocrinol Metab. 2004 Feb;89(2):503-10.
  228. Snyder PJ, Peachey H, Hannoush P, Berlin JA, Loh L, Holmes JH. Effect of testosterone treatment on bone mineral density in men over 65 years of age. J Clin Endocrinol Metab 1999; 84:1966-1972.
  229. Tracz MJ, Sideras K, Boloña ER, Haddad RM, Kennedy CC, Uraga MV, Caples SM, Erwin PJ, Montori VM. Testosterone use in men and its effects on bone health. A systematic review and meta-analysis of randomized placebo-controlled trials. J Clin Endocrinol Metab. 2006 Jun;91(6):2011-6.
  230. Anderson FH, Francis RM, Peaston RT, Eastell HJ. Androgen supplementation in eugonadal men with osteoporosis: effects of 6 months treatment on markers of bone formation and resorption. J Bone Mineral Res 1997;12:472-478.
  231. Leder BZ, LeBlanc KM, Schoenfeld DA, Eastell R, Finkelstein JS. Differential effects of androgens and estrogens on bone turnover in normal men. J Clin Endocrinol Metab. 2003 Jan;88(1):204-10.
  232. Falahati-Nini A, Riggs BL, Atkinson EJ, O'Fallon WM, Eastell R, Khosla S. Relative contributions of testosterone and estrogen in regulating bone resorption and formation in normal elderly men. J Clin Invest. 2000 Dec;106(12):1553-60.
  233. Wang C, Swerdloff RS, Iranmanesh A, Dobs A, Snyder PJ, Cunningham G, Matsumoto AM, Weber T, Berman N. Effects of transdermal testosterone gel on bone turnover markers and bone mineral density in hypogonadal men. Clin Endocrinol (Oxf). 2001 Jun;54(6):739-50.
  234. Riggs BL, Khosla S, Melton LJ 3rd. Sex steroids and the construction and conservation of the adult skeleton. Endocr Rev. 2002 Jun;23(3):279-302.
  235. Sims NA, Clement-Lacroix P, Minet D, Fraslon-Vanhulle C, Gaillard-Kelly M, Resche-Rigon M, Baron R. A functional androgen receptor is not sufficient to allow estradiol to protect bone after gonadectomy in estradiol receptor-deficient mice. J Clin Invest. 2003 May;111(9):1319-27.
  236. Syed FA, Modder UI, Fraser DG, Spelsberg TC, Rosen CJ, Krust A, Chambon P, Jameson JL, Khosla S. Skeletal effects of estrogen are mediated by opposing actions of classical and nonclassical estrogen receptor pathways. J Bone Miner Res. 2005 Nov;20(11):1992-2001.
  237. Vidal O, Lindberg MK, Hollberg K, Baylink DJ, Andersson G, Lubahn DB, Mohan S, Gustafsson JA, Ohlsson C. Estrogen receptor specificity in the regulation of skeletal growth and maturation in male mice. Proc Natl Acad Sci U S A. 2000 May 9;97(10):5474-9.
  238. Lindberg MK, Alatalo SL, Halleen JM, Mohan S, Gustafsson JA, Ohlsson C. Estrogen receptor specificity in the regulation of the skeleton in female mice. J Endocrinol. 2001 Nov;171(2):229-36.
  239. Vandenput L, Ederveen AG, Erben RG, Stahr K, Swinnen JV, Van Herck E, Verstuyf A, Boonen S, Bouillon R, Vanderschueren D. Testosterone prevents orchidectomy-induced bone loss in estrogen receptor-alpha knockout mice. Biochem Biophys Res Commun. 2001 Jul 6;285(1):70-6.
  240. Vandenput L, Swinnen JV, Boonen S, Van Herck E, Erben RG, Bouillon R, Vanderschueren D. Role of the androgen receptor in skeletal homeostasis: the androgen-resistant testicular feminized male mouse model. J Bone Miner Res. 2004 Sep;19(9):1462-70.
  241. Venken K, De Gendt K, Boonen S, Ophoff J, Bouillon R, Swinnen JV, Verhoeven G, Vanderschueren D. Relative impact of androgen and estrogen receptor activation in the effects of androgens on trabecular and cortical bone in growing male mice: a study in the androgen receptor knockout mouse model. J Bone Miner Res. 2006 Apr;21(4):576-85.
  242. Colvard DS, Eriksen EF, Keeting PE, Wilson EM, Lubahn DB, French FS, Riggs BL, Spelsberg TC. Identification of androgen receptors in normal human osteoblast-like cells. Proc Natl Acad Sci U S A. 1989;86:854-7.
  243. Kasperk CH, Wergedal JE, Farley JR, Linkhart TA, Turner RT, Baylink DJ 1989 Androgens directly stimulate proliferation of bone cells in vitro. Endocrinology 124:1576–1578
  244. Zhang XZ, Kalu DN, Erbas B, Hopper JL, Seeman E. The effects of gonadectomy on bone size, mass, and volumetric density in growing rats are gender-, site-, and growth hormone-specific.
    J Bone Miner Res. 1999 May;14(5):802-9.
  245. Kousteni S, Bellido T, Plotkin LI, O'Brien CA, Bodenner DL, Han L, Han K, DiGregorio GB, Katzenellenbogen JA, Katzenellenbogen BS, Roberson PK, Weinstein RS, Jilka RL, Manolagas SC. Nongenotropic, sex-nonspecific signaling through the estrogen or androgen receptors: dissociation from transcriptional activity. Cell. 2001; 104:719-730.
  246. Kousteni, S. et al.2002. Reversal of bone loss in mice by nongenotropic signaling of sex steroids. Science. 298:843-846.
  247. Maccoby EE JC. The Psychology of Sex Differences. Stanford, CA: Stanford University Press, 1974.
  248. Christiansen K. Sex hormone-related variations of cognitive performance in !Kung San hunter-gatherers of Namibia. Neuropsychobiology. 1993;27(2):97-107.
  249. Barrett-Connor E, Goodman-Gruen D, Patay B. Endogenous sex hormones and cognitive function in older men. J Clin Endocrinol Metab. 1999 Oct;84(10):3681-5.
  250. Moffat SD, Zonderman AB, Metter EJ, Blackman MR, Harman SM, Resnick SM. Longitudinal assessment of serum free testosterone concentration predicts memory performance and cognitive status in elderly men. J Clin Endocrinol Metab. 2002 Nov;87(11):5001-7.
  251. Gouchie C, Kimura D. The relationship between testosterone levels and cognitive ability patterns. Psychoneuroendocrinology. 1991;16(4):323-34.
  252. Silverman I, Kastuk D, Choi J, Phillips K. Testosterone levels and spatial ability in men. Psychoneuroendocrinology. 1999 Nov;24(8):813-22.
  253. Hooven CK, Chabris CF, Ellison PT, Kosslyn SM. The relationship of male testosterone to components of mental rotation. Neuropsychologia. 2004;42(6):782-90.
  254. Moffat SD, Hampson E. A curvilinear relationship between testosterone and spatial cognition in humans: possible influence of hand preference. Psychoneuroendocrinology. 1996 Apr;21(3):323-37.
  255. Nass R, Baker S. Androgen effects on cognition: congenital adrenal hyperplasia. Psychoneuroendocrinology. 1991;16(1-3):189-201.
  256. Jones BA, Watson NV. Spatial memory performance in androgen insensitive male rats. Physiol Behav. 2005 Jun 2;85(2):135-41.
  257. Janowsky JS, Oviatt SK, Orwoll ES. Testosterone influences spatial cognition in older men. Behav Neurosci 1994;108:325-32.
  258. Vaughan C, Goldstein FC, Tenover JL. Exogenous testosterone alone or with finasteride does not improve measurements of cognition in healthy older men with low serum testosterone.
    J Androl. 2007 Nov-Dec;28(6):875-82.
  259. Cherrier MM, Anawalt BD, Herbst KL, Amory JK, Craft S, Matsumoto AM, Bremner WJ.Cognitive effects of short-term manipulation of serum sex steroids in healthy young men. J Clin Endocrinol Metab. 2002 87:3090-3096.
  260. Cherrier MM, Matsumoto AM, Amory JK, Johnson M, Craft S, Peskind ER, Raskind MA. Characterization of verbal and spatial memory changes from moderate to supraphysiological increases in serum testosterone in healthy older men. Psychoneuroendocrinology. 2007 Jan;32(1):72-9.
  261. Cherrier MM, Matsumoto AM, Amory JK, Asthana S, Bremner W, Peskind ER, Raskind MA, Craft S. Testosterone improves spatial memory in men with Alzheimer disease and mild cognitive impairment. Neurology. 2005 Jun 28;64(12):2063-8.
  262. O'Connor DB, Archer J, Hair WM, Wu FC. Activational effects of testosterone on cognitive function in men. Neuropsychologia. 2001;39(13):1385-94.
  263. Alexander GM, Swerdloff RS, Wang C, Davidson T, McDonald V, Steiner B, Hines M. Androgen-behavior correlations in hypogonadal men and eugonadal men. II. Cognitive abilities. Horm Behav. 1998 Apr;33(2):85-94.
  264. Van Goozen SH, Cohen-Kettenis PT, Gooren LJ, Frijda NH, Van de Poll NE. Gender differences in behaviour: activating effects of cross-sex hormones. Psychoneuroendocrinology 1995; 20:343-63.
  265. Van Goozen SH, Cohen-Kettenis PT, Gooren LJ, Frijda NH, Van de Poll NE. Activating effects of androgens on cognitive performance: causal evidence in a group of female-to-male transsexuals. Neuropsychologia 1994; 32:1153-7.
  266. Sherwin BB. Estrogen and/or androgen replacement therapy and cognitive functioning in surgically menopausal women. Psychoneuroendocrinology 1988;13(3):345-357.
  267. Simerly RB, Chang C, Muramatsu M, Swanson LW. Distribution of androgen and estrogen receptor mRNA-containing cells in the rat brain: an in situ hybridization study. J Comp Neurol 1990; 294:76-95
  268. Lu S, Simon NG, Wang Y, Hu S. Neural androgen receptor regulation: effects of androgen and antiandrogen. J Neurobiol 1999; 41:505-11.
  269. Lustig RH. Sex hormone modulation of neural development in vitro. Horm Behav 1994;28:383-95.
  270. Rubinow DR, Schmidt PJ. Androgens, brain, and behavior. Am J Psychiatry 1996; 153:974-84.
  271. Tobin VA, Millar RP, Canny BJ. Testosterone acts directly at the pituitary to regulate gonadotropin- releasing hormone-induced calcium signals in male rat gonadotropes. Endocrinology 1997; 138:3314-9.
  272. Gray PB, Singh AB, Woodhouse LJ, Storer TW, Casaburi R, Dzekov J, Dzekov C, Sinha-Hikim I, Bhasin S. Dose-dependent effects of testosterone on sexual function, mood, and visuospatial cognition in older men. J Clin Endocrinol Metab. 2005 Jul;90(7):3838-46.
  273. Wang C, Alexander G, Berman N, Salehian B, Davidson T, McDonald V, Steiner B, Hull L, Callegari C, Swerdloff RS. Testosterone replacement therapy improves mood in hypogonadal men--a clinical research center study. J Clin Endocrinol Metab 1996;81(10):3578-83.
  274. Pope HG Jr, Cohane GH, Kanayama G, Siegel AJ, Hudson JI. Testosterone gel supplementation for men with refractory depression: a randomized, placebo-controlled trial. Am J Psychiatry. 2003 Jan;160(1):105-11.
  275. Seidman SN, Miyazaki M, Roose SP. Intramuscular testosterone supplementation to selective serotonin reuptake inhibitor in treatment-resistant depressed men: randomized placebo-controlled clinical trial. J Clin Psychopharmacol. 2005 Dec;25(6):584-8.
  276. Seidman SN, Spatz E, Rizzo C, Roose SP. Testosterone replacement therapy for hypogonadal men with major depressive disorder: a randomized, placebo-controlled clinical trial. J Clin Psychiatry. 2001 Jun;62(6):406-12.
  277. Grinspoon S, Corcoran C, Stanley T, Baaj A, Basgoz N, Klibanski A. Effects of hypogonadism and testosterone administration on depression indices in HIV-infected men. J Clin Endocrinol Metab 2000;85:60-5.
  278. Rabkin JG, Ferrando SJ, Wagner GJ, Rabkin R. DHEA treatment for HIV+ patients: effects on mood, androgenic and anabolic parameters. Psychoneuroendocrinology 2000;25(1):53-68.
  279. Bachman ES, Basaria S, Travison TG, Guo W, Li M, Bhasin S. Temporal changes in hepcidin and erythropoietin after testosterone administration suggest a dual mechanism of action. J Gerontol A Medical Sciences 2013 in press
  280. Guo W, Bachman E, Li M, Roy CN, Blusztajn J, Wong S, Chan SY, Serra C, Jasuja R, Travison TG, Muckenthaler MU, Nemeth E, Bhasin S. Testosterone administration inhibits hepcidin transcription and is associated with increased iron incorporation into red blood cells. Aging Cell 2013 epub doi: 10.1111/acel.12052.
  281. Guo W, Li M, Bhasin S. Testosterone supplementation improves anemia in aging male mice. J Gerontol: Biological Sciences 2013 in press
  282. Guo W, Wong S, Li M, Liang W, Liesa M, Serra C, Jasuja R, Bartke A, Kirkland JL, Shirihai O, Bhasin S. Testosterone plus low-intensity physical training in late life improves functional performance, skeletal muscle mitochondrial biogenesis, and mitochondrial quality control in male mice. PLoS One. 2012;7(12):e51180. doi: 10.1371/journal.pone.0051180.
  283. Kouri EM, Lukas SE, Pope HG Jr, Oliva PS. Increased aggressive responding in male volunteers following the administration of gradually increasing doses of testosterone cypionate. Drug Alcohol Depend. 1995;40:73-9.
  284. Pope HG Jr, Katz DL. Affective and psychotic symptoms associated with anabolic steroid use. Am J Psychiatry 1988;145:487-90.
  285. Wilson IB, Roubenoff R, Knox TA, Spiegleman D, Gorbach SL. Relation of lean body mass to health-related quality of life in persons with HIV. J Acquir Immune Defic Syndrome 2000;24:137-46.
  286. Litwin MS, Nied RJ, Dhandani N. Health-related quality of life in men with erectile dysfunction. J Gen Intern Med 1998;13:159-166.
  287. Morley JE, Charlton E, Patrick P, Kaiser FE, Cadeau P, McCready D, Perry HM 3rd. Validation of a screening questionnaire for androgen deficiency in aging males. Metabolism. 2000 Sep;49(9):1239-42.
  288. T'Sjoen G, Goemaere S, De Meyere M, Kaufman JM. Perception of males' aging symptoms, health and well-being in elderly community-dwelling men is not related to circulating androgen levels. Psychoneuroendocrinology. 2004 Feb;29(2):201-14.
  289. Smith KW, Feldman HA, McKinlay JB. Construction and field validation of a self-administered screener for testosterone deficiency (hypogonadism) in ageing men. Clin Endocrinol (Oxf). 2000 Dec;53(6):703-11.
  290. Calof OM, Singh AB, Lee ML, Kenny AM, Urban RJ, Tenover JL, Bhasin S. Adverse events associated with testosterone replacement in middle-aged and older men: a meta-analysis of randomized, placebo-controlled trials. J Gerontol A Biol Sci Med Sci. 2005 Nov;60(11):1451-7.
  291. Rolf C, Nieschlag E. Potential adverse effects of long-term testosterone therapy. Baillieres Clin Endocrinol Metab 1998; 12:521-34.
  292. Liu PY, Death AK, Handelsman DJ. Androgens and cardiovascular disease. Endocr Rev. 2003 Jun;24(3):313-40.
  293. Haffner SM, Laakso M, Miettinen H, Mykkanen L, Karhapaa P, Rainwater DL. Low levels of sex hormone-binding globulin and testosterone are associated with smaller, denser low density lipoprotein in normoglycemic men [see comments]. J Clin Endocrinol Metab 1996; 81:3697-701.
  294. Khaw KT, Barrett-Connor E. Endogenous sex hormones, high density lipoprotein cholesterol, and other lipoprotein fractions in men. Arterioscler Thromb. 1991 May-Jun;11(3):489-94.
  295. De Pergola G, De Mitrio V, Sciaraffia M, et al. Lower androgenicity is associated with higher plasma levels of prothrombotic factors irrespective of age, obesity, body fat distribution, and related metabolic parameters in men. Metabolism 1997; 46:1287-93.
  296. Crist DM, Peake GT, Stackpole PJ. Lipemic and lipoproteinemic effects of natural and synthetic androgens in humans. Clin Exp Pharmacol Physiol 1986; 13:513-8.
  297. Herbst KL, Amory JK, Brunzell JD, Chansky HA, Bremner WJ. Testosterone administration to men increases hepatic lipase activity and decreases HDL and LDL size in 3 wk. Am J Physiol Endocrinol Metab. 2003 284:E1112-E1118.
  298. Hurley BF, Seals DR, Hagberg JM, et al. High-density-lipoprotein cholesterol in bodybuilders v powerlifters. Negative effects of androgen use. JAMA 1984; 252:507-13.
  299. Thompson PD, Cullinane EM, Sady SP, et al. Contrasting effects of testosterone and stanozolol on serum lipoprotein levels. JAMA 1989; 261:1165-8.
  300. Langer C, Gansz B, Goepfert C, Engel T, Uehara Y, von Dehn G, Jansen H, Assmann G, von Eckardstein A: Testosterone up-regulates scavenger receptor BI and stimulates cholesterol efflux from macrophages. Biochem Biophys Res Commun 2002 296:1051–1057.
  301. Glueck CJ, Glueck HI, Stroop D, Speirs J, Hamer T, Tracy T: Endogenous testosterone, fibrinolysis, and coronary heart disease risk in hyperlipidemic men. J Lab Clin Med 1993 122:412–420.
  302. Ng MK, Liu PY, Williams AJ, Nakhla S, Ly LP, Handelsman DJ, Celermajer DS: Prospective study of effect of androgens on serum inflammatory markers in men. Arterioscler Thromb Vasc Biol 2002 22:1136–1141.
  303. Singh AB, Hsia S, Alaupovic P, Sinha-Hikim I, Woodhouse L, Buchanan TA, Shen R, Bross R, Berman N, Bhasin S. The effects of varying doses of T on insulin sensitivity, plasma lipids, apolipoproteins, and C-reactive protein in healthy young men. J Clin Endocrinol Metab. 2002 Jan;87(1):136-43.
  304. Phillips GB, Pinkernell BH, Jing TY. The association of hypotestosteronemia with coronary artery disease in men. Arterioscler Thromb 1994;14:701-6.
  305. Tibblin G, Adlerberth A, Lindstedt G, Bjorntorp P. The pituitary-gonadal axis and health in elderly men: a study of men born in 1913. Diabetes 1996; 45:1605-9.
  306. Arnlov J, Pencina MJ, Amin S, Nam BH, Benjamin EJ, Murabito JM, Wang TJ, Knapp PE, D'Agostino RB Sr, Bhasin S, Vasan RS. Endogenous sex hormones and cardiovascular disease incidence in men. Ann Intern Med. 2006 Aug 1;145(3):176-84.
  307. Wu SZ, Weng XZ. Therapeutic effects of an androgenic preparation on myocardial ischemia and cardiac function in 62 elderly male coronary heart disease patients. Chin Med J (Engl) 1993; 106:415-8.
  308. Jaffe MD. Effect of testosterone cypionate on postexercise ST segment depression. Br Heart J. 1977;39:1217-22.
  309. Ong PJ, Patrizi G, Chong WC, Webb CM, Hayward CS, Collins P. Testosterone enhances flow-mediated brachial artery reactivity in men with coronary artery disease. Am J Cardiol 2000 Jan 15;85(2):269-72.
  310. Rosano GM, Leonardo F, Pagnotta P, Pelliccia F, Panina G, Cerquetani E, della Monica PL, Bonfigli B, Volpe M, Chierchia SL. Acute anti-ischemic effect of testosterone in men with coronary artery disease. Circulation 1999:1666-70.
  311. Webb CM, Adamson DL, de Zeigler D, Collins P. Effect of acute testosterone on myocardial ischemia in men with coronary artery disease. Am J Cardiol 1999;83(3):437-9.
  312. English KM, Steeds RP, Jones TH, Diver MJ, Channer KS: Low-dose transdermal testosterone therapy improves angina threshold in men with chronic stable angina: a randomized, double-blind, placebo-controlled study. Circulation 102:1906–1911, 2000.
  313. Thompson PD, Ahlberg AW, Moyna NM, Duncan B, Ferraro-Borgida M, White CM, McGill CC, Heller GV: Effect of intravenous testosterone on myocardial ischemia in men with coronary artery disease. Am Heart J 143:249–256, 2002
  314. Yue P, Chatterjee K, Beale C, Poole-Wilson PA, Collins P: Testosterone relaxes rabbit coronary arteries and aorta. Circulation 1995 91:1154–1160.
  315. Nathan L, Shi W, Dinh H, Mukherjee TK, Wang X, Lusis AJ, Chaudhuri G. Testosterone inhibits early atherogenesis by conversion to estradiol: critical role of aromatase. Proc Natl Acad Sci U S A. 2001 98:3589-3593.
  316. Alexandersen P, Haarbo J, Byrjalsen I, Lawaetz H, Christiansen C: Natural androgens inhibit male atherosclerosis: a study in castrated, cholesterol-fed rabbits. Circ Res 1999 84:813–819.
  317. Fernández-Balsells MM, Murad MH, Lane M, Lampropulos JF, Albuquerque F, Mullan RJ, Agrwal N, Elamin MB, Gallegos-Orozco JF, Wang AT, Erwin PJ, Bhasin S, Montori VM. Clinical review 1: Adverse effects of testosterone therapy in adult men: a systematic review and meta-analysis. J Clin Endocrinol Metab. 2010 Jun;95(6):2560-75.
  318. Xu L, Freeman G, Cowling BJ, Schooling CM. Testosterone therapy and cardiovascular events among men: a systematic review and meta-analysis of placebo-controlled randomized trials. BMC Med. 2013 Apr 18;11:108.
  319. Vigen R, O'Donnell CI, Barón AE, Grunwald GK, Maddox TM, Bradley SM, Barqawi A, Woning G, Wierman ME, Plomondon ME, Rumsfeld JS, Ho PM. Association of testosterone therapy with mortality, myocardial infarction, and stroke in men with low testosterone levels. JAMA. 2013 Nov 6;310(17):1829-36.
  320. Shores, M.M., Smith, N.L., Forsberg, C.W., Anawalt, B.D. & Matsumoto, A.M. Testosterone treatment and mortality in men with low testosterone levels. J Clin Endocrinol Metab 97, 2050-8 (2012).
  321. Sattler, F.R. et al. Testosterone and growth hormone improve body composition and muscle performance in older men. J Clin Endocrinol Metab 94, 1991-2001 (2009).
  322. Mauras N, Hayes V, Welch S, et al. Testosterone deficiency in young men: marked alterations in whole body protein kinetics, strength, and adiposity. J Clin Endocrinol Metab 1998; 83:1886-92.
  323. Khaw KT, Barrett-Connor E. Lower endogenous androgens predict central adiposity in men. Ann Epidemiol 1992; 2:675-82.
  324. Seidell JC, Bjorntorp P, Sjostrom L, Kvist H, Sannerstedt R. Visceral fat accumulation in men is positively associated with insulin, glucose, and C-peptide levels, but negatively with testosterone levels. Metabolism 1990;
  325. Mårin P, Odén B, Björntorp P. Assimilation and mobilization of triglycerides in subcutaneous abdominal and femoral adipose tissue in vivo in men: Effects of androgens. J Clin Endocrinol Metab 1995; 80:239-243.
  326. Holmang A, Bjorntorp P. The effects of testosterone on insulin sensitivity in male rats. Acta Physiol Scand 1992; 146:505-10.
  327. Hobbs CJ, Jones RE, Plymate SR. Nandrolone, a 19-nortestosterone, enhances insulin-independent glucose uptake in normal men [see comments]. J Clin Endocrinol Metab 1996; 81:1582-5.
  328. Ramirez ME, McMurry MP, Wiebke GA. Evidence for sex steroid inhibition of lipoprotein lipase in men: comparison of abdominal and femoral adipose tissue. Metabolism 1997;46:179-85.
  329. Defay R, Papoz L, Barny S, Bonnot-Lours S, Caces E, Simon D. Hormonal status and NIDDM in the European and Melanesian populations of New Caledonia: a case-control study. The Caledonia Diabetes Mellitus (CALDIA) Study Group. Int J Obes Relat Metab Disord 1998; 22:927-34.
  330. Haffner SM, Shaten J, Stern MP, Smith GD, Kuller L. Low levels of sex hormone-binding globulin and testosterone predict the development of non-insulin-dependent diabetes mellitus in men. MRFIT Research Group. Multiple Risk Factor Intervention Trial. Am J Epidemiol 1996;143:889-97.
  331. Stellato RK, Feldman HA, Hamdy O, Horton ES, McKinlay JB. Testosterone, sex hormone-binding globulin, and the development of type 2 diabetes in middle-aged men: prospective results from the Massachusetts male aging study. Diabetes Care 2000;23(4):490-4.
  332. Oh JY, Barrett-Connor E, Wedick NM, Wingard DL; Rancho Bernardo Study. Endogenous sex hormones and the development of type 2 diabetes in older men and women: the Rancho Bernardo study. Diabetes Care. 2002 Jan;25(1):55-60.
  333. Endre T, Mattiasson I, Berglund G, Hulthen UL. Low testosterone and insulin resistance in hypertension-prone men. J Hum Hypertens 1996; 10:755-61.
  334. Haffner SM, Shaten J, Stern MP, Smith GD, Kuller L. Low levels of sex hormone-binding globulin and testosterone predict the development of non-insulin-dependent diabetes mellitus in men. MRFIT Research Group. Multiple Risk Factor Intervention Trial. Am J Epidemiol. 1996;143(9):889-97.
  335. Stellato RK, Feldman HA, Hamdy O, Horton ES, McKinlay JB. Testosterone, sex hormone-binding globulin, and the development of type 2 diabetes in middle-aged men: prospective results from the Massachusetts male aging study. Diabetes Care. 2000;23(4):490-4.
  336. Oh J-Y, Barrett-Connor E, Wedick NM, Wingard DL. Endogenous Sex Hormones and the Development of Type 2 Diabetes in Older Men and Women: the Rancho Bernardo Study. Diabetes Care. 2002;25(1):55-60.
  337. Svartberg J, Jenssen T, Sundsfjord J, Jorde R. The associations of endogenous testosterone and sex hormone-binding globulin with glycosylated hemoglobin levels, in community dwelling men. The Tromso Study. Diabetes Metab. 2004;30(1):29-34.
  338. Laaksonen DE, Niskanen L, Punnonen K, et al. Testosterone and sex hormone-binding globulin predict the metabolic syndrome and diabetes in middle-aged men. Diabetes Care. 2004;27(5):1036-41.
  339. Selvin E, Feinleib M, Zhang L, et al. Androgens and Diabetes in Men: Results from the Third National Health and Nutrition Examination Survey (NHANES III). Diabetes Care. 2007;30(2):234-238.
  340. Ding EL, Song Y, Malik VS, Liu S. Sex Differences of Endogenous Sex Hormones and Risk of Type 2 Diabetes: A Systematic Review and Meta-analysis. JAMA. 2006;295(11):1288-1299.
  341. Kapoor D, Aldred H, Clark S, Channer KS, Jones TH. Clinical and Biochemical Assessment of Hypogonadism in Men With Type 2 Diabetes: Correlations with bioavailable testosterone and visceral adiposity. Diabetes Care. 2007;30(4):911-917.
  342. Svartberg J, Schirmer H, Wilsgaard T, Mathiesen EB, Njølstad I, Løchen ML, Jorde R. Single-nucleotide polymorphism, rs1799941 in the sex hormone-binding globulin (SHBG) gene, related to both serum testosterone and SHBG levels and the risk of myocardial infarction, type 2 diabetes, cancer and mortality in men: the Tromsø Study. Andrology. 2013 Dec 10. doi: 10.1111/j.2047-2927.2013.00174.x. [Epub ahead of print]
  343. Perry JR, Weedon MN, Langenberg C, Jackson AU, Lyssenko V, Sparsø T, Thorleifsson G, Grallert H, Ferrucci L, Maggio M, Paolisso G, Walker M, Palmer CN, Payne F, Young E, Herder C, Narisu N, Morken MA, Bonnycastle LL, Owen KR, Shields B, Knight B, Bennett A, Groves CJ, Ruokonen A, Jarvelin MR, Pearson E, Pascoe L, Ferrannini E, Bornstein SR, Stringham HM, Scott LJ, Kuusisto J, Nilsson P, Neptin M, Gjesing AP, Pisinger C, Lauritzen T, Sandbaek A, Sampson M; MAGIC, Zeggini E, Lindgren CM, Steinthorsdottir V, Thorsteinsdottir U, Hansen T, Schwarz P, Illig T, Laakso M, Stefansson K, Morris AD, Groop L, Pedersen O, Boehnke M, Barroso I, Wareham NJ, Hattersley AT, McCarthy MI, Frayling TM. Genetic evidence that raised sex hormone binding globulin (SHBG) levels reduce the risk of type 2 diabetes. Hum Mol Genet. 2010 Feb 1;19(3):535-44. doi: 10.1093/hmg/ddp522.
  344. Saltiki K, Stamatelopoulos K, Voidonikola P, Lazaros L, Mantzou E, Georgiou I, Anastasiou E, Papamichael C, Alevizaki M. Association of the SHBG gene promoter polymorphism with early markers of atherosclerosis in apparently healthy women. Atherosclerosis. 2011 Nov;219(1):205-10.
  345. Lakshman KM, Bhasin S, Araujo AB Sex hormone-binding globulin as an independent predictor of incident type 2 diabetes mellitus in men. J Gerontol A Biol Sci Med Sci. 2010 May;65(5):503-9.
  346. Yialamas MA, Dwyer AA, Hanley E, Lee H, Pitteloud N, Hayes FJ. Acute sex steroid withdrawal reduces insulin sensitivity in healthy men with idiopathic hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 2007 Nov;92(11):4254-9.
  347. Basaria S, Muller DC, Carducci MA, Egan J, Dobs AS. Relation between duration of androgen deprivation therapy and degree of insulin resistance in men with prostate cancer. Arch Intern Med. 2007 Mar 26;167(6):612-3.
  348. Jones, T.H. et al. Testosterone replacement in hypogonadal men with type 2 diabetes and/or metabolic syndrome (the TIMES2 study). Diabetes Care 34, 828-37 (2011).
  349. Corona, G. et al. Type 2 diabetes mellitus and testosterone: a meta-analysis study. Int J Androl 34, 528-40 (2011).
  350. Basu R, Dalla Man C, Campioni M, Basu A, Nair KS, Jensen MD, Khosla S, Klee G, Toffolo G, Cobelli C, Rizza RA. Effect of 2 years of testosterone replacement on insulin secretion, insulin action, glucose effectiveness, hepatic insulin clearance, and postprandial glucose turnover in elderly men. Diabetes Care. 2007 Aug;30(8):1972-8.
  351. Singh AB, Hsia S, Alaupovic P, Sinha-Hikim I, Woodhouse L, Buchanan TA, Shen R, Bross R, Berman N, Bhasin S. The effects of varying doses of T on insulin sensitivity, plasma lipids, apolipoproteins, and C-reactive protein in healthy young men. J Clin Endocrinol Metab. 2002 Jan;87(1):136-43.
  352. Allan CA, Strauss BJ, Burger HG, Forbes EA, McLachlan RI. Testosterone therapy prevents gain in visceral adipose tissue and loss of skeletal muscle in nonobese aging men. J Clin Endocrinol Metab. 2008 Jan;93(1):139-46.
  353. Bhasin S, Singh AB, Mac RP, Carter B, Lee MI, Cunningham GR. Managing the risks of prostate disease during testosterone replacement therapy in older men: recommendations for a standardized monitoring plan. J Androl. 2003 May-Jun;24(3):299-311.
  354. Fowler, J. E., Jr. and W. F. Whitmore, Jr. (1981). The response of metastatic adenocarcinoma of the prostate to exogenous testosterone. J Urol 126(3): 372-5.
  355. Manni, A., M. Bartholomew, R. Caplan, A. Boucher, R. Santen, A. Lipton, H. Harvey, M. Simmonds, D. White-Hershey, R. Gordon. (1988). Androgen priming and chemotherapy in advanced prostate cancer: evaluation of determinants of clinical outcome. J Clin Oncol 6(9): 1456-66.
  356. Babaian R. J., D. A. Johnston, W. Naccarato, A. Ayala, V. A. Bhadkamkar, H. H. Fritsche. (2001). The incidence of prostate cancer in a screening population with a serum prostate specific antigen between 2.5 and 4.0 ng/ml: relation to biopsy strategy. J Urol 165(3): 757-60.
  357. Gatling, R. R. (1990). Prostate carcinoma: an autopsy evaluation of the influence of age, tumor grade, and therapy on tumor biology. South Med J 83(7): 782-4.
  358. Greenlee, R. T., T. Murray, S. Bolden, P. A. Wingo. (2000). Cancer statistics, 2000. CA Cancer J Clin 50(1): 7-33.
  359. Guileyardo, J. M., W. D. Johnson, R. A. Welsh, K. Akazaki, P. Correa. (1980). Prevalence of latent prostate carcinoma in two U.S. populations. J Natl Cancer Inst 65(2): 311-6.
  360. Hsing, A. W., L. Tsao and S. S. Devesa (2000). International trends and patterns of prostate cancer incidence and mortality. Int J Cancer 85(1): 60-7.
  361. Hsing, A. W. and S. S. Devesa (2001). Trends and patterns of prostate cancer: what do they suggest? Epidemiol Rev 23(1): 3-13.
  362. Landis, S. H., T. Murray, S. Bolden, P. A. Wingo. (1998). Cancer statistics, 1998. CA Cancer J Clin 48(1): 6-29.
  363. Landis, S. H., T. Murray, S. Bolden, P. A. Wingo. (1999). Cancer statistics, 1999. CA Cancer J Clin 49(1): 8-31, 1.
  364. Holmang, S., P. Marin, G. Lindstedt, H. Hedelin. (1993). Effect of long-term oral testosterone undecanoate treatment on prostate volume and serum prostate-specific antigen concentration in eugonadal middle-aged men. Prostate 23(2): 99-106.
  365. Morgentaler A, Bruning CO 3rd, DeWolf WC.  Occult prostate cancer in men with low serum testosterone levels. JAMA 1996 Dec 18;276(23):1904-6
  366. Hoffman, M. A., W. C. DeWolf and A. Morgentaler (2000). Is low serum free testosterone a marker for high grade prostate cancer? J Urol 163(3): 824-7.
  367. Schatzl, G., S. Madersbacher, T. Thurridl, J. Waldmuller, G. Kramer, A. Haitel, M. Marberger. (2001). High-grade prostate cancer is associated with low serum testosterone levels. Prostate 47(1): 52-8.
  368. Ahluwalia, B., M. A. Jackson, G. W. Jones, A. O. Williams, M. S. Rao, S. Rajguru. (1981). Blood hormone profiles in prostate cancer patients in high-risk and low- risk populations. Cancer 48(10): 2267-73.
  369. Andersson, S. O., H. O. Adami, R. Bergstrom, L. Wide. (1993). Serum pituitary and sex steroid hormone levels in the etiology of prostatic cancer--a population-based case-control study. Br J Cancer 68(1): 97-102.
  370. Demark-Wahnefried, W., S. M. Lesko, M. R. Conaway, C. N. Robertson, R. V. Clark, B. Lobaugh, B. J. Mathias, T. S. Strigo, D. F. Paulson. (1997). Serum androgens: associations with prostate cancer risk and hair patterning. J Androl 18(5): 495-500.
  371. Dorgan, J. F., D. Albanes, J. Virtamo, O. P., Heinonen, D. W. Chandler, M. Galmarini, L. M. McShane, M. J. Barrett, J. Tangrea, P. R. Taylor. (1998). Relationships of serum androgens and estrogens to prostate cancer risk: results from a prospective study in Finland. Cancer Epidemiol Biomarkers Prev 7(12): 1069-74.
  372. Eaton, N. E., G. K. Reeves, P. N. Appleby, T. J. Key. (1999). Endogenous sex hormones and prostate cancer: a quantitative review of prospective studies. Br J Cancer 80(7): 930-4.
  373. Gann, P. H., C. H. Hennekens, J. Ma, C. Longcope, M. J. Stampfer. (1996). Prospective study of sex hormone levels and risk of prostate cancer. J Natl Cancer Inst 88(16): 1118-26.
  374. Hawk, E., R. A. Breslow and B. I. Graubard (2000). Male pattern baldness and clinical prostate cancer in the epidemiologic follow-up of the first National Health and Nutrition Examination Survey. Cancer Epidemiol Biomarkers Prev 9(5): 523-7.
  375. Heikkila, R., K. Aho, M. Heliovaara, M. Hakama, J. Marniemi, A. Reunanen, P. Knekt. (1999). Serum testosterone and sex hormone-binding globulin concentrations and the risk of prostate carcinoma: a longitudinal study. Cancer 86(2): 312-5.
  376. Hsing, A. W. and G. W. Comstock (112). Serological precursors of cancer: serum hormones and risk of subsequent prostate cancer. Cancer Epidemiol Biomarkers Prev 2(1): 27-32.
  377. Hulka, B. S., J. E. Hammond, G. DiFerdinando, D. D. Mickey, F. A. Fried, H. Checkoway, W. E. Stumpf, W. C. Beckman, Jr., T. D. Clark. (76). Serum hormone levels among patients with prostatic carcinoma or benign prostatic hyperplasia and clinic controls. Prostate 11(2): 171-82.
  378. Nomura, A., L. K. Heilbrun, G. N. Stemmermann, H. L. Judd. (1988). Prediagnostic serum hormones and the risk of prostate cancer. Cancer Res 48(12): 3515-7.
  379. Nomura, A. M., G. N. Stemmermann, P. H. Chyou, B. E. Henderson, F. Z. Stanczyk. (1996). Serum androgens and prostate cancer. Cancer Epidemiol Biomarkers Prev 5(8): 621-5.
  380. Shaneyfelt, T., R. Husein, G. Bubley, C. S. Mantzoros. (2000). Hormonal predictors of prostate cancer: a meta-analysis. J Clin Oncol 18(4): 847-53.
  381. Slater S, Oliver RT. Testosterone: its role in development of prostate cancer and potential risk from use as hormone replacement therapy. Drugs Aging 2000: 17:431-9.
  382. Vatten, L. J., G. Ursin, R. K. Ross, F. Z. Stanczyk, R. A. Lobo, S. Harvei, E. Jellum. (1997). Androgens in serum and the risk of prostate cancer: a nested case- control study from the Janus serum bank in Norway. Cancer Epidemiol Biomarkers Prev 6: 967-969.
  383. Wolk, A, Andersson OS, Bergstrom R. 1997 Prospective study of sex hormone levels and risk of prostate cancer. J Natl Cancer Inst 89: 820.
  384. Barrett-Connor E, Garland C, McPhillips JB, Khaw KT, Wingard DL. 1990 A prospective, population-based study of androstenedione, estrogens, and prostatic cancer. Cancer Res 50: 169-173.
  385. Mohr, B. A., H. A. Feldman, L. A. Kalish, C. Longcope, J. B. McKinlay. (2001). Are serum hormones associated with the risk of prostate cancer? Prospective results from the Massachusetts Male Aging Study. Urology 57: 930-935.
  386. Platz EA, Leitzmann MF, Rifai N, Kantoff PW, Chen YC, Stampfer MJ, Willett WC, Giovannucci E. Sex steroid hormones and the androgen receptor gene CAG repeat and subsequent risk of prostate cancer in the prostate-specific antigen era. Cancer Epidemiol Biomarkers Prev. 2005 May;14(5):1262-9.
  387. Marks LS, Mazer NA, Mostaghel E, Hess DL, Dorey FJ, Epstein JI, Veltri RW, Makarov DV, Partin AW, Bostwick DG, Macairan ML, Nelson PS. Effect of testosterone replacement therapy on prostate tissue in men with late-onset hypogonadism: a randomized controlled trial. JAMA 2006;296:2351-61.
  388. Page ST, Lin DW, Mostaghel EA, Hess DL, True LD, Amory JK, Nelson PS, Matsumoto AM, Bremner WJ. Persistent intraprostatic androgen concentrations after medical castration in healthy men. J Clin Endocrinol Metab 2006;91:3850-6.
  389. Mostaghel EA, Page ST, Lin DW, Fazli L, Coleman IM, True LD, Knudsen B, Hess DL, Nelson CC, Matsumoto AM, Bremner WJ, Gleave ME, Nelson PS. Intraprostatic androgens and androgen-regulated gene expression persist after testosterone suppression: therapeutic implications for castration-resistant prostate cancer. Cancer Res. 2007 May 15;67(10):5033-41.
  390. Behre HM, B. J., Nieschlag E. (1994). Prostate volume in testosterone-treated and untreated hypogonadal men in comparison to age-matched normal controls. Clin Endocrinol (Oxf) 40: 341-349.
  391. Cooper CS, Perry PJ, Sparks AE, MacIndoe JH, Yates WR, Williams RD. Effect of exogenous testosterone on prostate volume, serum and semen prostate specific antigen levels in healthy young men. J Urol 1998;159:441-443.
  392. Douglas, T. H., R. R. Connelly, D. G. McLeod, S. J. Ericsson, R. Barren, 3rd, G. P. Murphy. (1995). Effect of exogenous testosterone replacement on prostate-specific antigen and prostate-specific membrane antigen levels in hypogonadal men. J Surg Oncol 59(4): 246-50.
  393. Guay, A. T., J. B. Perez, W. A. Fitaihi, M. Vereb. (2000). Testosterone treatment in hypogonadal men: prostate-specific antigen level and risk of prostate cancer. Endocr Pract 6(2): 132-8.
  394. Meikle AW, Arver S, Dobs AS, Adolfsson J, Sanders SW, Middleton RG et al. Prostate size in hypogonadal men treated with a nonscrotal permeation- enhanced testosterone transdermal system. Urology 1997; 49:191-196.
  395. Sasagawa I, Nakada T, Kazama T, Satomi S, Terada T, Katayama T. Volume change of the prostate and seminal vesicles in male hypogonadism after androgen replacement therapy. Int Urol Nephrol 1990; 22(3):279-284.
  396. Shibasaki, T., I. Sasagawa, Y. Suzuki, H. Yazawa, O. Ichiyanagi, S. Matsuki, M. Miura, T. Nakada. (2001). Effect of testosterone replacement therapy on serum PSA in patients with Klinefelter syndrome. Arch Androl 47(3): 173-6.
  397. Svetec DA, Canby ED, Thompson IM, Sabanegh ES, Jr. The effect of parenteral testosterone replacement on prostate specific antigen in hypogonadal men with erectile dysfunction. J Urol 1997;158:1775-1777.
  398. Gormley, G. J., E. Stoner, R. C. Bruskewitz, J. Imperato-McGinley, P. C. Walsh, J. D. McConnell, G. L. Andriole, J. Geller, B. R. Bracken, J. S. Tenover. (1992). “The effect of finasteride in men with benign prostatic hyperplasia. The Finasteride Study Group [see comments].” N Engl J Med 327(17): 1185-91.
  399. Thompson IM, Goodman PJ, Tangen CM, Lucia MS, Miller GJ, Ford LG, Lieber MM, Cespedes RD, Atkins JN, Lippman SM, Carlin SM, Ryan A, Szczepanek CM, Crowley JJ, Coltman CA Jr. The influence of finasteride on the development of prostate cancer. N Engl J Med. 2003;349:215-24.
  400. Manseck, A., C. Pilarsky, S. Froschermaier, M. Menschikowski, M. P. Wirth. (1998). Diagnostic significance of prostate-specific antigen velocity at intermediate PSA serum levels in relation to the standard deviation of different test systems. Urol Int 60(1): 25-27.
  401. Riehmann, M., P. R. Rhodes, T. D. Cook, G. S. Grose, R. C. Bruskewitz. (1993). Analysis of variation in prostate-specific antigen values. Urology 42(4): 390-7.
  402. Wenzl, T. B., M. Riehmann, T. D. Cook, R. C. Bruskewitz. (1998). Variation in PSA values in patients without malignancy. Urology 52(5): 932.
  403. Carter, H. B. and J. D. Pearson (1993). PSA velocity for the diagnosis of early prostate cancer. A new concept. Urol Clin North Am 20(4): 665-670.
  404. Carter HB, Pearson JD, Waclawiw Z, Metter EJ, Chan DW, Guess HA, Walsh PC. Prostate-specific antigen variability in men without prostate cancer: effect of sampling interval on prostate-specific antigen velocity. Urology 1995(b);45:591-6.
  405. Fang, J., E. J. Metter, P. Landis, H. B. Carter. (2002). PSA velocity for assessing prostate cancer risk in men with PSA levels between 2.0 and 4.0 ng/ml. Urology 59(6): 889-93; discussion 893-894.
  406. Fang, J., E. J. Metter, P. Landis, D. W. Chan, C. H. Morrell, H. B. Carter. (2001). Low levels of prostate-specific antigen predict long-term risk of prostate cancer: results from the Baltimore Longitudinal Study of Aging. Urology 58(3): 411-6.
  407. Bhasin S, Woodhouse L, Casaburi R, Singh AB, Mac RP, Lee M, Yarasheski KE, Sinha-Hikim I, Dzekov C, Dzekov J, Magliano L, Storer TW. Older men are as responsive as young men to the anabolic effects of graded doses of testosterone on the skeletal muscle. J Clin Endocrinol Metab. 2005 Feb;90(2):678-88.
  408. Wu FC, Farley TM, Peregoudov A, Waites GM. Effects of testosterone enanthate in normal men: experience from a multicenter contraceptive efficacy study. World Health Organization Task Force on Methods for the Regulation of Male Fertility. Fertil Steril 1996 Mar;65(3):626-36.
  409. Carter HB, Pearson JD, Metter EJ, Chan DW, Andres R, Fozard JL et al. Longitudinal evaluation of serum androgen levels in men with and without prostate cancer. Prostate 1995(a);27:25-31.
  410. Fried W, Marver D, Lange RD, Gurney CW. Studies on the erythropoietic stimulating factor in the plasma of mice after receiving testosterone. J Lab Clin Med 1966; 68:947-951.
  411. Rencricca NJ, Solomon J, Fimian WJ, Jr., Howard D, Rizzoli V, Stohlman F, Jr. The effect of testosterone on erythropoiesis. Scand J Haematol 1969; 6(6):431-436.
  412. Naets JP, Wittek M. The mechanism of action of androgens on erythropoiesis. Ann N Y Acad Sci 1968; 149(1):366-376.
  413. Coviello A, Lakshman K, Kaplan B, Singh AB, Chen T, Bhasin S. Dose-dependent effects of testosterone on erythropoiesis in young and older men. J Clin Endocrinol Metab 2008;93:914-9.
  414. Dobs AS, Meikle AW, Arver S, Sanders SW, Caramelli KE, Mazer NA. Pharmacokinetics, efficacy, and safety of a permeation-enhanced testosterone transdermal system in comparison with bi-weekly injections of testosterone enanthate for the treatment of hypogonadal men. J Clin Endocrinol Metab 1999; 84:3469-3478.

417. Bachman E, Feng R, Travison T, Li M, Olbina G, Ostland V, Ulloor J, Zhang A, Basaria S, Ganz T, Westerman M, Bhasin S. Testosterone suppresses hepcidin in men: a potential mechanism for testosterone-induced erythrocytosis. J Clin Endocrinol Metab 2010;95(10):4743-7.

418. Cistulli PA, Grunstein RR, Sullivan CE. Effect of testosterone administration on upper airway collapsibility during sleep. Am J Respir Crit Care Med 1994; 149:530-532.41Emery MJ, Hlastala MP, Matsumoto AM. Depression of hypercapnic ventilatory drive by testosterone in the sleeping infant primate. J Appl Physiol 1994; 76(4):1786-1793.

419.Emery MJ, Hlastala MP, Matsumoto AM. Depression of hypercapnic ventilatory drive by testosterone in the sleeping infant primate. J Appl Physiol 1994; 76(4):1786-1793.

420. Schiavi RC, White D, Mandeli J. Pituitary-gonadal function during sleep in healthy aging men. Psychoneuroendocrinology. 1992 Nov;17(6):599-609.

421. Hoyos, C.M., Killick, R., Yee, B.J., Grunstein, R.R. & Liu, P.Y. Effects of testosterone therapy on sleep and breathing in obese men with severe obstructive sleep apnoea: a randomized placebo-controlled trial. Clin Endocrinol (Oxf) 77, 599-607 (2012).

422.Matsumoto AM, Sandblom RE, Schoene RB, Lee KA, Giblin EC, Pierson DJ et al. Testosterone replacement in hypogonadal men: effects on obstructive sleep apnoea, respiratory drives, and sleep. Clin Endocrinol (Oxf) 1985; 22(6):713-721.

423.Schneider BK, Pickett CK, Zwillich CW, Weil JV, McDermott MT, Santen RJ et al. Influence of testosterone on breathing during sleep. J Appl Physiol 1986; 61(2):618-623.

424.Hoyos, C.M. et al. Body compositional and cardiometabolic effects of testosterone therapy in obese men with severe obstructive sleep apnoea: a randomised placebo-controlled trial. Eur J Endocrinol 167, 531-41 (2012).

425.Liu, P.Y. et al. The short-term effects of high-dose testosterone on sleep, breathing, and function in older men. J Clin Endocrinol Metab 88, 3605-13 (2003).

426.Kirbas, G. et al. Obstructive sleep apnoea, cigarette smoking and serum testosterone levels in a male sleep clinic cohort. J Int Med Res 35, 38-45 (2007).

427.Seymour FI, Duffy C, Korner A. 2002 A case of authentic fertility in a man of 94.  JAMA 105:1423-1424.

428.Plas E, Berger P, Hermann M, Pfluger H. 2000 Effects of aging on male fertility? Exp Gerontol 35:543-551.

429.Rolf C, Nieschlag E. 2001 Reproductive functions, fertility and genetic risks of ageing men. Exp Clin Endocrinol Diabetes 109:68-74.

430.American Fertility Society 1990: New guidelines for the use of semen donor insemination. Fertil Steril 53 (Suppl 1): 1S-13S.

431.American Society for Reproductive Medicine1998: Guidelines for gamete and embryo donation. http://www.asrm.org/Media/Practice/gamete.html

432..Silber SJ. 1991 Effects of age on male fertility. Semin Reprod Endocrinol 9:241-248.

433.Kidd SA, Eskenazi B, Wyrobek AJ. 2001 Effects of male age on semen quality and fertility: a review of the literature. Fertil Steril 75:237-248.

434.Sloter E, Nath J, Eskenazi B, Wyrobek AJ. Effects of male age on the frequencies of germinal and heritable chromosomal abnormalities in humans and rodents. Fertil Steril. 2004;81:925-43.

435.Eskenazi B, Wyrobek AJ, Sloter E, Kidd SA, Moore L, Young S, Moore D.  2003 The association of age and semen quality in healthy men. Hum Reprod 18(2): 447-454.

436.Imura M, Itoh N, Takagi S, Sasao T, Takahashi A, Masumori N, Tsukamoto T. Balance of apoptosis and proliferation of germ cells related to spermatogenesis in aged men. J Androl. 2003 Mar-Apr;24(2):185-91.

437.Luetjens CM, Rolf C, Gassner P, Werny JE, Nieschlag E. 2002 Sperm aneuploidy rates in younger and older men. Hum Reprod 17:1826-1832.

438.Sloter ED, Marchetti F, Eskenazi B, Weldon RH, Nath J, Cabreros D, Wyrobek AJ. Frequency of human sperm carrying structural aberrations of chromosome 1 increases with advancing age. Fertil Steril. 2007;87:1077-86.

439.Wiener-Megnazi Z, Auslender R, Dirnfeld M. Advanced paternal age and reproductive outcome. Asian J Andrology 2012;14:69-79.

440.Fonseka KGL, Griffin DK. Is there a paternal age effect for aneuploidy. Cytogent Genome Res 2011;133:280-291.

441.Fisch H, Hyun G, Golden R, Hensle TW, Olsson CA, Liberson GL. 2003 The influence of paternal age on Down syndrome. J Urology 169:2275-2278.

442.Stene E, Stene J, Stengel-Rutkowski S. A reanalysis of the New York State prenatal diagnosis data on Down's syndrome and paternal age effects. Hum Genet. 1987;77:299-302.

443.Olshan AF, Schnitzer PG, Baird PA. 1994 Paternal age and the risk of congenital heart defects. Teratol 50:80-84.

444.Lian Z-H, Zack MM, Erickson JD. 1986 Paternal age and the occurrence of birth defects. Am J Hum Genet 39:648-660.

445.Brown AS, Schaefer CA, Wyatt RJ, Begg MD, Goetz R, Bresnahan MA, Harkavy-Friedman J, Gorman JM, Malaspina D, Susser ES. 2002 Paternal age and risk of schizophrenia in adult offspring. Am J Psychiatry 159:1528-1533.

446.Goriely A, McGrath JJ, Hultman CM, Wilkie AO, Malaspina D. "Selfish spermatogonial selection": a novel mechanism for the association between advanced paternal age and neurodevelopmental disorders. Am J Psychiatry. 2013 Jun 1;170(6):599-608.

447.Goriely A, Wilkie AO. Paternal age effect mutations and selfish spermatogonial selection: causes and consequences for human disease. Am J Hum Genet. 2012 Feb 10;90(2):175-200.

448.Lim J, Maher GJ, Turner GD, Dudka-Ruszkowska W, Taylor S, Rajpert-De Meyts E, Goriely A, Wilkie AO. Selfish spermatogonial selection: evidence from an immunohistochemical screen in testes of elderly men. PLoS One. 2012;7(8):e42382. doi: 10.1371/journal.pone.0042382. Epub 2012 Aug 6.

449.Wyrobek AJ, Eskenazi B, Young S, Arnheim N, Tiemann-Boege I, Jabs EW, Glaser RL, Pearson FS, Evenson D. Advancing age has differential effects on DNA damage, chromatin integrity, gene mutations, and aneuploidies in sperm. Proc Natl Acad Sci U S A. 2006;103:9601-6.

450.Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, Magnusson G, Gudjonsson SA, Sigurdsson A, Jonasdottir A, Jonasdottir A, Wong WS, Sigurdsson G, Walters GB, Steinberg S, Helgason H, Thorleifsson G, Gudbjartsson DF, Helgason A, Magnusson OT, Thorsteinsdottir U, Stefansson K. Rate of de novo mutations and the importance of father's age to disease risk. Nature. 2012 Aug 23;488(7412):471-5.

451.Frans EM, Sandin S, Reichenberg A, Långström N, Lichtenstein P, McGrath JJ, Hultman CM. Autism risk across generations: a population-based study of advancing grandpaternal and paternal age. JAMA Psychiatry. 2013 May;70(5):516-21

452.Harlap S, Paltiel O, Deutsch L, Knaanie A, Masalha S, Tiram E, Caplan LS, Malaspina D, Friedlander Y. 2002 Paternal age and preeclampsia. Epidemiol 13:660-667

453.Hellstrom WJ, Overstreet JW, Sikka SC, Denne J, Ahuja S, Hoover AM, Sides GD, Cordell WH, Harrison LM, Whitaker JS. Semen and sperm reference ranges for men 45 years of age and older. J Androl. 2006 May-Jun;27(3):421-8.

454.Nieschlag E, Lammers U, Freischem CW, Langer K, Wickings EJ. 1982 Reproductive functions in young fathers and grandfathers. J Clin Endocrinol Metab 55:676-681.

456.Gallardo E, Simon C, Levy M, Guanes PP, Remohi J, Pellicer A. 1996 Effect of age on sperm fertility potential: oocyte donation as a model. Fertil Steril 66:260-264.

457.Jung A, Schuppe H-C, Schill W-B. 2002 Compariso of semen quality in older and younger men attending an andrology clinic. Andrologia 34:116-122.Haidl G, Badura B, Hinsch KD, Ghyczy M, Gareiss J, Schill WB. 1993 Disturbances of sperm flagella due to failure of epididymal maturation and their possible relationship to phospholipids. Hum Reprod 8:1070-1073.

458.Handelsman DJ, Staraj S. Testicular size: the effects of aging, malnutrition, and illness. J Androl. 1985 May-Jun;6(3):144-51.

459.Handelsman DJ. Male reproductive aging: human fertility, androgens and hormone dependent disease. Novartis Foundation Symposium 2002;242:66-77.

460.Johnson L, Zane RS, Petty CS, Neaves WB. 1984 Quantification of the human Sertoli cell population: its distribution, relation to germ cell numbers, and age-related decline. Biol Reprod. Nov; 31(4):785-95.

461.Tulina N, Matunis E 2001 Control of stem cell self-renewal in Drosophila spermatogenesis by JAK-STAT signaling. Science 294:2546-9.

462.Black A, Lane MA. 2002 Nonhuman primate models of skeletal and reproductive aging. Gerontology 48:72-80.

463.Greunewald DA, Naai MA, Hes DL, Matsumoto AM. 1994 The Brown Norway rat as a model of male reproductive aging: evidence for both primary and secondary testicular failure.  J Gerontol 49:B42-B50.

464.Wang C, Leung A, Sinha-Hikim AP. 1993 Reproductive aging in the male brown-Norway rat: a model for the human. Endocrinology 133:2773-2781.

465.Taylor GT, Weiss J, Pitha J. 1988 Epididymal sperm profiles in young adult, middle-aged, and testosterone-supplemented old rats. Gamete Res 19:401-409.

466.Oakes CC, Smiraglia DJ, Plass C, Trasler JM, Robaire B. 2003 Aging results in hypermethylation of ribosomal DNA in sperm and liver of male rats. Proc Natl Acad Sci U S A 100:1775-1780.

467.Zirkin BR, Santulli R, Strandberg JD, Wright WW, Ewing LL. Testicular steroidogenesis in the ageing Brown Norway rat. J Androl 2003;14:118-123.

468.Wright WW, Fiore C, Zirkin BR. The effect of aging on the seminiferous epithelium of the brown Norway rat. J Androl. 1993 Mar-Apr;14(2):110-7.

469.La Magueresse B, Jegou B. In vitro effects of germ cells on the secretory activity of Sertoli cells recovered from rats of different ages. Endocrinology 1988;22:1672-1680.

470.Humphreys PN. Histology of the testes in aging and senile rats. Exp Gerontol 1977;12:27-34.

471.Syntin P, Robaire B. Sperm structural and motility changes during aging in the Brown Norway rat. J Androl 2001 22:235-244

472.Tramer F, Rocco F, Micali F, Sandri G, Panfili E. Antioxidant systems in rat epididymal spermatozoa. Biol Reprod 1998 59:753-758.

473.Edelmann P, Gallant J. 1977 On the translational error theory of aging. Proc Natl Acad Sci U S A 74:3396-3398

474.Marden JH, Rogina B, Montooth KL, Helfand SL. 2003 Conditional tradeoffs between aging and organismal performance of Indy long-lived mutant flies. PNAS 100:3369-3373.

475.Jones LS, Berndtson WE. 1986 A quantitative study of Sertoli cell and germ cell populations as related to sexual development and aging in the stallion.  Biol Reprod Aug;35(1):138-48

476.Tenover JS, McLachlan RI, Dahl KD, Burger HG, de Kretser DM, Bremner WJ.  Decreased serum inhibin levels in normal elderly men: evidence for a decline in Sertoli cell function with aging. J Clin Endocrinol Metab. 1988 Sep;67(3):455-9.

477.Bohring C, Krause W. Serum levels of inhibin B in men of different age groups. Aging Male. 2003 Jun;6(2):73-8.

478.Mahmoud AM, Goemaere S, De Bacquer D, Comhaire FH, Kaufman JM. Serum inhibin B levels in community-dwelling elderly men. Clin Endocrinol (Oxf). 2000 Aug;53(2):141-7.

479.Haji M, Tanaka S, Nishi Y, Yanase T, Takayanagi R, Hasegawa Y, Sasamoto S, Nawata H. Sertoli cell function declines earlier than Leydig cell function in aging Japanese men. Maturitas. 1994 Feb;18(2):143-53.

480.Kaler LW, Neaves WB. Attrition of the human Leydig cell population with advancing age. Anat Rec 1978;192(4):513-8.;

481.Neaves WB, Johnson L, Petty CS. Age-related change in numbers of other interstitial cells in testes of adult men: evidence bearing on the fate of Leydig cells lost with increasing age. Biol Reprod. 1985 Aug;33(1):259-69;

482.Paniagua R, Amat P, Nistal M, Martin A. Ultrastructure of Leydig cells in human ageing testes. J Anat. 1986 Jun;146:173-83.

483.Neaves WB, Johnson L, Porter JC, Parker CR Jr, Petty CS.  Leydig cell numbers, daily sperm production, and serum gonadotropin levels in aging men. J Clin Endocrinol Metab. 1984 Oct;59(4):756-63.

484.Nistal M, Santamaria L, Paniagua R, Regadera J, Codesal J. Multinucleate Leydig cells in normal human testes. Andrologia. 1986 May-Jun;18(3):268-72.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

LESSONS FROM RODENT MODELS OF OBESITY

LESSONS FROM RODENT MODELS OF OBESITY

Martin G. Myers, Jr., MD, PhD, Professor of Diabetes Research, Departments of Internal Medicine and Physiology, Director, Michigan Diabetes Research Center, University of Michigan
Rudolph L. Leibel, MD, Christopher J. Murphy Professor of Diabetes Research, Co-Director Naomi Berrie Diabetes Center, Russ Berrie Medical Science Pavilion, Columbia University

Updated: 6 January 2015.

INTRODUCTION

TAKE HOME POINTS:
1-
Rodent models in which monogenic alterations cause obesity in the absence of environmental changes have confirmed earlier inferences regarding the biologic/genetic control of energy balance in mammals.
2-Spontaneous models of obesity in rodents have identified fundamental molecular/cellular systems underlying the control of feeding and energy homeostasis, including the melanocortin system and CNS circuits that respond to leptin.
3-Engineered mutations in rodents have revealed additional genes and pathways participating in the control of body weight.
4-Most genes that impact body weight and adiposity affect the brain systems that control “regulatory” and/or hedonic aspects of feeding behavior. Effects on energy expenditure are frequently present, but of smaller magnitude.
5-Differences in adipocyte physiology between obese and lean individuals appear to be largely secondary phenomena
6-Most of the rodent monogenic obesities have human orthologs that result in comparably severe obesity-related phenotypes.  While these mutations are rare, they confirm that the molecular predicates for the control of body weight in humans are fundamentally the same as those in the rodents.
7-Engineered mutants in rodents have also permitted the analysis of putative genetic contributors to obesity in humans, as well as providing an initial blueprint of the molecular components, pathways and physical interconnections of these systems.

 

Historically, obesity has been considered a disorder of voluntary behaviors, exacerbated by the ready availability of food and reduced need for energy expenditure afforded by modern societies. Rodent models in which monogenic alterations provoke obesity in the absence of environmental changes have, however, conclusively demonstrated the biologic control of energy balance in mammals.  Indeed, orthologs of many of these genes cause or contribute to obesity in humans.  Spontaneous mono and polygenic models of obesity in rodents (along with obesity phenotypes of engineered mutations) have identified fundamental molecular/cellular systems underlying the control of feeding and energy homeostasis.  Importantly, most genes that impact body weight and adiposity affect the brain systems that control feeding.  Genetic studies in rodents have provided an initial blueprint of the molecular constituents and interconnection of these systems.

The first law of thermodynamics and body weight regulation.

The first law of thermodynamics (the conservation of energy) dictates that body energy stores reflect the difference between energy taken in and energy expended.  More intake with relatively less expenditure leads to energy storage (generally, in adipose tissue); conversely, when expenditure exceeds intake, energy/fat stores decline.  In this context, energy is generally taken in by eating (or drinking calorie-containing beverages). Behaviorally, two related systems govern eating (1): The circuits that control the incentive and reward values (wanting and liking) of food, and the satiety system, which promotes meal termination associated with the sensation of “fullness.”  Each of these systems is subject to long and short-term regulation. They are physiologically integrated, but for simplicity frequently studied and described as distinct entities.

Energy loss/expenditure includes energy consumed but not absorbed (and which therefore passes out of the body in the stool; this is ~5% under most normal circumstances); energy metabolized in the process of breaking down and storing nutrients (diet-induced thermogenesis); energy metabolized to maintain baseline cellular functions  at rest (resting metabolic rate, BMR); and energy consumed in  physical activity (non-resting energy expenditure; NREE) (2).  RMR accounts for about 70% of total energy expenditure (TEE) in sedentary adults, and is determined by body composition, age, wakefulness, and genetic factors. NREE is the next largest contributor, averaging approximately 20% of TEE. In sedentary individuals, low-level physical activities (fidgeting, short bouts of ambulation, etc.) make up most of NREE. Smaller amounts of energy (7%) are accounted for by diet-induced thermogenesis.  Recently, there has been growing interest in the contributions of the gut microbiome to systemic energy homeostasis by possible effects on the efficiency of nutrient utilization in the gut and the consequences of such bacterial metabolism for release of metabolites that could affect energy intake or expenditure (3).

Adipose tissue represents the major repository for ingested energy that exceeds immediate needs (2). The energy density of adipose tissue is nearly 10-fold greater than muscle (protein) or liver (glycogen). The ability to store such energy protects against environmental vicissitudes that might result in starvation, fetal wastage, and inability to provide sufficient breast milk to the young. Therefore, it is likely that evolution has promoted genes/alleles that favor energy storage and conservation. The existence of environments in which excess calories are readily available with minimum or no effort is a very recent occurrence in human evolution. Human genetic “makeup” is presumably designed for the opposite circumstance.  Contrary to earlier prevailing views, adipose tissue is not a passive energy depot, but participates in homeostatic processes that regulate food intake (e.g. production of leptin), and storage (e.g., insulin) and release (e.g., catecholamines) of the acylglycerides stored within them (4).

Spontaneous obesity in rodents provides the first clues to the genetic underpinnings of energy balance

In 1902, French geneticist L. Cuenot described the obese Yellow (Ay/a) mouse, which had been bred and maintained by European mouse fanciers since the 1800s (5). This was the first report of a spontaneously obese mouse, which prompted investigation of additional spontaneous obese mouse models, including by investigators at the Jackson Laboratories (Jax) in Bar Harbor, Maine.

The autosomal recessive obese (ob) mutation was discovered at Jax in 1949-50, after spontaneously arising in a non-inbred strain (6). Sixteen years later, a phenotypically similar mouse was identified (7). The diabetic state of these latter animals (studied on the diabetes-prone coisogenic KsJ background) distinguished them from ob (studied on the B6 background) and hence the mutation was designated diabetes (db).  In 1990, Coleman and colleagues described additional, milder recessive obese mutations in mice: tubby (tub), and fat (fat)(8,9).  Fewer obesity-related spontaneous mutations have been detected in rats, due in part to the absence of explicit screening of large numbers of progeny for such phenotypes, and the greater cost of rat husbandry. In addition to identifying mutations in genes identical to some of the murine genes above (e.g. leptin receptor mutations in db mice and the Zucker and Koletsky rats(10,11)), however, the OLETF obese rat has also been described (12).

Each of these mutant animals is hyperphagic compared to controls.  Furthermore, classic genetic studies revealed that each of these obese phenotypes had predictable, monogenetic heritability, demonstrating the genetic underpinnings of feeding, as well as overall energy balance.  The subsequent finding that some of these rodent obesity genes control body weight in humans confirms that biologic/genetic factors control feeding and the predisposition to obesity in humans, as well as in rodents (13).

Transgenic models

In addition to spontaneous models of obesity, genetic engineering (generally in mice) has provided many examples in which genetic alterations modulate body weight and adiposity.  In the early 1980s, genetic manipulation techniques  became available in rodents, enabling the analysis of systemic and organ-specific effects on physiology of one or more genes selected by the investigator (14).  Several genes important for energy balance that have been examined by such approaches are discussed below.  Historically, gene manipulation in mammals has been accomplished by one of two distinct means: standard transgenesis and gene targeting.

In standard transgenesis, artificial genes (often comprising promoter sequences designed to produce desired patterns of expression plus the coding sequences for the molecule to be expressed) are directly introduced into a fertilized oocyte, which is then implanted in a female surrogate to permit the development of the transgenic animal (15).  In this method, the site of the transgene insertion in the genome is random; hence, the insertion may inadvertently disrupt endogenous genes, and the expression pattern may be influenced by the site of insertion as well as by the promoter sequences used.  The use of very large genomic regions (such as those derived from bacterial artificial chromosomes (BACs)) to drive the gene of interest can mitigate some, but not all, of these expression issues (16).  Multiple independent transgenic lines must therefore be screened to identify correctly-expressing progeny.

Gene targeting is accomplished by using homologous recombination to introduce genetic sequences designed to modify specific genes, while leaving the rest of the genome intact (17).  Generally, manipulated DNA sequences are introduced into undifferentiated murine embryonic stem (ES) cells to recombine with native DNA, producing a modified ES cell line in which a specific gene is inactivated (a “knockout” or KO) or altered by editing or by the introduction of new genetic material (a “knockin” or KI) .  The modified ES cells are then injected into blastocysts, which are implanted into surrogate mothers.  The resultant pups generally contain cells derived from the ES cells along with cells donated by the recipient blastocyst (hence, these animals are termed “chimeras”).  The chimeras are then bred in an effort to obtain germline transmission of the ES cell-derived genes.  Thus, while gene targeting is quite specific in terms of the types and locations of manipulations, the use of ES cells requires a greater up-front investment of time and resources (generating the homologous targeting construct, screening ES cell clones for correct targeting, breeding for germline transmission, etc.) than does standard transgenesis.

New technologies have emerged over the past several years that promise to facilitate gene targeting.  These generally involve the use of site-specific nucleases (zinc finger nucleases, TALENs, CRISPRs, etc.) to create breaks in the genomic DNA of fertilized oocytes or ES cells (18,19).  These site-specific breaks can be employed not only to produce KO animals, but also to increase the efficiency of site-specific recombination within the oocyte, so that the co-injection of homologous templates can be used to generate KI animals without the use of ES cells.

In addition to the production of standard KO animals, and animals with edited coding or regulatory sequences in specific genes, gene targeting is also commonly used to produce conditional null or specific gene-expressing animals, often by employing the Cre recombinase/LoxP system (19).  This bacteriophage-derived system is composed of two components- Cre (a site-specific DNA recombinase) and LoxP (the short DNA sequences recognized by Cre).  In most versions of the system, Cre removes the DNA sequences that are flanked by LoxP sites (“floxed”).  By combining tissue-specific Cre expression with floxed genes of interest, the floxed genes may be disrupted in a tissue- and time-specific manner.  Cre-expressing animals may be generated by standard transgenesis (with the caveats, above, regarding site of integration), or can be delivered to specific sites in the genome by homologous targeting.  Animals carrying floxed alleles are produced by delivering LoxP sites to the desired locations in the genome by homologous targeting.

While many homologous targeting events are designed to be benign, they may have unintended consequences for the expression of the targeted gene (e.g., alterations in expression) or surrounding genes; these must be controlled for carefully.  Another important consideration in mouse models of obesity generated by gene targeting is genetic background effects. Targeting has been most consistently successful in the 129 mouse strain because the ES cells of 129 mice are relatively easy to culture and manipulate. However,  this strain exhibits increased levels of anxiety in response to environmental  stressors that could potentially distort food intake and metabolic phenotypes (20). ES cells from C57BL/6 require carefully controlled in vitro conditions, and even then, often fail to transmit the introduced mutation to the germ line. Serial backcrossing of a progenitor 129 or other strains, e.g., C57BL/6, can be used to transfer the mutation to a more suitable “background”. For example, mice overexpressing melanin-concentrating hormone (MCH) have an obese phenotype only when backcrossed for 7 generations onto the C57BL/6J background. On an FVB background, they appear to have a normal phenotype with regard to body composition (21).

PATHWAYS INITIALLY REVEALED BY SPONTANEOUSLY OCCURRING MUTATIONS

Obese and Diabetes reveal the endocrine control of energy balance.

The obese (ob; now Lepob) mouse was identified as an autosomal recessive mutation in a noninbred strain (Stock V) at Jackson Laboratory in 1949 (6). Mice segregating for the obese gene were backcrossed for many generations to generate a congenic line on the C57BL/6J background strain. Lepob mice on C57BL/6J, despite their early-onset obesity and transient glucose intolerance, are not diabetic; however, the Lepob mutation coisogenic on the diabetes-prone C57BL/KsJ line results in severe, early-onset type 2 diabetes (22).

Diabetes (db; now Leprdb) is a spontaneous recessive mutation that was first noted in a C57BL/KsJ mouse colony at the Jackson Laboratory. The KsJ Leprdb mutant is hyperphagic and obese, but also develops severe type 2 diabetes (7). Backcrossing Leprdb onto the C57BL/6J background attenuates the diabetic phenotype; C57BL/6J-Leprdb is virtually identical to C57BL/6J-Lepob.

Douglas Coleman, at Jackson Labs, looking for the molecular predicates of the lipostatic system posited by Kennedy (23) and Hervey (24), performed parabiosis (joined circulation) studies coupling Lepob mice to either wild-type or Leprdb mice (25). The Lepob mouse became lean when joined to a wild type, but, when joined to a Leprdb mouse, the Lepob mouse died of starvation. These findings led Coleman to hypothesize that a blood-borne factor regulating body weight might be deficient in Lepob, but circulating at high levels in the blood of Leprdb mice. He suggested that obese was the secreted factor and diabetes its receptor (25,26).

In 1994, the gene encoding Lepob was isolated by positional cloning (27) by a group at Rockefeller University, and its product, leptin, was shown to be produced primarily in adipocytes. Leptin is a type 1 cytokine, similar in structure to IL-6.  Lepob mice lack circulating leptin by virtue of an R105X mutation that creates a premature stop codon sequence in the leptin gene, resulting in a truncated protein that is rapidly degraded (27).

The gene (Lepr) that encodes the receptor for leptin (LepR) was identified by expression cloning (28); the first genetic mutation in Lepr was identified in the Leprdb mouse (10,29,30), thus confirming the conceptual model proposed by Coleman: the obesity of the Leprdb mouse was due to a mutation that precluded leptin signaling by the receptor that binds the ob gene product.


Alternative splicing of the Lepr transcript produces multiple isoforms of the receptor (which is a cytokine receptor similar to members of the IL6 receptor family): LepRa, -b, -c, -d, and so forth (Figure 1).  The mutation in the Leprdb mouse results from a splicing defect that causes the 3′ terminal exon (18a) of leptin receptor isoform a (Lepr-a) to be inserted into Lepr-b. A stop codon at the end of exon 18a prevents transcription of the Lepr-b terminal exon, so that LepRa is produced in place of LepRb (10,29,30). Because the Leprdb mouse synthesizes all leptin receptor isoforms except LepRb, it is clear that this isoform (which contains JAK box and STAT3 domains) is critical to the control of energy homeostasis (31).  Indeed, restoration of LepRb on a background null for all other LepR isoforms restores energy balance (32).

Figure 1. LepR isoforms and function. LepRa (Ra) represents the mostly highly expressed short form of LepR; LepRb (Rb) is the long form. Exon 17 contains half of a Jak docking site (BOX1) common to Ra, Rb and Rc, while exon 18b contains additional motifs required for full Jak2 binding (BOX2) and STAT3 signaling (31,33). Circulating leptin binding protein consists of extracellular domain that has been cleaved from the cell surface, along with the LepRe splice variant that lacks a transmembrane domain. Humans do not generate the splice variant, so that all LepRe is produced by cell surface cleavage, presumably by membrane associated metalloproteases (33).

LepRa, -c, -d and the other so-called “short” isoforms contain the same first 17 exons as LepRb, but diverge within the intracellular domain.  LepRb is the only isoform that mediates classical Jak-STAT signaling, as this isoform alone contains the motifs required to interact with Jak2 and to bind STAT proteins for downstream signaling (Figure 1)(34).  While the function of LepRb is clear, the functions of the short isoforms are not, although they have been speculated to function in leptin transport into the brain and/or a source of cleaved, circulating extracellular LepR (which, along with LepRe comprises the major circulating leptin-binding protein) (35). The biological role(s) of soluble LepR isoforms (sLEPR) are unclear. Human obesity and fasting are associated with decreased circulating sLEPR; pregnancy with increased sLEPR. sLEPR can block LEP transport across the BBB(36-40).

The physiologic function of leptin.

Disruption of Lep function results in hyperphagia and obesity, and leptin administration to Lepob mice (but not Leprdb animals), reduces food intake and adiposity, sparing lean tissue (41-43).  Thus, Lepob and Leprdb mice demonstrate that fat mass (along with both energy intake and expenditure) can be controlled by a single molecule.  Leptin represents a powerful biologic controller of feeding and energy balance, revealing the existence of an endocrine system that controls feeding and energy balance.  In humans, leptin deficiency also elicits a severe obesity phenotype: A rare, recessively inherited LEP mutation was discovered in two children who are members of a highly consanguineous Pakistani family (44). As with the Lepob mutation in mice, this frameshift mutation introduces a premature stop codon that truncates the leptin protein. While rare, additional leptin-deficient individuals (all of whom are severely obese) have been identified.  Daily subcutaneous administration of recombinant leptin dramatically and selectively reduces body fat to normal levels in these individuals (45).  A few humans homozygous for LEPR mutations have also been identified; these individuals present a severe obese phenotype similar to those lacking leptin, although – as anticipated -  they are not responsive to exogenous leptin (46).  In these patients, growth hormone deficiency and central hypothyroidism are phenotypes seen more frequently than in leptin deficiency per se.   It is important to note that mice (47) and humans (48) heterozygous for null mutations of either LEPR or LEP are more obese than suitable controls.  It is thus possible that individuals heterozygous for functionally null mutations of these and other genes encoding molecular components of the various signaling pathways regulating energy homeostasis discussed in this review constitute a significant proportion of the very obese.  Additionally, heterozygosity for several of these mutations would be expected to produce even greater levels of obesity.  The increasing use of exome sequencing in evaluating instances of severe obesity will lead to the detection of more instances of obesity caused by such oligogenic mechanisms.

While the role for leptin in the control of appetite and adiposity initially dominated the thinking about its biology, it rapidly became clear that leptin has other functions, and that the effects of high leptin are not as dramatic as those of low leptin.  Indeed, obese rodents and humans exhibit high circulating concentrations of leptin, commensurate with their high levels of leptin-producing adipose tissue (49,50).  Similarly, in contrast to the Lepob mice, increasing leptin to supraphysiologic levels in normal animals modestly and briefly blunts food intake and body weight [effect may be more striking than this]. Likewise, supraphysiological doses of leptin have only modest effects on body weight in obese and non-obese humans (51).  Thus, the absence of leptin appears to convey a more powerful signal than does its excess.  Also, Lepob mice (and their human counterparts) display additional phenotypes, including impaired growth and gonadal axis function, diminished immune function, infertility, and decreased energy expenditure due to low sympathetic nervous system tone and thyroid function- all of which are reversed by leptin treatment (52).  The lack of leptin also promotes increased hepatic glucose production, and leptin treatment suppresses hyperglycemia in models of several diabetes, including T1D (53).  This constellation of phenotypes resulting from low leptin mirrors the physiologic response to starvation; indeed, leptin treatment attenuates many of these consequences of very low adiposity (54).  Thus, normal leptin concentrations signal the repletion of energy (fat) stores to mitigate hunger and enable energy expenditure, while low leptin indicates the dearth of adipose reserves and promotes food-seeking and the conservation of remaining fat by reducing energy expenditure.  The concentration of leptin constituting such a signal of adequacy of fat stores may differ among individuals, reflecting genetic, developmental and acquired differences in the CNS molecules and circuits comprising this system (55).

Transgenic animals that lack adipose tissue exhibit a syndrome that mirrors that of lipodystrophic humans (who lack adipose tissue on a congenital or acquired basis): In spite of their leanness, lipodystrophic people and animals exhibit hyperphagia along with a predisposition to insulin resistance,  diabetes and other endocrine and metabolic abnormalities that are not corrected even with caloric restriction (56,57).  Due to their dearth of adipose tissue, leptin levels are low and leptin treatment improves their hunger and endocrine/metabolic abnormalities.  Indeed, leptin was recently approved for the treatment of lipodystrophy syndromes in humans (58).

A leptin-regulated neural network underlies energy balance.

The similar phenotypes of Lepob and Leprdb mice (along with the inability of leptin to alter physiology in Leprdb mice) indicates that leptin action on LepRb-expressing cells must mediate its effects.  Consistent with its behavioral effects (e.g., on feeding) and its effects on the neuroendocrine and autonomic systems, most LepRb-expressing cells lie in the brain.  Indeed, transgenic overexpression of LepRb throughout the central nervous system (CNS) partially corrects the obesity syndrome of Leprdb-3J mice (which lack all LepR isoforms) (32). Similarly, ablation of CNS LepRb using a neuron-specific Cre in combination with a floxed (Leprflox) allele promotes hyperphagia, neuroendocrine failure, and obesity (59).  Some tissues outside of the CNS express LepRb, but the physiologic role for leptin action on these non-CNS cells remains unclear.

Within the brain, the majority of LepRb-expressing neurons are found within the hypothalamus and brainstem, consistent with the known roles for these structures in the control of feeding and endocrine and autonomic function (60-62).  While LepRb ablation in the nucleus tractus solitarius (NTS) in the brainstem reduces satiety (consistent with the known role of this brain structure), pan-hypothalamic ablation of LepRb promotes a phenotype very similar in quality and magnitude to that of whole-body null Leprdb animals (63).  Furthermore, ablation of LepRb from broadly-distributed hypothalamic vGat- or Nos1-expressing neurons promotes dramatic hyperphagia and obesity (64,65).  Smaller, more circumscribed sets of hypothalamic LepRb neurons have also been implicated, as well. Within the arcuate nucleus (ARC), an important satiety center in the brain, excision of Leprflox by Pomccre and Agrpcre modestly increases feeding and adiposity (66,67).  Ablation of LepRb in the SF1-expressing ventromedial hypothalamic nucleus (VMH) blunts the increase in energy expenditure that accompanies increased adiposity, and deletion of SF1 in the lateral hypothalamic area (LHA, which is associated with motivation) diminishes motor activity and promotes obesity (68,69).  LepRb neurons in the ventral premammillary nucleus (PMv) play roles in reproduction (70).  Importantly, many additional groups of LepRb cells in the hypothalamus (especially the ARC and dorsomedial hypothalamic nucleus (DMH)) and brainstem have as yet undetermined functions.

Spontaneously-arising Agouti mice reveal the crucial role for the hypothalamic melanocortin system in energy balance.

Expression of the agouti gene (a) normally occurs intermittently in the hair follicle resulting in the production of alternate yellow and black pigment bands of the resulting hair; this admixture produces the agouti coat color (71). The molecule acts as primarily as an inverse agonist at the melanocortin receptor (MC1R in skin).

The Yellow mutation of the agouti locus (Ay/a) is also termed ‘lethal yellow’, since homozygotes for the allele are prenatal lethal. Yellow was bred by mouse fanciers in Europe beginning in the 1800s, and was notable for the dominant inheritance of its striking yellow coat color and obesity proportional to the intensity of the yellow coat (5). In 1960, another spontaneous agouti mutation was detected in the Jackson Laboratory colony; viable yellow (Avy) (72). The original, lethal, yellow mutation is a deletion of the Raly gene, which causes a fusion of the constitutively active Raly promoter to the agouti gene, resulting in ectopic continuous overexpression of agouti in all somatic (including brain) cells.  Avy/a is also the result of ectopic overexpression of agouti; this mutation results from insertion of a retrovirus-like repetitive intracisternal A particle (IAP) into a noncoding exon of agouti. The resulting splice variant fuses the constitutively active Raly promoter to the agouti gene, allowing constitutive overexpression of agouti in all somatic (including brain) cells.

The increased body weight of Ay/a and Avy/a mice results mainly from hyperphagia, and reflects both increased fat mass and lean body mass (with increased body length) (73). By contrast, Lepob and Leprdb mice have a selective expansion of fat mass due to increased food intake, decreased energy expenditure, preferential storage of excess calories as fat, decreased body length and lean body mass (74). Thus, agouti overexpression affects food intake similarly to leptin, but alters energy expenditure less dramatically.

The agouti gene encodes agouti signaling protein (ASP), a peptide with a high affinity for melanocortin receptors. The yellow coat color of the Ay/a mouse results from continuous overexpression of agouti in the skin which blocks (mainly by inverse agonist effects) alpha-melanocyte-stimulating hormone (-MSH) signaling at melanocortin-1 receptors (MC1R) in the hair follicle (71,75). Since -MSH activates melanocytes to initiate synthesis of eumelanin (black pigment) instead of phaeomelanin (yellow pigment), antagonism of -MSH by ASP elicits a yellow coat color. ICV administration of -MSH and -MSH agonists decreases food intake and body weight; overexpression of agouti in the Ay/a brain antagonizes the anorectic action of -MSH signaling as well as blunting the endogenous activity of the receptor, thus causing hyperphagia.

Melanocortin receptors- the role for MC4R in energy balance.
The brain contains two predominant melanocortin receptor isoforms- melanocortin receptor-3 and -4 (MC3R and MC4R, respectively) (76).  Both isoforms are potently activated by -MSH. MC4R is expressed in the PVN, DMH, VMH and LHA(75), all of which are hypothalamic sites crucial for the control of food intake.  Mice homozygous for a targeted deletion of Mc4r display substantial hyperphagia, 3- to 5-fold increased adiposity, and 50-100% increased body weight compared to littermate controls, while maintaining the same absolute lean body mass as +/+ littermates (77). Heterozygosity for the Mc4r null mutation elicits an intermediate phenotype.

Mc4r-/- mice also have increased linear growth, as is characteristic of Ay/a mice (77). Mc4r-deficient mice maintain core body temperature when exposed to a cold challenge (78), suggesting that sympathetic tone is not reduced to the same extent as it is in Lepob and Leprdb mice.  Oxygen consumption of Mc4r-/-mice is reduced by 20% as compared to weight-matched controls, however, indicating that MC4R mediates some control of energy expenditure, in addition to affecting feeding.

Approximately 4% of morbid human obesity (BMI > 40 kg/m2) is due to mutations in MC4R (79-81). Preserved lean mass and increased stature are also evident in the human MC4R deficiency syndrome, as in rodent models (82).  Most obesity associated with MC4R mutations has been attributed to heterozygosity for such mutations (83). Severe childhood obesity results from a null MC4R receptor, generated by missense, frame shift, deletion, and nonsense mutations (82). MC4R mutations are codominantly inherited, and heterozygous family members are overweight, suggesting that these mutations impair the function of the normal gene product, unlike null mutations in Mc4r-/- mice.  Genome-wide association studies (GWAS) have revealed common non-coding polymorphisms within MC4R that are associated with increased adiposity (84). These are likely variants affecting the transcription rate of the gene.

Agouti-related peptide (AgRP) blocks MC4R signaling.
 An homology search to identify a protein with a normal physiological function comparable to ASP in the brain revealed a candidate with 25% amino acid homology to ASP (Agouti-related peptide; AgRP) (71).  Eutopic Agrp expression is restricted to a set of neurons in the ARC that contain the orexigenic neuropeptide Y (NPY), and which are activated by fasting or leptin deficiency, consistent with a role in controlling (promoting) food intake (85).  In vitro binding studies showed that AgRP binds the MC1, MC3 and MC4 receptors, and mice globally overexpressing AgRP (like Agouti mice) are hyperphagic and obese compared to nontransgenic littermates (86).  AgRP differs from ASP in that it does not block eumelanin synthesis to elicit a yellow coat color when transgenically overexpressed in mice.  MC4R exhibits significant constitutive activity in vitro; ASP and AgRP not only block binding of -MSH to the MC4R, but also suppress MC4R constitutive signaling, i.e., act as inverse agonists (71).  In support of a physiological role for the orexigenic action of AgRP, fasted animals show increased expression of ARC Agrp mRNA, as do Lep ob and Lepr db mice (86).  ICV administration of AGRP to rats elicits a long-lasting hyperphagic response (71).

Interestingly, however, mice congenitally  null for Agrp or Npy exhibit minimal alterations in energy balance, as do compound Npy-/-;Agrp-/- mice (87). This lack of phenotype apparently reflects developmental compensation/reprogramming, however, since mice in which AgRP neurons are ablated early in development exhibit normal energy balance, while ablation of these cells in adults results in aphagia and death by starvation (88,89).  Thus, AgRP likely plays an important physiologic role in the promotion of feeding, presumably by blocking melanocortin receptor action.

MC3R also contributes to the control of adiposity.
Centrally administered AgRP causes hyperphagia in Mc4r-/- mice (90), supporting a role for an additional brain melanocortin receptor in the regulation of body weight. The MC3R is expressed in the ARC, DMH and VMH (75). In comparison to the MC4R, the MC3R has reduced affinity for AGRP, and increased affinity for -MSH. Mc3r-/- mice develop late onset obesity accompanied by a 2-fold increase in fat mass (91,92). These effects are considerably smaller than the 3- to 5-fold increased adiposity observed the Mc4r null mouse.  The Mc3r-/- mouse displays normal food intake, but reduced locomotor activity and increased respiratory quotient (reduced fatty acid oxidation) on high-fat chow as compared to contols, suggesting alterations in energy partitioning when challenged with a high-fat diet. Thus, while MC3R does not impact feeding to the same extent as MC4R, it nonetheless plays a role in the control of energy expenditure/metabolism, and nutrient partitioning, and thus plays a role in the control of adiposity.  GWAS have not identified SNPs in the region of MC3R as risk alleles for increased body weight, however.

Proopiomelanocortin (POMC).
POMC, the precursor peptide for melanocortin receptor agonists and endorphins with effects on ingestive behaviors, is expressed in both the anterior pituitary and the hypothalamus (76). In the anterior pituitary, POMC is processed to ACTH and -lipotropin. In the intermediate lobe of the pituitary, and in the hypothalamus, ACTH is processed further to -MSH and CLIP.  Pomc-/- mice generated by gene targeting weigh twice as much as wild-type littermates at 12 weeks of age, and are hyperphagic when presented with either standard or high-fat chow (93). Daily intraperitoneal injection of -MSH to Pomc-/- mice caused a 46% weight loss over a 2-week period with a concomitant darkening of coat color. To investigate specifically the role of -MSH in body weight regulation, Pomc null mice have been rescued by transgenic overexpression of POMC in the pituitary but not the brain (94). Homozygous pituitary rescue mice are 33% heavier than Pomc null mice, indicating that -MSH deficiency in the brain causes the obesity phenotype and that circulating glucocorticoids restored with pituitary Pomc replacement have an additive effect to increase body weight.

Functionally consequential mutations in POMC have been identified in humans. Human subjects have been described who are 1) compound heterozygous for mutations in exon 2 of POMC that result in premature termination of transcription as well as a frameshift mutation that disrupts the common binding site of -MSH and ACTH, or 2) homozygous for a nucleotide transversion mutation in exon 3 that truncates POMC protein at codon 79, resulting in trace or undetectable amounts of circulating -MSH and ACTH (95). These individuals exhibit early onset obesity and red hair because of the -MSH deficiency, and are adrenal insufficient due to a lack of circulating ACTH.  In a number of instances the accompanying adrenal insufficiency has led to death in infancy. Hence, suspicion of this diagnosis should be considered to constitute an urgent medical issue.

Syndecans.
Cell surface heparan sulfate proteoglycans (HSPGs) modulate ligand-receptor interactions at neural synapses (96). In vitro studies suggest that HSPG syndecan-1 may bind to AgRP, facilitating AgRP binding to MC4R. In accord with this model, transgenic mice overexpressing syndecan-1 exhibit late-onset obesity. The endogenous hypothalamic analogue of syndecan-1 is syndecan-3, and fasted mice show a four-fold induction of syndecan-3 mRNA in hypothalamic areas involved in energy balance. When challenged by a 16-hour fast, syndecan-3-/- mice exhibit blunted refeeding, presumably due to the decreased binding of AgRP to MC4R.

Fat reveals roles for peptide processing systems in the control of energy balance

Fat.
The fat mutation was first identified at the Jackson Laboratory in a colony of inbred HRS/J mice (9). Cpe fat mice exhibit apparent hyperinsulinemia as early as four weeks of age followed by obesity at 8 to 12 weeks. Approximately 77% of the measured insulin is proinsulin, thus bioactive insulin levels are normal.  By a positional candidate gene approach, a TàC point mutation in Cpefat (which results in a S202P transversion) was identified in carboxypeptidase E (CPE) (97). CPE is an enzyme that cleaves COOH-terminal dibasic residues arginine and lysine in prohormone precursors of insulin, enkephalin, POMC, NPY, melanin concentrating hormone (MCH), cholecystokinin (CCK), oxytocin (OXT), and vasopressin (AVP). Transgenic overexpression of insulin in Cpefat mice does not correct the obesity, suggesting that aberrant processing of one of the other peptides targets of CPE contributes to their increased fat mass (98).  POMC is cleaved by prohormone convertase 1 (PC1, also called PCSK1 or PC1/3) and PC2 (PCSK2) to generate a precursor peptide that is further cleaved by CPE to generate active MSH; miscleavage of POMC may account for the obesity in Cpefat mice.

One missense polymorphism in human CPE has been identified: a CàT transversion (99). This mutation results in a non-conservative R283W amino acid substitution that greatly decreases CPE enzymatic activity and is associated with early onset type 2 diabetes.

Prohormone convertases.
Mice null for PC1 are obese, hypoadrenal, and hypogonadal, presumably due to impaired processing of POMC and GnRH (100).  Proinsulin, prothyrotropin releasing hormone, progastrin, proneurotensin and prodynorphin are also incompletely processed in these animals.  Human PC1 deficiency caused by missense and splice site mutations in the PC1 gene also results in a disorder characterized by obesity and hypocortisolemia as well as hypogonadism (101).

Animals null for PC2 are not obese, although they exhibit phenotypes consistent with other defects in peptide processing (102); this lack of obesity may result from the partial activity of the POMC PC1 product on MC3/4R even in the absence of PC2, combined with other hormonal and neural changes in these animals.  No humans defective in PC2 function have been identified to this point.

Prolyl carboxypeptidase.
PRCP is a serine protease that cleaves the COOH-terminal amino acid from substrate proteins where the penultimate amino acid is a proline residue (substrate preferences are X-P-F-COOH and X-P-V-COOH) (103).  In general, PRCP inactivates biologically active peptides.  The COOH-terminal sequence of the 13 amino acid α-MSH molecule is PV; removal of V abrogates the ability of α-MSH to decrease food intake. Prcp-null mice are lean with increased sensitivity to exogenous α-MSH.  Presumably, mutations in PRCP also alter the inactivation of other peptides.

OTHER SPONTANEOUSLY-OCCURRING RODENT MUTATIONS LEADING TO OBESITY

Tubby.
 The tubby mutation arose spontaneously at Jackson Laboratory in the C57BL/6 strain (9). These mice have a mild, late-onset obesity apparent by 8 to 12 weeks of age that is associated with hyperinsulinemia without hyperglycemia. Tubby results from a GàT transversion that interferes with normal intron excision (104). The result is an aberrant transcript in which a 44-base pair deletion at the 3′ end of the gene is replaced with a 24-base pair intronic segment that is usually spliced out. Tub-/- mice are phenotypically indistinguishable from tubby, suggesting that the phenodeviant that occurred in the Jackson Laboratory colony had a loss-of-function mutation in the tubby gene. Lack of detectable tubby protein or transcript in tubby mice further supported this finding.

The precise physiological mechanism(s) underlying the obesity of tubby mice remain unknown (104), but the tubby mutation also produces retinal and cochlear degeneration, which is seen in primary ciliopathies such as the Bardet-Biedl and Alstrom syndromes.  (See below).  The tubby gene product, TUB, binds to membrane phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) and is released upon PtdIns(4,5)P2 hydrolysis.  A variety of data suggest an important role for TUB in GPCR signaling and trafficking, as well as insulin and leptin signaling, via the primary cilium (105).

Mahogany and mahoganoid.
 The spontaneous, autosomal recessive coat color mutations, mahogany (mg) and mahoganoid (md), were first reported over forty years ago (71). When crossed to Ay/a mice, both mg and md darken coat color and attenuate obesity. Positional cloning of mg identified a gene orthologous to the human immune-response protein attractin (atrn), the gene product of which accumulates on the surface of activated T cells and subsequently facilitates the interaction between T cells and antigen by “attracting” macrophages (106).  Atrn encodes a single transmembrane protein with a glycosaminoglycan side chain that has been suggested to chaperone agouti to melanocortin receptors. Atrn is expressed widely in the brain, as well as the skin, heart, kidney, liver and lung of wild-type mice. The Atrnmg mutation in mice is due to a ~5-kb retroviral insertion in intron 11 that disrupts Atrn expression, and which permits a relative increase in eumelanin expression in the hair follicle, resulting in dark fur, presumably as a consequence of increased melanocortin signaling at the melanocortin-1 receptor (MC1R).  Atrnmg mutants also have 10-15% reduction in body weight and have 20-40% less body fat content than littermate controls.  Deficiency of Atrn in Ay/a mice reduces body weight and adiposity by increasing energy expenditure rather than reducing food intake, suggesting melanocortin-independent mechanisms of action.  Indeed, Atrn mg mice may have neurological deficits (107), and the zitter mutation, which causes hypomyleination in rats, results from an 8-bp deletion in Atrn at a splice donor site that decreases Atrn expression.  Thus, Atrn function may not be entirely Ay/melanocortin-dependent.

The mahoganoid locus, Mgrn1 (mahogunin; RING finger 1), is located  2 cM from the centromere of chromosome 16 (108,109). Five mutations at this locus have been identified: md, md3J, md4J, md5J, and md6J. The Mgrn1md mutation is a 5-kB IAP element intronic insertion between exons 11 and 12, which attenuates expression. Similar to the mg phenotype, the Mgrn1md mouse is lean, and the allele reduces body weight and darkens the yellow coat color of Ay/a mice. The Mgrn gene product contains a RING finger domain consistent with ubiquitin E3 ligase function. Although the spontaneous mutations at the Mgrn1 locus do not cause neurological degeneration, the Mgrn1md-nc mutation generated by caffeine mutagenesis results in histopathological changes similar to those seen in Atrnmg and the zitter rat (110).  Both mahoganoid and mahogany convey their effects on ASP signaling by effects on endosomal trafficking of MC4R (111).

THE USE OF TRANSGENIC MODELS TO STUDY SYSTEMS INVOLVED IN ENERGY BALANCE

Hypothalamic circuits important for energy balance

AgRP/NPY neurons, their mediators and modulators

The AgRP-expressing neurons of the ARC also contain NPY, as well as the fast inhibitory neurotransmitter, GABA (112).  These neurons are inhibited by leptin and activated by fasting and leptin deficiency; their activation promotes feeding and decreases energy expenditure, while their ablation results in lethal anorexia (88,89).

Mediators of AgRP/NPY neuron function.
As noted above, Agrp and Npy proteins have been ablated individually and in combination, with little effect upon energy balance in wild-type animals, although their ablation modestly attenuates the obesity of Lepob/obanimals (87).  In contrast, blockade of GABA release from these neurons, via the cre-mediated deletion of the vesicular GABA transporter (vGat) results in leanness and interferes with the response to ghrelin or food restriction, suggesting that these neurons (and especially GABA release therefrom) is crucial for promoting food intake, especially in response to signals of negative energy balance (112).  Detailed studies of animals ablated for AgRP neurons have also suggested that GABA release from AgRP cells into the brainstem parabrachial nucleus is especially important for the stimulation of feeding by AgRP neurons (113).

NPY receptors.
NPY receptors are G-protein coupled receptors; six NPY receptors have been identified: Y1, Y2, Y3, Y4, Y5, and Y6 (114).  Y1 and Y5 are localized to the hypothalamus and ICV administration of Y1- and Y5 receptor antagonists reduce food intake. Mice with targeted Y1 disruption show a variable and sex-dependent alterations in energy balance [139]; however, Y5-deficient mice develop mild obesity (115).  Indeed, Lepob/ob;Y5-/- mice are  not different than Lepob/ob mice in terms of energy balance. In addition, both Y1- and Y5-deficient mice are hyperphagic in response to centrally administered NPY, suggesting the existence of an additional NPY receptor or receptors that regulate food intake.

Hypothalamus-specific deletion of the Y2 receptor by viral delivery of Cre recombinase in Y2Flox mice results in a significant decrease in food intake and body weight(116).  The endogenous peptide YY1-36, a Y2 ligand co-localized with GLP-1 in the L-type endocrine cells of the GI mucosa, stimulates food intake; however, its cleavage product peptide YY3-36 (PYY3-36), a Y2 agonist primarily secreted from endocrine cells lining the gastrointestinal tract, decreases food intake (117). Thus, the Y2 receptor may have dual functionality that is determined by the PYY moiety that binds to it.

 

Ghrelin, GHSR, and GOAT.
Ghrelin is a hormone released from cells in the epithelium of the stomach, duodenum, ileum, cecum and colon (118); its pharmacologic administration promotes dramatic feeding (119).  The receptor for ghrelin is the growth hormone secretagogue receptor (GHSR), and ghrelin acylation is required for GHSR activation.  Ghrelin is acylated (octanoylated) by ghrelin O-acyl transferase (GOAT) in the cells that synthesize it (120).  Diurnal release of ghrelin into the circulation coincides with the initiation of meals, and decreases over the course of each meal (121); ingested fatty acids are required for ghrelin acylation, so that active ghrelin only increases prior to meals in animals that have fed over the prior 24 hours.  GHSR is highly expressed on AgRP/NPY cells (as well as some other cells) in the hypothalamus, and ghrelin activates AgRP/NPY cells.  Ghrelin is also expressed in the epsilon cells of the islets of Langerhans where it may act as a brake on glucose-induced insulin release by direct effects on the beta cell and/or antagonism of GLP1 secretogogues (122).

Consistent with the modest baseline phenotypes of mice null for the individual neurotransmitters employed by AgRP/NPY neurons, mice null for ghrelin, GHSR, or GOAT exhibit no detectable alterations in baseline energy balance, and only modest defects in refeeding (123).  It is possible that some of the lack of effect of Npy, Agrp, Ghrelin, Ghsr, or Goat deletion reflects developmental reprogramming that occurs with defects in AgRP/NPY neurons during circuit formation, however, since ablation of these neurons early in development produces little effect on body weight, while their ablation in adults results in  lethal anorexia (88,89).

Serotonin (5HT) receptor 2c.

The 5HT2cR is expressed in the ARC, PVN, LHA, and anterior hypothalamic nucleus (AH) of the hypothalamus (124). Agonists of the 5HT2cR promote weight loss, and several are in clinical trials or approved for the treatment of obesity.  Deletion of 5ht2cr produces hyperphagic obesity that is accentuated by high fat diet.  A subset of ARC POMC neurons express 5ht2cr, and the Pomccre-mediated reactivation of a null 5ht2cr allele in these cells attenuates the food intake and obesity in the 5ht2cr null mice (125).  The effect of 5HT2cR activation may vary by nucleus, but, in aggregate, 5ht2cr mutant mice confirm the important role for this receptor in energy balance.

Single-minded-1 (SIM1) and the PVH.

In mice, Sim1 encodes a transcription factor required for the development of the PVH, an integrative hypothalamic nucleus in which -MSH, NPY and 5-HT are released (126). Many PVH neurons contain MC4R, Y1R and/or 5-HT2cR.  Ablation of the PVH in rodents produces a profound hyperphagia (127).  The PVH contains a diverse constellation of neuronal subtypes, including those that express oxytocin (OXT), corticotropin releasing hormone (CRH), vasopressin (AVP), thyroid hormone releasing hormone (TRH), and others.  Many of these molecules are thought to participate in energy balance, as well as their well-recognized neuroendocrine functions.

Human Single-minded-1 (SIM1) deficiency was discovered by karyotyping in three case studies of young obese patients with small deletions or translocations at the human SIM1 locus on chromosome 6 (126). Homozygous deletion of Sim1 is embryonic lethal in mice. Sim1+/- mice are normal until 4 weeks of age, when they develop hyperphagic obesity (128).  These mice display reduced numbers of neuronal nuclei in the PVH with a proportional decrease in overall size of the PVH. Presumably, the decreased number of PVH neurons in these mice diminishes anorexic “tone” from the PVH, leading to hyperphagia and obesity in the Sim1+/- mice, as well as in rare human patients with SIM1 mutations.  Additionally, deletion of Mc4r with Sim1cre recapitulates the hyperphagia and obesity of Mc4r-/- mice, as does the ablation of Mc4r in the PVH by virus-mediated cre delivery (129,130).  Thus, PVH SIM1-expressing neurons represent crucial direct targets for MC4R action and for energy balance.  Understanding the roles for the various subsets of PVH SIM1 neurons in the control of energy balance has been more difficult, however.  Ablation of Mc4r from OXT, AVP, and CRH neurons does not alter energy balance (130).

Oxytocin.
A variety of pharmacologic data suggest important roles for PVH-derived OXT in the control of feeding; the injection of OXT into the region of the NTS promotes satiation (131).  However, genetic data argue against an important role of OXT or OXT neurons in energy balance.  Not only do Oxt-/- animals display no alteration in feeding or energy balance, but neither the activation nor the ablation of PVH OXT neurons in adult animals alters food intake (132,133).

Corticotropin releasing hormone.
CRH increases glucocorticoid secretion via the hypothalamic-pituitary-adrenal axis, but also acts on a number of CNS circuits. Centrally administered CRH produces decreased food intake and weight loss (134); conversely, elevated CRH promotes activation of the HPA axis and promotes Cushing’s syndrome with increased central adiposity due to peripheral glucocorticoid excess.  While inactivation of Crh causes glucocorticoid deficiency, it has no impact on energy homeostasis (135).  Similarly, while antagonism of the receptors for CHR (CRH1 and CRH2) leads to increased food intake, decreased energy expenditure and increased body weight (136), CRH receptor-deficient mice display normal regulation of body weight (137).  Thus, while PVH CRH neurons and CRH signaling are crucial for the control of the HPA axis and for stress responses, CRH and its receptors do not appear to play an important role in the control of energy balance by the PVH.

Vasopressin (AVP).
In addition to magnocellular AVP neurons (mainly located in the SON) that project to the posterior pituitary to control fluid balance, PVH AVP cells project widely throughout the brain.  While the deletion of Mc4r from these cells does not alter energy balance, the pharmacogenetic activation of these cells modestly suppresses food intake, suggesting that these cells may play some role in the control of energy balance, even though they do not represent direct targets of melanocortin action (130,138).

Steroidogenic factor-1 (SF1) and the VMH.

Steroidogenic factor 1 (Sf1; Nr5a1) is a transcriptional modulator expressed in the dorsomedial portion of the VMH- a hypothalamic nucleus implicated in the regulation of body weight (139). The VMH contains neurons that express LepRb, MC3R and other receptors involved in body weight regulation. Although Sf1-deficient mice were first described in 1994, early death due to adrenal insufficiency prevented characterization of this mouse in adulthood. By performing adrenal transplantation, it was possible to observe late-onset obesity in Sf1 -deficient mice, which resembles the mild obesity phenotype of MC3R deficiency. Sf1-deficient, adrenal-transplanted mice appear normal until 8 weeks of age, when their body weights diverge from their wild-type littermates. By 6 months, Sf1-/- mice are 72% heavier than controls due primarily to increased body fat, with no differences observed in linear growth. The obesity of these animals appears to result largely from decreased energy expenditure.  Sf1cre has been used to delete LepRb from the VMH; this manipulation decreases energy expenditure and accentuates obesity in high-fat diet-fed animals (68).  Many SF1-containing VMH neurons also contain the neuropeptide PACAP (the product of the Adcyap gene), which may contribute to the control of energy expenditure (140).  Thus, Sf1-mediated manipulation of the dorsomedial VMH has revealed a crucial role for this region in the control of energy expenditure and thus overall energy balance.

Reward circuitry:
The lateral hypothalamic area (LHA) and mesolimbic dopamine (DA) system.

While the ARC, PVH, and (to a lesser extent) VMH mediate net anorexic tone, the LHA (together with the mesolimbic DA system) modulates behavioral incentive- including the drive to eat (1).  While lesions of the ARC or LHA promote hyperphagia and obesity, destruction of the LHA abolishes the motivation to feed, resulting in starvation.  While many details of these reward circuits remain to be discovered, LHA neurons modulate the mesolimbic DA system by projections to the ventral tegmental area (VTA; where the DA cell bodies lie) and the striatum (a crucial target of VTA DA neurons).

The VTA and DA.
Mice that lack tyrosine hydroxylase (TH) cannot make the precursor for catecholamines and are deficient in DA, noradrenaline and adrenaline; these animals die between embryonic day 11.5 and 15.5; restoring TH in noradrenergic neurons generates viable mice that synthesize noradrenaline and adrenaline normally, but do not synthesize DA in neurons of the mesolimbic DA system (141). DA-deficient pups nurse normally until 2 weeks of age, but thereafter fail to thrive due an inability to wean themselves onto solid food unless supplemented with the DA precursor, L-DOPA, suggesting that DA is required for normal ingestive behavior (as well as activity). However, ingestive behavior data that implicate dopamine as a stimulator of food intake may be confounded by the roles of dopamine in the initiation of motor activity and reward mechanisms.

The LHA.
Both leptin and the melanocortins have been implicated in the control of two important sets of neurons that lie within the LHA.  One population contains the neuropeptide melanin concentrating hormone (MCH; not related to POMC or any of its derivative peptides) (142).  MCH promotes feeding, and animals null for MCH (or its receptor) are lean (143).  The MCH receptor is located on the primary cilium, and some of the effects of ciliopathies on adiposity may be conveyed by effects on this receptor (see discussion of ciliopathies below).   A distinct set of LHA neurons express the neuropeptide hypocretin (HCRT; also known as orexin) (144,145).  Based upon early acute pharmacologic studies, HCRT was originally conceived of as an orexigen; subsequent work has revealed animals null for HCRT or its receptors to be mildly obese, however (146).  Indeed, narcolepsy, which results from the loss of HCRT action in mice and humans, is associated with increased adiposity (147).  Most of the effect of HCRT administration or Hcrt mutation on energy balance results from decreased physical activity and energy expenditure.  Similarly, the LepRb-containing neurons that control HCRT neurons have been identified- these contain neurotensin (NT) and lie in the LHA, intermingled with the HCRT cells (148-150).  Ablation of LepRb from these LHA cells prevents the normal regulation of HCRT neurons and results in decreased motor activity and energy expenditure.  Both LHA LepRb neurons and HCRT cells project to the VTA, and parameters of DA neuron function are altered in mice lacking LepRb in NT neurons, as well as in Leprob/oband Leprdb/dbmice.

Genes involved in insulin and leptin signaling.

Transcription factors involved in leptin signaling.
LepRb, like other Type 1 cytokine receptors, activates signal transducers and activators of transcription (STATs) as a major component of its signaling pathway (31).  During leptin signaling, tyrosine phosphorylated residues on LepRb recruit STAT3 and STAT5, which are then phosphorylated by Jak2 to promote their trafficking to the nucleus.  In the nucleus, STATs bind DNA and modulate gene expression.  STAT3 mediates the majority of leptin action, since disruption of the binding site for STAT3 on LepRb causes a severe obesity phenotype in mice that is similar to the obesity syndrome of Leprdb/db mice (151).  Similarly, disruption of Stat3 in the forebrain or in LepRb-expressing POMC, or AgRP neurons results in obesity in mice (152,153).  While the brain-wide disruption of the genes encoding both isoforms of STAT5 (STAT5a and STAT5b) causes mild late-onset obesity, the deletion of Stat5a/b specifically in LepRb neurons produces no detectable phenotype, suggesting that STAT5 signaling is not required for leptin action in vivo (154-156).  STAT5 represents a major mediator of GM-CSF signaling, however, and mice null for GM-CSFR in the brain animals are obese, suggesting that the role for STAT5 in energy balance may be linked to the action of GM-CSF or other cytokines different than leptin (155).

Insulin receptor.
Like leptin, insulin circulates in proportion to fat mass, and alters neuropeptide expression in the hypothalamus via receptors located in the ARC, PVN, and DMH (157).  ICV insulin has been reported to decrease food intake in rats and mice.  Furthermore, mice deleted for insulin receptor (Insr) throughout the CNS display a modest late-onset obesity (more prominent in females), and are more susceptible to diet-induced obesity than wild-type mice (158).  Thus, CNS INSR signaling plays a role in energy balance.  Deletion of Insr in skeletal muscle causes modest increases in adiposity, presumably by decreasing insulin-stimulated glycogen storage in muscle and concomitantly increasing glucose uptake in adipose tissue (hence, due to changes in nutrient partitioning) (159).  Conversely, deletion of Insr from adipose tissue produces lipodystrophy (160).

The IRS-protein/PI 3-kinase pathway.
The tyrosine phosphorylation of insulin receptor substrate proteins (IRS-proteins; IRS-1, -2, -3, and -4) represents the first downstream step in insulin signaling (161).  Tyrosine phosphorylated IRS-proteins engage downstream molecules, such as those in the phosphatidylinositol 3-kinase (PI3-kinase) pathway, to mediate insulin action.  While deletion of Irs1 interferes primarily with peripheral insulin action and the growth axis, deletion of Irs2 affects pancreatic beta cells and the brain to cause insulin deficient diabetes (due to islet failure) and obesity.  Restoration of Irs2 in the islets of Irs2-/- mice or brain-specific ablation of Irs2 results in normoglycemic obesity, consistent with a role for brain IRS2 signaling in energy balance (162).  Indeed, deletion of Irs2 from LepRb-expressing neurons promotes obesity, albeit a milder obesity than observed in animals deleted for Irs2 throughout the brain.  While leptin modulates the IRS-proteinàPI3-kinase pathway, deletion of Irs2 itself does not interfere with leptin action, suggesting that IRS2 may primarily play a role in brain insulin action (163).  Deletion of Irs4, which is expressed in neurons of the hypothalamus, modestly alters energy balance.  A variety of subunits and downstream effectors of the PI3-kinase signaling pathway have also been deleted in several neuronal populations in mice (35).  These produce phenotypes generally consistent with the notion that PI3-kinase is important for the proper function of the POMC and AgRP neurons that modulate energy balance- at least in part by controlling the firing of these important neurons.  Similarly, ablation of the gene encoding the PI3-kinase inhibited transcription factor, FOXO1, tends to augment insulin and leptin action in vivo (164).

mTOR and autophagy.
The mechanistic target of rapamycin complex 1 (mTORC1) is activated by PI3-kinase signaling and nutrient (especially amino acid) availability to promote cellular anabolic processes while blunting autophagy (165).  ICV amino acids activate hypothalamic mTOR and promote satiation, while blockade of hypothalamic mTORC1 using the inhibitor, rapamycin, promotes hyperphagia- suggesting a role for mTORC1 in producing satiety (166).  The role for mTORC1 in the hypothalamic control of energy balance may be complicated, however, as neuronal firing also activates mTORC1, and mTORC1 is increased in AgRP/NPY neurons during fasting (167).  Furthermore, lifelong activation of mTORC1 or inactivation of autophagy (via deletion of Atg7) in POMC neurons promotes hyperphagic obesity in mice (168,169).

Tyrosine phosphatases and other inhibitors of insulin and leptin signaling.
Protein tyrosine phosphatase-1B (PTP1B) dephosphorylates cognate tyrosine kinases (including those associated with INSR and LepRb) to terminate signaling (170,171).  In addition to exhibiting increased insulin sensitivity, Ptp1b-/- mice are lean compared to controls and are resistant to weight gain on a high-fat diet, suggesting increased leptin action in these animals.  Indeed, animals in which Ptp1b is disrupted throughout the brain, or specifically in LepRb or POMC neurons demonstrate increased leanness and enhanced leptin action (172,173).  Other phosphatases may also limit insulin and/or leptin signaling:  Mice null for Rptpe or Tcptp also demonstrate leanness and increased leptin sensitivity (174).

Suppressors of Cytokine Signaling (SOCS proteins), including SOCS1 and SOCS3, bind to activated cytokine receptor/Jak2 kinase complexes (including the LepRb/Jak2 complex) to mediate their inhibition and degradation (175).  SOCS proteins may also inhibit INSR and other related tyrosine kinases.  Leptin signaling via STAT3 promotes Socs3 expression in hypothalamic LepRb neurons; SOCS3 protein binds to phosphorylated Tyr985 of LepRb to attenuate LepRb signaling (176).  The physiologic importance of this pathway is demonstrated by the leanness of mice containing a substitution mutation of LepRb Tyr985  (binding site for SOCS3 on LEPR; see figure above) and the similar phenotype of mice lacking Socs3 in the brain or in LepRb neurons (177,178).  While LepRb Tyr985 also mediates the recruitment of the tyrosine phosphatase SHP2 (aka, PTPN1), data from cultured cells suggest that SHP2 mediates ERK pathway signaling by LepRb, and disruption of Ptpn1 in the brain, in LepRb neurons, or in POMC neurons, promotes obesity (31).

SH2B1.
SH2B1 binds to activated Jak2, as well as to the INSR, TrkB, and a few other receptor tyrosine kinase complexes to increase their activity and mediate aspects of downstream signaling (179).  Sh2b1-/- mice display a complex phenotype that includes obesity; brain-specific absence of Sh2b1 also promotes obesity in mice (180,181).  Thus, SH2B1 signaling in the brain is required for energy balance, perhaps due to its requirement for correct signaling by multiple receptors involved in energy homeostasis.  Furthermore, the phenotype of several human patients with morbid obesity, developmental delay, and behavioral disorders are associated with chromosomal deletions (16p11.2)  or coding variants involving SH2B1 (182).  Indeed, GWAS studies have suggested a role for common variants in SH2B1 in human obesity (84).

Other transcription factors.
The transcription factor BSX is found in AgRP neurons, where its expression is regulated by leptin and feeding status (183).  Deletion of Bsx in mice reduces the increase in Npy and Agrp expression, and associated hyperphagia, in food-deprived mice, suggesting a role for BSX in the function of AgRP neurons and in feeding control.  While its role in leptin action is not known, the disruption of Atf3 (which encodes a STAT3-responsive member of the AP1 family of transcription factors) in a poorly-characterized set of hypothalamic neurons also causes obesity in mice (184).

The peroxisome-proliferator activated receptor (PPAR) family of nuclear transcription factors modulates genes encoding proteins involved in lipid homeostasis (185). PPAR induces hepatic genes that promote mitochondrial uptake and beta-oxidation of free fatty acids, and PPAR agonists (fibrates) are used clinically to lower circulating triglycerides and free fatty acids. Ppara-/- mice exhibit mild late-onset obesity that may partially result from decreased energy expenditure, but food intake is increased in these animals, and no increase in feed efficiency is observed, suggesting that increased food intake may ultimately drive this phenotype (186).  PPAR is expressed primarily in adipose tissue, and promotes adipocyte differentiation and storage of triacylglycerols in adipose depots (185).  PPAR agonists (thiazolidinediones, TZDs) increase insulin sensitivity and have been used in the treatment of type 2 diabetes.  PPAR agonists promote weight gain and, although Pparg-/- is embryonic lethal), Pparg+/- mice weigh 14% less than wild-type C57BL/6 mice, and have a 70% reduction in WAT mass. These mice display an elevated metabolic rate, and also a decrease in food intake.  Indeed, while the primary effect of PPAR manipulation of body weight and adiposity was previously assumed to results from direct adipose tissue action, genetic and pharmacologic manipulation of PPAR in the brain has revealed that brain PPAR action promotes increased feeding, which accounts for the energy balance effects of PPAR(187-189).  A common polymorphism of PPARG has been associated with BMI in GWAS studies (84).

Mouse Models of human obesity syndromes.

Brain-derived neurotrophic factor (BDNF)/TrkB signaling.
BDNF, a member of the neurotropin family, is widely expressed in the nervous system during development, as well as being expressed within several brain regions important for energy homeostasis in adults (190).  It acts via its receptor, TrkB, to control a variety of basic neural processes, including proliferation, survival, and plasticity.  Given its many important roles in the CNS, alteration in BDNF expression (or that of its receptor, TrkB) would be predicted to interfere with multiple processes.  Indeed, humans haploinsufficient for BDNF display impaired cognitive function and hyperactivity, in addition to hyperphagic obesity (191,192).  Mutations in TrkB produce similar hyperphagia and obesity in rare human patients, along with impaired cognitive function and nociception (193).  Interestingly, a coding polymorphism in BDNF (Val66Met) is associated both with obesity and with binge eating disorders in humans (194), consistent with the role for BDNF/TrkB signaling in energy balance, and suggesting a broader role for this system in the genetic determination of adiposity in humans.  Indeed, alteration of TrkB and/or BDNF function in the hypothalamus of mice promotes obesity (195,196).  Furthermore, polymorphisms in BDNF are associated with risk for obesity in human GWAS studies (84).

Ciliopathies.
A subset of mutations causing defects in primary cilia promote obesity syndromes (197,198).  The primary cilium is found on most cells; while structurally related to motile cilia (such as flagella), the primary cilium is immotile and does not participate in propulsion.  The primary cilium plays a crucial sensory role in cells, including cell-specific sensing, such as olfaction in sensory epithelium, photoreception in retinal cells, mechanical transduction in kidney cells, and signaling via a variety of cell surface receptors, including many GPCRs.  A broad group of disease-causing human mutations have now been recognized to result from mutations in genes affecting ciliary functions (the “ciliopathies”).  The clinical presentation of these diseases variably includes anosmia, retinal degeneration, kidney malformations, and a variety of developmental and neural defects, many of which are idiosyncratic to the particular gene that is mutated.  A number of these mutations produce obesity in addition to the other phenotypes noted above, both in mice and humans.  Included in these obesity-causing ciliopathies are Bardet-Biedel Syndrome (BBS), McKusic-Kaufman Syndrome, Alström Syndrome, and, possibly,  Joubert Syndrome.

Structural defects are apparent in the primary cilia of humans with BBS and the mice segregating for mutations in these genes (199,200).  The structural changes may not themselves account for the functional derangements associated with these mutations.  The BBS proteins, which constitute a “BBS-some” complex associated with the base of the primary cilium/basal body, participate in the trafficking of proteins to and within the cilium.  Indeed, mutations affecting IFT88, a protein specific for trafficking within the cilium, in mice results in an obesity phenotype similar to that produced by mutations in BBS genes (201).  While the particular protein(s) whose impaired trafficking may underlie this obesity is not yet clear, the primary cilium is crucial for signaling via a variety of receptor signaling pathways, including the WNT and SHH pathways, tyrosine kinases (such as the receptor for PDGF), MCHR, and numerous GPCRs.  There is also evidence for impaired leptin receptor signaling in mice segregating for Bbs (202)– though attributed by some to the consequences of weight gain per se (203)ref] - and Rpgrip1l mutations (204).   Such alterations could impair the development or function of a variety of neural circuits important for the regulation of energy balance.  The deletion of Ift88 from POMC neurons produces a portion of the obesity phenotype observed in the complete null, suggesting roles for multiple cell types (and perhaps multiple signaling pathways) in the complete ciliopathy phenotype (201).

FTO.
The human locus with the strongest GWAS linkage to adiposity (a polymorphism located in the human FTO locus) also contributes the largest amount to the genetic component of polygenic human obesity (205).  Multiple mechanisms for the regulation of energy balance have been proposed for this alteration.  In mice, Fto is expressed in the brain, including in hypothalamic feeding centers, where its expression is modulated by leptin and feeding status (206).  Furthermore, its role in controlling food intake and body weight is suggested by the lean phenotype of Fto-/- mice, although these animals present a complex phenotype that includes runting (207).  In contrast, mice ubiquitously overexpressing Fto or overexpressing Fto in the brain demonstrate increased food intake and adiposity (208).  Alternatively, the Fto locus is adjacent to the Rpgrip1l gene, which encodes a protein involved in primary cilium function; the non-coding sequence variant in intron 1 of Fto are physically associated (in linkage disequilibrium) with alleles of a transcription factor (Cux1) binding site that, by binding in the intron affects expression of a ciliary gene, Rpgrip1l (204). Hence, the effects of the FTO alleles may be conveyed via effects on RPGRIP1L.  Recently, it has been suggested that these Fto-associated polymorphisms lie within a long-range enhancer element that modulates the expression of a downstream gene, Irx3 (209).  The Fto polymorphism predicts the expression of hypothalamic Irx3, not Fto, in mice.  Mice null for Irx3 or that overexpress a dominant negative Irx3 mutant in the hypothalamus demonstrate increased leanness. It is possible, of course, that the Fto intronic variants are affecting the expression of multiple genes other than Fto itself. In fact, the strength of the association is consistent with that possibility.

Prader-Willi syndrome (PWS)
PWS presents in infancy with low birth weight, hypotonia and poor feeding, with a progressive transition to hyperphagia and obesity starting after age 2 or 3 years.  Additional features include short stature (correctible with growth hormone therapy), central hypogonadism, characteristic behaviors (especially around feeding), and often cognitive impairment (210,211).  Most instances result from a 5-7 Mb deletion of an imprinted region (PWS region) on the paternal chromosome 15 (15q11-q13) and are non-recurrent.  Within this deletion lie a number of genetic elements, including the genes encoding MAGEL2 and NECDIN, which are thought to be involved in neural development and function, and a complex non-coding locus.  Non protein -coding genes in this interval  include a transcribed non-coding gene (SNURF-SNRPN) that encodes a multitude of C/D box small nucleolar (sno-) RNA genes, including SNORD116.  The RNA products of these SNORD genes are thought to be involved in RNA editing, perhaps of specific mRNA species.  A small number of individuals with PWS phenotypes associated with microdeletions of the implicated region on chromosome 15 have reduced the number of candidate genes for this syndrome (210).  Some of these patients have demonstrated obesity and developmental delay in the absence of many of the other features of PWS. These lesions primarily affect SNORD loci, prominently including SNORD116, suggesting an important role for this SNORD in the obesity of PWS.  The Snord116 locus has been deleted from mouse models, which display a growth defect and behavioral abnormalities, including a relative hyperphagia that develops after weaning, but which is balanced by increased motor activity (212).  Thus, the effects of SNORD116 likely contribute to PWS, but may not account for all of the phenotypes.  The functions of Necdin and Magel2 have also been examined in genetically targeted mouse models- Magel2-/- mice display early growth retardation with late-onset obesity, and Necdin-/- mice display early postnatal respiratory failure along with a subset of PWS-associated behaviors (213-215).  Thus, the full PWS likely results from the combined effects of multiple genes; several genes within the PWS region also likely contribute to the maximal obesity phenotype.  It is not yet clear how each of the loci within the PWS alter neurophysiology and/or which neurons they might specifically affect energy balance.  As with BBS, some of these genes are likely to be affecting brain structural development/connectivity as well as more conventional signaling pathways.   Understanding the molelcular physiology of PWS (and BBS) is likely to identify novel genes in the control of energy homeostasis in non-syndromic obesities.

MODELS THAT PROBE ROLES FOR SATIETY SYSTEMS IN ENERGY BALANCE.

A variety of gut-derived signals including peptides (such as cholecystokinin (CCK), glucagon-like peptide-1 (GLP1), and amylin (a.k.a., islet amyloid polypeptide)) and vagal signals converge on hindbrain circuits in the NTS to promote satiation and meal termination (216).  These systems are crucial for the short-term control of feeding, and their pharmacologic manipulation may be therapeutically important.  Injection of CCK, GLP-1, or amylin, including into the hindbrain, promotes meal termination.  Furthermore, supraphysiologic/pharmacologic agonism of GLP1 and amylin receptors not only induces satiation, but promotes modest weight loss.  It is not clear that these systems modulate long-term feeding under physiologic conditions, however.

CCK.
CCK is a gastrointestinal hormone secreted in response to ingestion of a meal by enteroendocrine “I” cells located primarily in the duodenum. Many of these cells co-express other peptides affecting ingestive behavior (ghrelin, GIP, PYY). CCK induces a transitory sensation of satiety, secretion of pancreatic enzymes and gallbladder contraction. CCK-A receptors are located on vagal afferents of the stomach and the liver and transduce signals via the vagal nerve to satiety centers in the brainstem, eliciting a brief reduction in food intake (for a review, see (217)). CCK-B receptors are located diffusely throughout the brain, but their role in the satiety effect of CCK has not been demonstrated. The Otsuka Long-Evans Tokushima Fatty (OLETF) rat is an outbred strain of Long-Evans rats used experimentally as a model of type 2 diabetes. This animal has a 34% increase in food intake resulting from larger meal size, accompanied by a 23% increase in body weight at 15 weeks as compared to lean Long-Evans rats (218). In 1994, a mutation in the CCK-A receptor (CCKAR) of OLETF rats was identified; this 6847-base-pair deletion disrupts the Cckar promoter, reducing receptor expression.  While CCK decreases meal size and duration, compensatory increases in meal frequency prevent CCK from producing long term effects on total food intake or body weight.  Indeed, deletion of Cckar in mice does not cause obesity, suggesting that the OLETF phenotype results from a number of genetic variants that act in concert with the Cckar mutation.

GLP1.
 GLP-1 functions as an incretin (stimulator of insulin secretion) following its release from L-cells of the duodenum after nutrients enter the intestine (219).  GLP1 can also modulate satiety: ICV GLP-1 (or GLP1R agonists) potently suppress food intake in rats and mice, while the GLP-1 receptor antagonist, exendin (9-37), increases short-term food intake. Some of the effects of GLP1 on food intake may be due to delayed gastric emptying.  Glp1r-/- mice exhibit decreased circulating insulin concentrations during a glucose tolerance test, suggesting that GLP1R is important for glucose-stimulated insulin secretion (GLP-1 acting as an “incretin”).  Body weight and food intake are unaffected by ablation of GLP-1R, however, suggesting that (like CCK and CCKAR) this system primarily modulates short-term satiation, rather than long-term energy balance, under normal physiologic circumstances.

INTERACTIONS OF THE IMMUNE SYSTEM AND ENERGY BALANCE

Inflammatory signals are proposed to mediate several distinct metabolic responses.  Clearly, strong acute inflammatory stimuli (including those associated with systemic infection, cancer, etc.) decrease appetite and increase energy expenditure, promoting cachexia.  Conversely, obesity is associated with increased low-grade inflammation that appears limited to particular tissues, such as adipose tissue and the hypothalamus.  This low-grade “metabolic inflammation” is associated with insulin resistance and obesity.  A variety of animal models have been employed to explore the interaction of inflammatory signals and energy balance/metabolism.

Systemic immune signaling promotes negative energy balance.
Lipopolysaccharide (LPS) administration, which produces some of the metabolic consequences of bacterial infection, blunts appetite; the mechanism of this hypophagia overlaps with the systems that control energy balance, as the LPS-induced anorexia requires the melanocortin system (220).  Consistent with the induction of negative energy balance by systemic inflammation, alterations that blunt inflammation generally blunt inflammatory anorexia.  While not altering baseline energy balance in chow-fed animals, deletion of IL-1 converting enzyme (ICE; which is essential for IL-1 activity), prevents LPS-induced anorexia in mice (221).  The inflammatory system may also contribute to the control of energy balance under normal physiology, as well: adiposity is increased in Il6-/- and Gmcsf-/- mice, and in mice with impaired macrophage function due to the targeted deletion of Mac-1 or LFA-1 (or their receptor, ICAM-1) (222).  Conversely, mice with constitutively increased IL-1 receptor signaling induced by targeted deletion of the endogenous IL-1 receptor antagonist, Il1ra, display reduced body mass compared to wild-type littermates (223).

Metabolic inflammation.
Obesity is associated with increased production of a number of cytokines (including TNF) in adipose tissue, resulting primarily from the activation of adipose tissue macrophages and other immune cells (224,225).  Indeed, a number of manipulations that decrease adipose tissue inflammation ameliorate the metabolic dysfunction associated with obesity.  While interference with generalized macrophage function may increase adiposity, as noted above, other manipulations that alter their pro-inflammatory (versus anti-inflammatory) nature increase leanness and improve metabolic function (226,227).  Similarly, interfering with the Nf-kb pathway (which is crucial for the response to a variety of inflammatory stimuli) in the liver improves metabolic function in obese mice.  Some data also suggest a contributory role of hypothalamic inflammation, including gliosis, in promoting obesity.  However, debate continues regarding whether this inflammation provokes or attenuates obesity, virus-mediated interference with Nf-kb signaling in the hypothalamus ameliorates obesity and metabolic dysfunction (228).  The ER stress in adipose tissue and the hypothalamus, potentially a consequence of metabolic inflammation, is also associated with obesity (229).  Genetic or pharmacologic interference with ER stress ameliorates obesity and insulin resistance in rodent models.

ENERGY EXPENDITURE AS A DETERMINANT OF ADIPOSITY

With few exceptions, most of the systems that dramatically alter energy balance act primarily via the control of feeding; isolated alterations in energy expenditure promote more modest changes in energy balance that may be synergistic with effects on ingestive behavior, and may be detected under specific environmental and experimental conditions.  Increases in energy expenditure and negative energy balance promote a compensatory increase in feeding.  Similarly, decreased energy expenditure will cause the accretion of adipose mass, which tends to restrain feeding.  For instance, interference with normal VMH function (discussed above) decreases diet-induced energy expenditure, and promotes increased adiposity only when animals are provided high caloric density diets. The adipokine leptin, which is responsive to acute and chronic changes in adipose tissue energy stores plays an important role in promoting these reciprocal responses.

However, the physiological responses to reductions in energy stores – increased drive to eat, reduced energy expenditure – are much stronger than the response to increased energy stores (55).

Animal models with altered energy expenditure.
Uncoupling protein 1 (UCP1, which is found primarily in brown and beige adipose tissue (BAT)) allows dissipation of the electrochemical gradient across the inner mitochondrial membrane, releasing energy as heat (230).  Ablation of BAT in mice expressing diphtheria toxin A driven from the UCP1 promoter or congenital deletion of Ucp1 fails to alter adiposity at thermoneutrality, although adiposity increases slightly relative to controls in animals raised at temperatures colder than thermoneutrality, since these animals fail to substantially increase energy expenditure in response to the cold challenge (231).  Similarly, the phenotype of mice null for the 3-AR was not as severe as predicted: fat mass in male mice is only slightly increased, even in animals consuming a high-energy diet under non-thermoneutral conditions (232). Also, “-less” mice, with a global targeted deletion of all three b-adrenergic receptor isoforms, have only slightly increased body fat (22.2 % ± 0.9 as compared to 16.2% ± 1.9 for wild type controls) on high fat diet under non-thermoneutral (232).

Increased sympathoadrenal activity in adipose tissue activates a signaling cascade that induces phosphorylation of regulatory subunits of protein kinase A (PKA), which in turn inhibits lipogenesis and increases lipolysis. Deletion of the regulatory subunit II b of PKA (RII b ) , found mainly in WAT, BAT and brain, causes a compensatory increase in RI a , a subunit isoform that constitutively upregulates PKA activity (233). RII b -/- mice therefore have constitutively increased cAMP in response to sympathetic activation in adipose depots, with secondary elevation of metabolic rate and body temperature, and a 50% reduction in WAT pad weight despite a compensatory hyperphagia.

ALTERATIONS IN ADIPOSE TISSUE THAT AFFECT ENERGY BALANCE

Glucocorticoids and adipocyte 11-β–hydroxysteroid dehydrogenase type 1 (11βHSD-1).
 11βHSD-1 is the enzyme that catalyzes the conversion of cortisone to biologically active cortisol. Visceral obesity is associated with elevated cortisol secretion due to increased local activity of 11βHSD-β1 in adipose tissue, but normal levels of circulating glucocorticoids (234). Transgenic overexpression of 11βHSD-1 driven by the aP2 promoter (to confer adipose tissue specificity) produced mice that consumed 17.1% more calories than lean controls and thus gained 16% more weight by 9 weeks of age, and weighed 21% more calories than controls when administered a high-fat diet. A 3.7-fold increase in the mesenteric (visceral) fat pad weight was detected (235).  These findings suggest that increased production of glucocorticoids in adipose tissue drives the visceral obesity syndrome, although the increased food intake in these animals implicates potential non-local mechanisms.

CONCLUSIONS

The mice and rats described in this essay provide proof that body weight and composition are regulated by specific genes that participate in complex neural and metabolic pathways that determine energy intake and expenditure. The identification of molecules responsible for the single gene obesities in these animals has expedited the discovery of many other molecules, pathways and developmental processes that constitute the still only incompletely understood mechanisms for energy homeostasis that interact with developmental and environmental processes to determine body mass and composition. Their existence provides definitive refutation of vitalist/psychological notions that have permeated the field of energy intake and metabolism, and provides the heuristic, reductionist framework in which ongoing research on these questions should be conducted. It is likely that major genes and their modifiers, as well as allelic variants of a larger number of genes with lesser individual impact, will eventually account for both qualitative and quantitative aspects of the critical phenotypes in rodents and humans. As this chapter demonstrates, mice and rats provide a powerful resource for the discovery and study of the constituent molecules, and for hypothesis generation regarding the same processes in humans. The ability to refine the characterization of the behavioral and metabolic phenotypes that are controlled by these genes –in rodents and humans– will add greatly to the power of genetics to reduce the complex continuous phenotypes that are the physiologic “stuff” of energy homeostasis to their constituent molecular events.

 

 

References

1.         Berthoud HR. Interactions between the "cognitive" and "metabolic" brain in the control of food intake. Physiol Behav. 2007.

2.         Leibel RL. Energy in, energy out, and the effects of obesity-related genes. The New England journal of medicine. 2008;359(24):2603-2604.

3.         Ridaura VK, Faith JJ, Rey FE, Cheng J, Duncan AE, Kau AL, Griffin NW, Lombard V, Henrissat B, Bain JR, Muehlbauer MJ, Ilkayeva O, Semenkovich CF, Funai K, Hayashi DK, Lyle BJ, Martini MC, Ursell LK, Clemente JC, Van Treuren W, Walters WA, Knight R, Newgard CB, Heath AC, Gordon JI. Gut microbiota from twins discordant for obesity modulate metabolism in mice. Science. 2013;341(6150):1241214.

4.         Ahima RS, Flier JS. Adipose tissue as an endocrine organ. Trends in endocrinology and metabolism: TEM. 2000;11(8):327-332.

5.         cuenot l. [Les races pures et leur combinaisons chez les souris]. Arch Zool Exper Gener. 1905;3:123-132.

6.         Ingalls AM, Dickie MM, Snell GD. Obese, a new mutation in the house mouse. The Journal of heredity. 1950;41(12):317-318.

7.         Hummel KP, Dickie MM, Coleman DL. Diabetes, a new mutation in the mouse. Science. 1966;153(3740):1127-1128.

8.         Coleman DL. Diabetes-obesity syndromes in mice. Diabetes. 1982;31(Suppl 1 Pt 2):1-6.

9.         Coleman DL, Eicher EM. Fat (fat) and tubby (tub): two autosomal recessive mutations causing obesity syndromes in the mouse. The Journal of heredity. 1990;81(6):424-427.

10.       Chua SC, Jr., Chung WK, Wu-Peng XS, Zhang Y, Liu SM, Tartaglia LA, Leibel RL. Phenotypes of mouse  diabetes  and rat  fatty  due to mutations in the OB (Leptin) receptor. Science. 1996;271:994-996.

11.       Phillips MS, Liu Q, Hammond HA, Dugan V, Hey PJ, Caskey CJ, Hess JF. Leptin receptor missense mutation in the fatty Zucker rat [letter]. Nature genetics. 1996;13(1):18-19.

12.       Kawano K, Hirashima T, Mori S, Saitoh Y, Kurosumi M, Natori T. Spontaneous long-term hyperglycemic rat with diabetic complications. Otsuka Long-Evans Tokushima Fatty (OLETF) strain. Diabetes. 1992;41(11):1422-1428.

13.       Ramachandrappa S, Farooqi IS. Genetic approaches to understanding human obesity. The Journal of clinical investigation. 2011;121(6):2080-2086.

14.       Palmiter RD, Brinster RL, Hammer RE, Trumbauer ME, Rosenfeld MG, Birnberg NC, Evans RM. Dramatic growth of mice that develop from eggs microinjected with metallothionein-growth hormone fusion genes. Nature. 1982;300(5893):611-615.

15.       Brinster RL, Chen HY, Trumbauer M, Senear AW, Warren R, Palmiter RD. Somatic expression of herpes thymidine kinase in mice following injection of a fusion gene into eggs. Cell. 1981;27(1 Pt 2):223-231.

16.       Yang XW, Gong S. An overview on the generation of BAC transgenic mice for neuroscience research. Current protocols in neuroscience / editorial board, Jacqueline N Crawley  [et al]. 2005;Chapter 5:Unit 5 20.

17.       Capecchi MR. Altering the genome by homologous recombination. Science. 1989;244:1288-1292.

18.       Hsu PD, Lander ES, Zhang F. Development and applications of CRISPR-Cas9 for genome engineering. Cell. 2014;157(6):1262-1278.

19.       Rossant J, Nagy A. Genome engineering: the new mouse genetics. Nature medicine. 1995;1(6):592-594.

20.       Dockstader CL, Rubinstein M, Grandy DK, Low MJ, van der Kooy D. The D2 receptor is critical in mediating opiate motivation only in opiate-dependent and withdrawn mice. The European journal of neuroscience. 2001;13(5):995-1001.

21.       Ludwig DS, Tritos NA, Mastaitis JW, Kulkarni R, Kokkotou E, Elmquist J, Lowell B, Flier JS, Maratos-Flier E. Melanin-concentrating hormone overexpression in transgenic mice leads to obesity and insulin resistance. J ClinInvest. 2001;107(3):379-386.

22.       Naggert JK, Mu JL, Frankel W, Bailey DW, Paigen B. Genomic analysis of the C57BL/Ks mouse strain. Mammalian genome : official journal of the International Mammalian Genome Society. 1995;6(2):131-133.

23.       Kennedy GC. The role of depot fat in the hypothalamic control of food intake in rats. ProcRSoc. 1953;140:578-592.

24.       Hervey GR. A hypothetical mechanism for the regulation of food intake in relation to energy balance. The Proceedings of the Nutrition Society. 1969;28(2):54A-55A.

25.       Coleman DL. Effects of parabiosis of obese with diabetic and normal mice. Diabetologia. 1973;4:294-298.

26.       Coleman DL, Hummel KP. The effects of hypothalamics lesions in genetically diabetic mice. Diabetologia. 1970;6(3):263-267.

27.       Zhang Y, Proenca R, Maffei M, Barone M, Leopold L, Friedman JM. Positional cloning of the mouse obese gene and its human homologue. Nature. 1994;372(6505):425-432.

28.       Tartaglia LA, Dembski M, Weng X, Deng N, Culpepper J, Devos R, Richards GJ, Campfield LA, Clark FT, Deeds J, Muir C, Sanker S, Moriarty A, Moore KJ, Smutko JS, Mays GG, Woolf EA, Monroe CA, Tepper RI. Identification and expression cloning of a leptin receptor, OB-R. Cell. 1995;83(7):1263-1271.

29.       Chen H, Charlat O, Tartaglia LA, Woolf EA, Weng X, Ellis SJ, Lakey ND, Culpepper J, Moore KJ, Breitbart RE, Duyk GM, Tepper RI, Morgenstern JP. Evidence that the diabetes gene encodes the leptin receptor: Identification of a mutatioan in the leptin receptor gene in db/db mice. Cell. 1996;84(3):491-495.

30.       Lee GH, Proenca R, Montez JM, Carroll KM, Darvishzadeh JG, Lee JI, Friedman JM. Abnormal splicing of the leptin receptor in  diabetic  mice. Nature. 1996;379:632-635.

31.       Robertson SA, Leinninger GM, Myers MG, Jr. Molecular and neural mediators of leptin action. Physiology & behavior. 2008;94(5):637-642.

32.       Kowalski TJ, Liu SM, Leibel RL, Chua SC, Jr. Transgenic complementation of leptin-receptor deficiency. I. Rescue of the obesity/diabetes phenotype of LEPR-null mice expressing a LEPR-B transgene. Diabetes. 2001;50(2):425-435.

33.       Chua SC, Jr., Koutras IK, Han L, Liu SM, Kay J, Young SJ, Chung WK, Leibel RL. Fine structure of the murine leptin receptor gene: Splice site suppression is required to form two alternatively spliced transcripts. Genomics. 1997;45:264-270.

34.       Kloek C, Haq AK, Dunn SL, Lavery HJ, Banks AS, Myers MG, Jr. Regulation of Jak kinases by intracellular leptin receptor sequences. J Biol Chem. 2002;277(44):41547-41555.

35.       Allison MB, Myers MG, Jr. 20 years of leptin: connecting leptin signaling to biological function. The Journal of endocrinology. 2014;223(1):T25-35.

36.       Chan JL, Bluher S, Yiannakouris N, Suchard MA, Kratzsch J, Mantzoros CS. Regulation of circulating soluble leptin receptor levels by gender, adiposity, sex steroids, and leptin: observational and interventional studies in humans. Diabetes. 2002;51(7):2105-2112.

37.       Gavrilova O, Barr V, Marcus-Samuels B, Reitman M. Hyperleptinemia of pregnancy associated with the appearance of a circulating form of the leptin receptor. Journal of Biological Chemistry. 1997;272(48):30546-30551.

38.       Tu H, Kastin AJ, Hsuchou H, Pan W. Soluble receptor inhibits leptin transport. Journal of cellular physiology. 2008;214(2):301-305.

39.       Banks WA, Kastin AJ, Huang W, Jaspan JB, Maness LM. Leptin enters the brain by a saturable system independent of insulin. Peptides. 1996;17(2):305-311.

40.       Huang L, Wang Z, Li C. Modulation of circulating leptin levels by its soluble receptor. J Biol Chem. 2001;276(9):6343-6349.

41.       Halaas JL, Gajiwala KS, Maffei M, Cohen SL, Chait BT, Rabinowitz D, Lallone RL, Burley SK, Friedman JM. Weight-reducing effects of the plasma protein encoded by the obese gene. Science. 1995;269:543-546.

42.       Pelleymounter MA, Cullen MJ, Baker MB, Hecht R, Winters D, Boone T, Collins F. Effects of the obese gene product on body weight regulation in ob/ob mice. Science. 1995;269(5223):540-543.

43.       Campfield LA, Smith FJ, Guisez Y, Devos R, Burn P. Recombinant mouse OB protein: Evidence for a peripheral signal linking adiposity and central neural networks. Science. 1995;269:546-549.

44.       Montague CT, Farooqi IS, Whitehead JP, Soos MA, Rau H, Wareham NJ, Sewter CP, Digby JE, Mohammed SN, Hurst JA, Cheetham CH, Earley AR, Barnett AH, Prins JB, O'Rahilly S. Congenital leptin deficiency is associated with severe early-onset obesity in humans. Nature. 1997;387(6636):903-908.

45.       Farooqi IS, Jebb SA, Langmack G, Lawrence E, Cheetham CH, Prentice AM, Hughes IA, McCamish MA, O'Rahilly S. Effects of recombinant leptin therapy in a child with congenital leptin deficiency. NEnglJMed. 1999;341(12):879-884.

46.       Clement K, Vaisse C, Lahlou N, Cabrol S, Pelloux V, Cassuto D, Gourmelen M, Dina C, Chambaz J, Lacorte JM, Basdevant A, Bougneres P, leBouc Y, Froguel P, Guy-Grand B. A mutation in the human leptin receptor gene causes obesity and pituitary dysfunction. Nature. 1998;392:398-401.

47.       Chung WK, Belfi K, Chua M, Wiley J, Mackintosh R, Nicolson M, Boozer CN, Leibel RL. Heterozygosity for Lep(ob) or Lep(rdb) affects body composition and leptin homeostasis in adult mice. The American journal of physiology. 1998;274(4 Pt 2):R985-990.

48.       Farooqi IS, Wangensteen T, Collins S, Kimber W, Matarese G, Keogh JM, Lank E, Bottomley B, Lopez-Fernandez J, Ferraz-Amaro I, Dattani MT, Ercan O, Myhre AG, Retterstol L, Stanhope R, Edge JA, McKenzie S, Lessan N, Ghodsi M, De Rosa V, Perna F, Fontana S, Barroso I, Undlien DE, O'Rahilly S. Clinical and molecular genetic spectrum of congenital deficiency of the leptin receptor. The New England journal of medicine. 2007;356(3):237-247.

49.       Considine RV, Sinha MK, Heiman ML, Kriauciunas A, Stephens TW, Nyce MR, Ohannesian JP, Marco CC, McKee LJ, Bauer TL, Caro JF. Serum immunoreactive-leptin concentrations in normal-weight and obese humans. NEnglJMed. 1996;334(5):292-295.

50.       Frederich RC, Hamann A, Anderson S, Lollmann B, Lowell BB, Flier JS. Leptin levels reflect body lipid content in mice: evidence for diet-induced resistance to leptin action. NatMed. 1995;1(12):1311-1314.

51.       Heymsfield SB, Greenberg AS, Fujioka K, Dixon RM, Kushner R, Hunt T, Lubina JA, Patane J, Self B, Hunt P, McCamish M. Recombinant leptin for weight loss in obese and lean adults - A randomized, controlled, dose-escalation trial. Jama-Journal of the American Medical Association. 1999;282(16):1568-1575.

52.       Ahima RS, Saper CB, Flier JS, Elmquist JK. Leptin regulation of neuroendocrine systems. Front Neuroendocrinol. 2000;21(3):263-307.

53.       Schwartz MW, Seeley RJ, Tschop MH, Woods SC, Morton GJ, Myers MG, D'Alessio D. Cooperation between brain and islet in glucose homeostasis and diabetes. Nature. 2013;503(7474):59-66.

54.       Ahima RS, Prabakaran D, Mantzoros CS, Qu D, Lowell BB, Maratos-Flier E, Flier JS. Role of leptin in the neuroendocrine response to fasting. Nature. 1996;382:250-252.

55.       Leibel RL. Molecular physiology of weight regulation in mice and humans. Int J Obes (Lond). 2008;32 Suppl 7:S98-108.

56.       Shimomura I, Hammer RE, Ikemoto S, Brown MS, Goldstein JL. Leptin reverses insulin resistance and diabetes mellitus in mice with congenital lipodystrophy [In Process Citation]. Nature. 1999;401(6748):73-76.

57.       Ebihara K, Ogawa Y, Masuzaki H, Shintani M, Miyanaga F, Aizawa-Abe M, Hayashi T, Hosoda K, Inoue G, Yoshimasa Y, Gavrilova O, Reitman ML, Nakao K. Transgenic overexpression of leptin rescues insulin resistance and diabetes in a mouse model of lipoatrophic diabetes. Diabetes. 2001;50(6):1440-1448.

58.       Morrow T. Myalept approved for treatment of disorders marked by loss of body fat. Managed care. 2014;23(6):50-51.

59.       Cohen P, Zhao C, Cai X, Montez JM, Rohani SC, Feinstein P, Mombaerts P, Friedman JM. Selective deletion of leptin receptor in neurons leads to obesity. The Journal of clinical investigation. 2001;108(8):1113-1121.

60.       Patterson CM, Leshan RL, Jones JC, Myers MG, Jr. Molecular mapping of mouse brain regions innervated by leptin receptor-expressing cells. Brain research. 2011.

61.       Elmquist JK, Bjorbaek C, Ahima RS, Flier JS, Saper CB. Distributions of leptin receptor mRNA isoforms in the rat brain. The Journal of comparative neurology. 1998;395(4):535-547.

62.       Scott MM, Lachey JL, Sternson SM, Lee CE, Elias CF, Friedman JM, Elmquist JK. Leptin targets in the mouse brain. J Comp Neurol. 2009;514(5):518-532.

63.       Ring LE, Zeltser LM. Disruption of hypothalamic leptin signaling in mice leads to early-onset obesity, but physiological adaptations in mature animals stabilize adiposity levels. The Journal of clinical investigation. 2010;120(8):2931-2941.

64.       Vong L, Ye C, Yang Z, Choi B, Chua S, Jr., Lowell BB. Leptin Action on GABAergic Neurons Prevents Obesity and Reduces Inhibitory Tone to POMC Neurons. Neuron. 2011;71(1):142-154.

65.       Leshan RL, Greenwald-Yarnell M, Patterson CM, Gonzalez IE, Myers MG, Jr. Leptin action through hypothalamic nitric oxide synthase-1-expressing neurons controls energy balance. Nature medicine. 2012.

66.       van de Wall E, Leshan R, Xu AW, Balthasar N, Coppari R, Liu SM, Jo YH, MacKenzie RG, Allison DB, Dun NJ, Elmquist J, Lowell BB, Barsh GS, de Luca C, Myers MG, Jr., Schwartz GJ, Chua SC, Jr. Collective and individual functions of leptin receptor modulated neurons controlling metabolism and ingestion. Endocrinology. 2008;149(4):1773-1785.

67.       Balthasar N, Coppari R, McMinn J, Liu SM, Lee CE, Tang V, Kenny CD, McGovern RA, Chua SC, Jr., Elmquist JK, Lowell BB. Leptin Receptor Signaling in POMC Neurons Is Required for Normal Body Weight Homeostasis. Neuron. 2004;42(6):983-991.

68.       Dhillon H, Zigman JM, Ye C, Lee CE, McGovern RA, Tang V, Kenny CD, Christiansen LM, White RD, Edelstein EA, Coppari R, Balthasar N, Cowley MA, Chua S, Jr., Elmquist JK, Lowell BB. Leptin directly activates SF1 neurons in the VMH, and this action by leptin is required for normal body-weight homeostasis. Neuron. 2006;49(2):191-203.

69.       Kim KW, Zhao L, Donato J, Jr., Kohno D, Xu Y, Elias CF, Lee C, Parker KL, Elmquist JK. Steroidogenic factor 1 directs programs regulating diet-induced thermogenesis and leptin action in the ventral medial hypothalamic nucleus. Proceedings of the National Academy of Sciences of the United States of America. 2011.

70.       Donato J, Jr., Cravo RM, Frazao R, Gautron L, Scott MM, Lachey J, Castro IA, Margatho LO, Lee S, Lee C, Richardson JA, Friedman J, Chua S, Jr., Coppari R, Zigman JM, Elmquist JK, Elias CF. Leptin's effect on puberty in mice is relayed by the ventral premammillary nucleus and does not require signaling in Kiss1 neurons. The Journal of clinical investigation. 2011;121(1):355-368.

71.       Barsh GS, He L, Gunn TM. Genetic and biochemical studies of the Agouti-attractin system. J Recept Signal Transduct Res. 2002;22(1-4):63-77.

72.       Dickie MM. Mutations at the agouti locus in the mouse. The Journal of heredity. 1969;60(1):20-25.

73.       Lu D, Willard D, Patel IR, Kadwell SH, Overton L, Kost T, Luther M, Chen W, Woychik RP, Wilkison WO, Cone RD. Agouti protein is an antagonist of the melanocyte-stimulating-hormone receptor. Nature. 1994;371:799-802.

74.       Coleman DL. Obese and diabetes: two mutant genes causing diabetes-obesity syndromes in mice. Diabetologia. 1978;14(3):141-148.

75.       Cone RD. The melanocortin receptors: Agonists, antagonists, and the hormonal control of pigmentation. Recent ProgHormRes. 1996;51:287-318.

76.       Butler AA, Cone RD. The melanocortin receptors: lessons from knockout models. Neuropeptides. 2002;36(2-3):77-84.

77.       Huszar D, Lynch CA, Fairchild-Huntress V, Dunmore JH, Fang Q, Berkemeier LR, Gu W, Kesterson RA, Boston BA, Cone RD, Smith FJ, Campfield LA, Burn P, Lee F. Targeted disruption of the melanocortin-4 receptor results in obesity in mice. Cell. 1997;88:131-141.

78.       Ste ML, Miura GI, Marsh DJ, Yagaloff K, Palmiter RD. A metabolic defect promotes obesity in mice lacking melanocortin-4 receptors. ProcNatlAcadSciUSA. 2000;97(22):12339-12344.

79.       Vaisse C, Clement K, Guy-Grand B, Froguel P. A frameshift mutation in human MC4R is associated with a dominant form of obesity. Nature genetics. 1998;20:113-114.

80.       Loos RJ, Lindgren CM, Li S, Wheeler E, Zhao JH, Prokopenko I, Inouye M, Freathy RM, Attwood AP, Beckmann JS, Berndt SI, Jacobs KB, Chanock SJ, Hayes RB, Bergmann S, Bennett AJ, Bingham SA, Bochud M, Brown M, Cauchi S, Connell JM, Cooper C, Smith GD, Day I, Dina C, De S, Dermitzakis ET, Doney AS, Elliott KS, Elliott P, Evans DM, Sadaf Farooqi I, Froguel P, Ghori J, Groves CJ, Gwilliam R, Hadley D, Hall AS, Hattersley AT, Hebebrand J, Heid IM, Lamina C, Gieger C, Illig T, Meitinger T, Wichmann HE, Herrera B, Hinney A, Hunt SE, Jarvelin MR, Johnson T, Jolley JD, Karpe F, Keniry A, Khaw KT, Luben RN, Mangino M, Marchini J, McArdle WL, McGinnis R, Meyre D, Munroe PB, Morris AD, Ness AR, Neville MJ, Nica AC, Ong KK, O'Rahilly S, Owen KR, Palmer CN, Papadakis K, Potter S, Pouta A, Qi L, Randall JC, Rayner NW, Ring SM, Sandhu MS, Scherag A, Sims MA, Song K, Soranzo N, Speliotes EK, Syddall HE, Teichmann SA, Timpson NJ, Tobias JH, Uda M, Vogel CI, Wallace C, Waterworth DM, Weedon MN, Willer CJ, Wraight, Yuan X, Zeggini E, Hirschhorn JN, Strachan DP, Ouwehand WH, Caulfield MJ, Samani NJ, Frayling TM, Vollenweider P, Waeber G, Mooser V, Deloukas P, McCarthy MI, Wareham NJ, Barroso I, Kraft P, Hankinson SE, Hunter DJ, Hu FB, Lyon HN, Voight BF, Ridderstrale M, Groop L, Scheet P, Sanna S, Abecasis GR, Albai G, Nagaraja R, Schlessinger D, Jackson AU, Tuomilehto J, Collins FS, Boehnke M, Mohlke KL. Common variants near MC4R are associated with fat mass, weight and risk of obesity. Nature genetics. 2008;40(6):768-775.

81.       O'Rahilly S, Farooqi IS. Human obesity as a heritable disorder of the central control of energy balance. Int J Obes (Lond). 2008;32 Suppl 7:S55-61.

82.       Farooqi IS, Keogh JM, Yeo GS, Lank EJ, Cheetham T, O'Rahilly S. Clinical spectrum of obesity and mutations in the melanocortin 4 receptor gene. The New England journal of medicine. 2003;348(12):1085-1095.

83.       Farooqi IS, Yeo GS, Keogh JM, Aminian S, Jebb SA, Butler G, Cheetham T, O'Rahilly S. Dominant and recessive inheritance of morbid obesity associated with melanocortin 4 receptor deficiency. The Journal of clinical investigation. 2000;106(2):271-279.

84.       Speliotes EK, Willer CJ, Berndt SI, Monda KL, Thorleifsson G, Jackson AU, Allen HL, Lindgren CM, Luan J, Magi R, Randall JC, Vedantam S, Winkler TW, Qi L, Workalemahu T, Heid IM, Steinthorsdottir V, Stringham HM, Weedon MN, Wheeler E, Wood AR, Ferreira T, Weyant RJ, Segre AV, Estrada K, Liang L, Nemesh J, Park JH, Gustafsson S, Kilpelainen TO, Yang J, Bouatia-Naji N, Esko T, Feitosa MF, Kutalik Z, Mangino M, Raychaudhuri S, Scherag A, Smith AV, Welch R, Zhao JH, Aben KK, Absher DM, Amin N, Dixon AL, Fisher E, Glazer NL, Goddard ME, Heard-Costa NL, Hoesel V, Hottenga JJ, Johansson A, Johnson T, Ketkar S, Lamina C, Li S, Moffatt MF, Myers RH, Narisu N, Perry JR, Peters MJ, Preuss M, Ripatti S, Rivadeneira F, Sandholt C, Scott LJ, Timpson NJ, Tyrer JP, van Wingerden S, Watanabe RM, White CC, Wiklund F, Barlassina C, Chasman DI, Cooper MN, Jansson JO, Lawrence RW, Pellikka N, Prokopenko I, Shi J, Thiering E, Alavere H, Alibrandi MT, Almgren P, Arnold AM, Aspelund T, Atwood LD, Balkau B, Balmforth AJ, Bennett AJ, Ben-Shlomo Y, Bergman RN, Bergmann S, Biebermann H, Blakemore AI, Boes T, Bonnycastle LL, Bornstein SR, Brown MJ, Buchanan TA, Busonero F, Campbell H, Cappuccio FP, Cavalcanti-Proenca C, Chen YD, Chen CM, Chines PS, Clarke R, Coin L, Connell J, Day IN, Heijer M, Duan J, Ebrahim S, Elliott P, Elosua R, Eiriksdottir G, Erdos MR, Eriksson JG, Facheris MF, Felix SB, Fischer-Posovszky P, Folsom AR, Friedrich N, Freimer NB, Fu M, Gaget S, Gejman PV, Geus EJ, Gieger C, Gjesing AP, Goel A, Goyette P, Grallert H, Grassler J, Greenawalt DM, Groves CJ, Gudnason V, Guiducci C, Hartikainen AL, Hassanali N, Hall AS, Havulinna AS, Hayward C, Heath AC, Hengstenberg C, Hicks AA, Hinney A, Hofman A, Homuth G, Hui J, Igl W, Iribarren C, Isomaa B, Jacobs KB, Jarick I, Jewell E, John U, Jorgensen T, Jousilahti P, Jula A, Kaakinen M, Kajantie E, Kaplan LM, Kathiresan S, Kettunen J, Kinnunen L, Knowles JW, Kolcic I, Konig IR, Koskinen S, Kovacs P, Kuusisto J, Kraft P, Kvaloy K, Laitinen J, Lantieri O, Lanzani C, Launer LJ, Lecoeur C, Lehtimaki T, Lettre G, Liu J, Lokki ML, Lorentzon M, Luben RN, Ludwig B, Manunta P, Marek D, Marre M, Martin NG, McArdle WL, McCarthy A, McKnight B, Meitinger T, Melander O, Meyre D, Midthjell K, Montgomery GW, Morken MA, Morris AP, Mulic R, Ngwa JS, Nelis M, Neville MJ, Nyholt DR, O'Donnell CJ, O'Rahilly S, Ong KK, Oostra B, Pare G, Parker AN, Perola M, Pichler I, Pietilainen KH, Platou CG, Polasek O, Pouta A, Rafelt S, Raitakari O, Rayner NW, Ridderstrale M, Rief W, Ruokonen A, Robertson NR, Rzehak P, Salomaa V, Sanders AR, Sandhu MS, Sanna S, Saramies J, Savolainen MJ, Scherag S, Schipf S, Schreiber S, Schunkert H, Silander K, Sinisalo J, Siscovick DS, Smit JH, Soranzo N, Sovio U, Stephens J, Surakka I, Swift AJ, Tammesoo ML, Tardif JC, Teder-Laving M, Teslovich TM, Thompson JR, Thomson B, Tonjes A, Tuomi T, van Meurs JB, van Ommen GJ, Vatin V, Viikari J, Visvikis-Siest S, Vitart V, Vogel CI, Voight BF, Waite LL, Wallaschofski H, Walters GB, Widen E, Wiegand S, Wild SH, Willemsen G, Witte DR, Witteman JC, Xu J, Zhang Q, Zgaga L, Ziegler A, Zitting P, Beilby JP, Farooqi IS, Hebebrand J, Huikuri HV, James AL, Kahonen M, Levinson DF, Macciardi F, Nieminen MS, Ohlsson C, Palmer LJ, Ridker PM, Stumvoll M, Beckmann JS, Boeing H, Boerwinkle E, Boomsma DI, Caulfield MJ, Chanock SJ, Collins FS, Cupples LA, Smith GD, Erdmann J, Froguel P, Gronberg H, Gyllensten U, Hall P, Hansen T, Harris TB, Hattersley AT, Hayes RB, Heinrich J, Hu FB, Hveem K, Illig T, Jarvelin MR, Kaprio J, Karpe F, Khaw KT, Kiemeney LA, Krude H, Laakso M, Lawlor DA, Metspalu A, Munroe PB, Ouwehand WH, Pedersen O, Penninx BW, Peters A, Pramstaller PP, Quertermous T, Reinehr T, Rissanen A, Rudan I, Samani NJ, Schwarz PE, Shuldiner AR, Spector TD, Tuomilehto J, Uda M, Uitterlinden A, Valle TT, Wabitsch M, Waeber G, Wareham NJ, Watkins H, Wilson JF, Wright AF, Zillikens MC, Chatterjee N, McCarroll SA, Purcell S, Schadt EE, Visscher PM, Assimes TL, Borecki IB, Deloukas P, Fox CS, Groop LC, Haritunians T, Hunter DJ, Kaplan RC, Mohlke KL, O'Connell JR, Peltonen L, Schlessinger D, Strachan DP, van Duijn CM, Wichmann HE, Frayling TM, Thorsteinsdottir U, Abecasis GR, Barroso I, Boehnke M, Stefansson K, North KE, McCarthy MI, Hirschhorn JN, Ingelsson E, Loos RJ. Association analyses of 249,796 individuals reveal 18 new loci associated with body mass index. Nature genetics. 2010;42(11):937-948.

85.       Barsh G, Gunn T, He L, Wilson B, Lu X, Gantz I, Watson S. Neuroendocrine regulation by the Agouti/Agrp-melanocortin system. Endocrine research. 2000;26(4):571.

86.       Ollmann MM, Wilson BD, Yang YK, Kerns JA, Chen Y, Gantz I, Barsh GS. Antagonism of central melanocortin receptors in vitro and in vivo by agouti-related protein. Science. 1997;278(5335):135-138.

87.       Qian S, Chen H, Weingarth D, Trumbauer ME, Novi DE, Guan X, Yu H, Shen Z, Feng Y, Frazier E, Chen A, Camacho RE, Shearman LP, Gopal-Truter S, MacNeil DJ, Van Der Ploeg LH, Marsh DJ. Neither Agouti-Related Protein nor Neuropeptide Y Is Critically Required for the Regulation of Energy Homeostasis in Mice. MolCell Biol. 2002;22(14):5027-5035.

88.       Luquet S, Perez FA, Hnasko TS, Palmiter RD. NPY/AgRP neurons are essential for feeding in adult mice but can be ablated in neonates. Science. 2005;310(5748):683-685.

89.       Gropp E, Shanabrough M, Borok E, Xu AW, Janoschek R, Buch T, Plum L, Balthasar N, Hampel B, Waisman A, Barsh GS, Horvath TL, Bruning JC. Agouti-related peptide-expressing neurons are mandatory for feeding. NatNeurosci. 2005;8(10):1289-1291.

90.       Marsh DJ, Hollopeter G, Huszar D, Laufer R, Yagaloff KA, Fisher SL, Burn P, Palmiter RD. Response of melanocortin-4 receptor-deficient mice to anorectic and orexigenic peptides. NatGenet. 1999;21(1):119-122.

91.       Butler AA, Kesterson RA, Khong K, Cullen MJ, Pelleymounter MA, Dekoning J, Baetscher M, Cone RD. A unique metabolic syndrome causes obesity in the melanocortin-3 receptor-deficient mouse. Endocrinology. 2000;141(9):3518-3521.

92.       Chen AS, Marsh DJ, Trumbauer ME, Frazier EG, Guan XM, Yu H, Rosenblum CI, Vongs A, Feng Y, Cao L, Metzger JM, Strack AM, Camacho RE, Mellin TN, Nunes CN, Min W, Fisher J, Gopal-Truter S, MacIntyre DE, Chen HY, Van Der Ploeg LH. Inactivation of the mouse melanocortin-3 receptor results in increased fat mass and reduced lean body mass. NatGenet. 2000;26(1):97-102.

93.       Smart JL, Low MJ. Lack of proopiomelanocortin peptides results in obesity and defective adrenal function but normal melanocyte pigmentation in the murine C57BL/6 genetic background. Annals of the New York Academy of Sciences. 2003;994:202-210.

94.       Smart JL, Tolle V, Otero-Corchon V, Low MJ. Central dysregulation of the hypothalamic-pituitary-adrenal axis in neuron-specific proopiomelanocortin-deficient mice. Endocrinology. 2007;148(2):647-659.

95.       Krude H, Biebermann H, Luck W, Horn R, Brabant G, Gruters A. Severe early-onset obesity, adrenal insufficiency and red hair pigmentation caused by POMC mutations in humans. Nature genetics. 1998;19(2):155-157.

96.       Reizes O, Benoit SC, Strader AD, Clegg DJ, Akunuru S, Seeley RJ. Syndecan-3 modulates food intake by interacting with the melanocortin/AgRP pathway. Annals of the New York Academy of Sciences. 2003;994:66-73.

97.       Naggert JK, Fricker LD, Varlamov O, Nishina PM, Rouille Y, Steiner DF, Carroll RJ, Paigen BJ, Leiter EH. Hyperproinsulinaemia in obese fat/fat mice associated with a carboxypeptidase E mutation which reduces enzyme activity. Nature genetics. 1995;10(2):135-142.

98.       Fricker LD, Leiter EH. Peptides, enzymes and obesity: new insights from a 'dead' enzyme. Trends Biochem Sci. 1999;24(10):390-393.

99.       Chen H, Jawahar S, Qian Y, Duong Q, Chan G, Parker A, Meyer JM, Moore KJ, Chayen S, Gross DJ, Glaser B, Permutt MA, Fricker LD. Missense polymorphism in the human carboxypeptidase E gene alters enzymatic activity. Human mutation. 2001;18(2):120-131.

100.     Zhu X, Zhou A, Dey A, Norrbom C, Carroll R, Zhang C, Laurent V, Lindberg I, Ugleholdt R, Holst JJ, Steiner DF. Disruption of PC1/3 expression in mice causes dwarfism and multiple neuroendocrine peptide processing defects. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(16):10293-10298.

101.     Jackson RS, Creemers JWM, Ohagi S, Raffin-Sanson ML, Sanders L, Montague CT, Hutton JC, O'Rahilly S. Obesity and impaired prohormone processing associated with mutations in the human prohormone convertase I gene. Nature genetics. 1997;16:303-306.

102.     Berman Y, Mzhavia N, Polonskaia A, Furuta M, Steiner DF, Pintar JE, Devi LA. Defective prodynorphin processing in mice lacking prohormone convertase PC2. Journal of neurochemistry. 2000;75(4):1763-1770.

103.     Wallingford N, Perroud B, Gao Q, Coppola A, Gyengesi E, Liu ZW, Gao XB, Diament A, Haus KA, Shariat-Madar Z, Mahdi F, Wardlaw SL, Schmaier AH, Warden CH, Diano S. Prolylcarboxypeptidase regulates food intake by inactivating alpha-MSH in rodents. The Journal of clinical investigation. 2009;119(8):2291-2303.

104.     Noben-Taruth K, Naggert JK, North MA, Nishina PM. A candidate gene for the mouse mutation  tubby Nature. 1996;380(6574):534-538.

105.     Mukhopadhyay S, Jackson PK. Cilia, tubby mice, and obesity. Cilia. 2013;2:1.

106.     Gunn TM, Miller KA, He L, Hyman RW, Davis RW, Azarani A, Schlossman SF, Duke-Cohan JS, Barsh GS. The mouse mahogany locus encodes a transmembrane form of human attractin. Nature. 1999;398(6723):152-156.

107.     Kuramoto T, Kitada K, Inui T, Sasaki Y, Ito K, Hase T, Kawagachi S, Ogawa Y, Nakao K, Barsh GS, Nagao M, Ushijima T, Serikawa T. Attractin/mahogany/zitter plays a critical role in myelination of the central nervous system. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(2):559-564.

108.     Phan LK, Lin F, LeDuc CA, Chung WK, Leibel RL. The mouse mahoganoid coat color mutation disrupts a novel C3HC4 RING domain protein. The Journal of clinical investigation. 2002;110(10):1449-1459.

109.     He L, Gunn TM, Bouley DM, Lu XY, Watson SJ, Schlossman SF, Duke-Cohan JS, Barsh GS. A biochemical function for attractin in agouti-induced pigmentation and obesity. Nature genetics. 2001;27(1):40-47.

110.     He L, Lu XY, Jolly AF, Eldridge AG, Watson SJ, Jackson PK, Barsh GS, Gunn TM. Spongiform degeneration in mahoganoid mutant mice. Science. 2003;299(5607):710-712.

111.     Overton JD, Leibel RL. Mahoganoid and mahogany mutations rectify the obesity of the yellow mouse by effects on endosomal traffic of MC4R protein. J Biol Chem. 2011;286(21):18914-18929.

112.     Tong Q, Ye CP, Jones JE, Elmquist JK, Lowell BB. Synaptic release of GABA by AgRP neurons is required for normal regulation of energy balance. NatNeurosci. 2008.

113.     Wu Q, Boyle MP, Palmiter RD. Loss of GABAergic signaling by AgRP neurons to the parabrachial nucleus leads to starvation. Cell. 2009;137(7):1225-1234.

114.     Chambers AP, Woods SC. The role of neuropeptide Y in energy homeostasis. Handbook of experimental pharmacology. 2012(209):23-45.

115.     Marsh DJ, Hollopeter G, Kafer KE, Palmiter RD. Role of the Y5 neuropeptide Y receptor in feeding and obesity. Nature medicine. 1998;4(6):718-721.

116.     Sainsbury A, Schwarzer C, Couzens M, Fetissov S, Furtinger S, Jenkins A, Cox HM, Sperk G, Hokfelt T, Herzog H. Important role of hypothalamic Y2 receptors in body weight regulation revealed in conditional knockout mice. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(13):8938-8943.

117.     Batterham RL, Cowley MA, Small CJ, Herzog H, Cohen MA, Dakin CL, Wren AM, Brynes AE, Low MJ, Ghatei MA, Cone RD, Bloom SR. Gut hormone PYY(3-36) physiologically inhibits food intake. Nature. 2002;418(6898):650-654.

118.     Kirchner H, Heppner KM, Tschop MH. The role of ghrelin in the control of energy balance. Handbook of experimental pharmacology. 2012(209):161-184.

119.     Tschop M, Smiley DL, Heiman ML. Ghrelin induces adiposity in rodents. Nature. 2000;407(6806):908-913.

120.     Yang J, Brown MS, Liang G, Grishin NV, Goldstein JL. Identification of the acyltransferase that octanoylates ghrelin, an appetite-stimulating peptide hormone. Cell. 2008;132(3):387-396.

121.     Heppner KM, Tong J, Kirchner H, Nass R, Tschop MH. The ghrelin O-acyltransferase-ghrelin system: a novel regulator of glucose metabolism. Curr Opin Endocrinol Diabetes Obes. 2011;18(1):50-55.

122.     Tong J, Prigeon RL, Davis HW, Bidlingmaier M, Tschop MH, D'Alessio D. Physiologic concentrations of exogenously infused ghrelin reduces insulin secretion without affecting insulin sensitivity in healthy humans. The Journal of clinical endocrinology and metabolism. 2013;98(6):2536-2543.

123.     Castaneda TR, Tong J, Datta R, Culler M, Tschop MH. Ghrelin in the regulation of body weight and metabolism. Front Neuroendocrinol. 2010;31(1):44-60.

124.     Marston OJ, Garfield AS, Heisler LK. Role of central serotonin and melanocortin systems in the control of energy balance. Eur J Pharmacol. 2011;660(1):70-79.

125.     Sohn JW, Xu Y, Jones JE, Wickman K, Williams KW, Elmquist JK. Serotonin 2C receptor activates a distinct population of arcuate pro-opiomelanocortin neurons via TRPC channels. Neuron. 2011;71(3):488-497.

126.     Holder JL, Jr., Butte NF, Zinn AR. Profound obesity associated with a balanced translocation that disrupts the SIM1 gene. Human molecular genetics. 2000;9(1):101-108.

127.     Elmquist JK, Elias CF, Saper CB. From lesions to leptin: hypothalamic control of food intake and body weight. Neuron. 1999;22(2):221-232.

128.     Holder JL, Jr., Zhang L, Kublaoui BM, DiLeone RJ, Oz OK, Bair CH, Lee YH, Zinn AR. Sim1 gene dosage modulates the homeostatic feeding response to increased dietary fat in mice. Am J Physiol Endocrinol Metab. 2004;287(1):E105-113.

129.     Balthasar N, Dalgaard LT, Lee CE, Yu J, Funahashi H, Williams T, Ferreira M, Tang V, McGovern RA, Kenny CD, Christiansen LM, Edelstein E, Choi B, Boss O, Aschkenasi C, Zhang CY, Mountjoy K, Kishi T, Elmquist JK, Lowell BB. Divergence of melanocortin pathways in the control of food intake and energy expenditure. Cell. 2005;123(3):493-505.

130.     Shah BP, Vong L, Olson DP, Koda S, Krashes MJ, Ye C, Yang Z, Fuller PM, Elmquist JK, Lowell BB. MC4R-expressing glutamatergic neurons in the paraventricular hypothalamus regulate feeding and are synaptically connected to the parabrachial nucleus. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(36):13193-13198.

131.     Atasoy D, Betley JN, Su HH, Sternson SM. Deconstruction of a neural circuit for hunger. Nature. 2012;488(7410):172-177.

132.     Sutton AK, Pei H, Burnett KH, Myers MG, Jr., Rhodes CJ, Olson DP. Control of food intake and energy expenditure by nos1 neurons of the paraventricular hypothalamus. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2014;34(46):15306-15318.

133.     Wu Z, Xu Y, Zhu Y, Sutton AK, Zhao R, Lowell BB, Olson DP, Tong Q. An obligate role of oxytocin neurons in diet induced energy expenditure. PloS one. 2012;7(9):e45167.

134.     Morley JE, Levine AS. Corticotrophin releasing factor, grooming and ingestive behavior. Life sciences. 1982;31(14):1459-1464.

135.     Muglia L, Jacobson L, Dikkes P, Majzoub JA. Corticotropin-releasing hormone deficiency reveals major fetal but not adult glucocorticoid need. Nature. 1995;373(6513):427-432.

136.     Heinrichs SC, Lapsansky J, Lovenberg TW, De Souza EB, Chalmers DT. Corticotropin-releasing factor CRF1, but not CRF2, receptors mediate anxiogenic-like behavior. Regulatory peptides. 1997;71(1):15-21.

137.     Bale TL, Picetti R, Contarino A, Koob GF, Vale WW, Lee KF. Mice deficient for both corticotropin-releasing factor receptor 1 (CRFR1) and CRFR2 have an impaired stress response and display sexually dichotomous anxiety-like behavior. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2002;22(1):193-199.

138.     Pei H, Sutton AK, Burnett KH, Fuller PM, Olson DP. AVP neurons in the paraventricular nucleus of the hypothalamus regulate feeding. Molecular metabolism. 2014;3(2):209-215.

139.     Tran PV, Lee MB, Marin O, Xu B, Jones KR, Reichardt LF, Rubenstein JR, Ingraham HA. Requirement of the orphan nuclear receptor SF-1 in terminal differentiation of ventromedial hypothalamic neurons. Molecular and cellular neurosciences. 2003;22(4):441-453.

140.     Hawke Z, Ivanov TR, Bechtold DA, Dhillon H, Lowell BB, Luckman SM. PACAP neurons in the hypothalamic ventromedial nucleus are targets of central leptin signaling. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2009;29(47):14828-14835.

141.     Robinson S, Sotak BN, During MJ, Palmiter RD. Local dopamine production in the dorsal striatum restores goal-directed behavior in dopamine-deficient mice. BehavNeurosci. 2006;120(1):196-200.

142.     Qu D, Ludwig DS, Gammeltoft S, Piper M, Pelleymounter MA, Cullen MJ, Mathes WF, Przypek J, Kanarek R, Maratos-Flier E. A role for melanin-concentrating hormone in the central regulation of feeding behavior. Nature. 1996;380(6571):243-247.

143.     Shimada M, Tritos NA, Lowell BB, Flier JS, Maratos-Flier E. Mice lacking melanin-concentrating hormone are hypophagic and lean. Nature. 1998;396(6712):670-674.

144.     de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C, Battenberg EL, Gautvik VT, Bartlett FS, II, Frankel WN, van den Pol AN, Bloom FE, Gautvik KM, Sutcliffe JG. The hypocretins: Hypothalamus-specific peptides with neuroexcitatory activity. ProcNatlAcadSciUSA. 1998;95(1):322-327.

145.     Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams SC, Richardson JA, Kozlowski GP, Wilson S, Arch JR, Buckingham RE, Haynes AC, Carr SA, Annan RS, McNulty DE, Liu WS, Terrett JA, Elshourbagy NA, Bergsma DJ, Yanagisawa M. Orexins and orexin receptors: A family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell. 1998;92(4):573-585.

146.     Funato H, Tsai AL, Willie JT, Kisanuki Y, Williams SC, Sakurai T, Yanagisawa M. Enhanced orexin receptor-2 signaling prevents diet-induced obesity and improves leptin sensitivity. Cell Metab. 2009;9(1):64-76.

147.     de Lecea L, Sutcliffe JG. The hypocretins and sleep. FEBS J. 2005;272(22):5675-5688.

148.     Louis GW, Leinninger GM, Rhodes CJ, Myers MG, Jr. Direct innervation and modulation of orexin neurons by lateral hypothalamic LepRb neurons. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2010;30(34):11278-11287.

149.     Goforth PB, Leinninger GM, Patterson CM, Satin LS, Myers MG, Jr. Leptin acts via lateral hypothalamic area neurotensin neurons to inhibit orexin neurons by multiple GABA-independent mechanisms. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2014;34(34):11405-11415.

150.     Leinninger GM, Opland DM, Jo YH, Faouzi M, Christensen L, Cappellucci LA, Rhodes CJ, Gnegy ME, Becker JB, Pothos EN, Seasholtz AF, Thompson RC, Myers MG, Jr. Leptin action via neurotensin neurons controls orexin, the mesolimbic dopamine system and energy balance. Cell Metab. 2011;14(3):313-323.

151.     Bates SH, Stearns WH, Schubert M, Tso AWK, Wang Y, Banks AS, Dundon TA, Lavery HJ, Haq AK, Maratos-Flier E, Neel BG, Schwartz MW, Myers MG, Jr. STAT3 signaling is required for leptin regulation of energy balance but not reproduction. Nature. 2003;421:856-859.

152.     Xu AW, Ste-Marie L, Kaelin CB, Barsh GS. Inactivation of signal transducer and activator of transcription 3 in proopiomelanocortin (Pomc) neurons causes decreased pomc expression, mild obesity, and defects in compensatory refeeding. Endocrinology. 2007;148(1):72-80.

153.     Piper ML, Unger EK, Myers MG, Jr., Xu AW. Specific physiological roles for Stat3 in leptin receptor-expressing neurons. Mol Endocrinol. 2007.

154.     Lee JY, Muenzberg H, Gavrilova O, Reed JA, Berryman D, Villanueva EC, Louis GW, Leinninger GM, Bertuzzi S, Seeley RJ, Robinson GW, Myers MG, Hennighausen L. Loss of Cytokine-STAT5 Signaling in the CNS and Pituitary Gland Alters Energy Balance and Leads to Obesity. PLoSONE. 2008;3(2):e1639.

155.     Reed JA, Clegg DJ, Smith KB, Tolod-Richer EG, Matter EK, Picard LS, Seeley RJ. GM-CSF action in the CNS decreases food intake and body weight. J ClinInvest. 2005;115(11):3035-3044.

156.     Singireddy AV, Inglis MA, Zuure WA, Kim JS, Anderson GM. Neither signal transducer and activator of transcription 3 (STAT3) or STAT5 signaling pathways are required for leptin's effects on fertility in mice. Endocrinology. 2013;154(7):2434-2445.

157.     Sipols AJ, Baskin DG, Schwartz MW. Effect of intracerebroventricular insulin infusion on diabetic hyperphagia and hypothalamic neuropeptide gene expression. Diabetes. 1995;44:147-151.

158.     Bruning JC, Gautam D, Burks DJ, Gillette J, Schubert M, Orban PC, Klein R, Krone W, Muller-Wieland D, Kahn CR. Role of brain insulin receptor in control of body weight and reproduction [see comments]. Science. 2000;289(5487):2122-2125.

159.     Bruning JC, Michael MD, Winnay JN, Hayashi T, Horsch D, Accili D, Goodyear LJ, Kahn CR. A muscle-specific insulin receptor knockout exhibits features of the metabolic syndrome of NIDDM without altering glucose tolerance. MolCell. 1998;2(5):559-569.

160.     Nandi A, Kitamura Y, Kahn CR, Accili D. Mouse models of insulin resistance. Physiol Rev. 2004;84(2):623-647.

161.     White MF. Insulin signaling in health and disease. Science. 2003;302(5651):1710-1711.

162.     Lin X, Taguchi A, Park S, Kushner JA, Li F, Li Y, White MF. Dysregulation of insulin receptor substrate 2 in beta cells and brain causes obesity and diabetes. J ClinInvest. 2004;114(7):908-916.

163.     Sadagurski M, Dong XC, Myers MG, Jr., White MF. Irs2 and Irs4 synergize in non-LepRb neurons to control energy balance and glucose homeostasis. Molecular metabolism. 2014;3(1):55-63.

164.     Ren H, Orozco IJ, Su Y, Suyama S, Gutierrez-Juarez R, Horvath TL, Wardlaw SL, Plum L, Arancio O, Accili D. FoxO1 Target Gpr17 Activates AgRP Neurons to Regulate Food Intake. Cell. 2012;149(6):1314-1326.

165.     Fingar DC, Blenis J. Target of rapamycin (TOR): an integrator of nutrient and growth factor signals and coordinator of cell growth and cell cycle progression. Oncogene. 2004;23(18):3151-3171.

166.     Cota D, Proulx K, Smith KA, Kozma SC, Thomas G, Woods SC, Seeley RJ. Hypothalamic mTOR signaling regulates food intake. Science. 2006;312(5775):927-930.

167.     Villanueva EC, Munzberg H, Cota D, Leshan RL, Kopp K, Ishida-Takahashi R, Jones JC, Fingar DC, Seeley RJ, Myers MG, Jr. Complex regulation of mammalian target of rapamycin complex 1 in the basomedial hypothalamus by leptin and nutritional status. Endocrinology. 2009;150(10):4541-4551.

168.     Mori H, Inoki K, Munzberg H, Opland D, Faouzi M, Villanueva EC, Ikenoue T, Kwiatkowski D, MacDougald OA, Myers MG, Jr., Guan KL. Critical role for hypothalamic mTOR activity in energy balance. Cell Metab. 2009;9(4):362-374.

169.     Quan W, Kim HK, Moon EY, Kim SS, Choi CS, Komatsu M, Jeong YT, Lee MK, Kim KW, Kim MS, Lee MS. Role of hypothalamic proopiomelanocortin neuron autophagy in the control of appetite and leptin response. Endocrinology. 2012;153(4):1817-1826.

170.     Klaman LD, Boss O, Peroni OD, Kim JK, Martino JL, Zabolotny JM, Moghal N, Lubkin M, Kim YB, Sharpe AH, Stricker-Krongrad A, Shulman GI, Neel BG, Kahn BB. Increased energy expenditure, decreased adiposity, and tissue-specific insulin sensitivity in protein-tyrosine phosphatase 1B-deficient mice. Molecular and cellular biology. 2000;20(15):5479-5489.

171.     Zabolotny JM, Bence-Hanulec KK, Stricker-Krongrad A, Haj F, Wang Y, Minokoshi Y, Kim YB, Elmquist JK, Tartaglia LA, Kahn BB, Neel BG. PTP1B regulates leptin signal transduction in vivo. Dev Cell. 2002;2(4):489-495.

172.     Banno R, Zimmer D, De Jonghe BC, Atienza M, Rak K, Yang W, Bence KK. PTP1B and SHP2 in POMC neurons reciprocally regulate energy balance in mice. The Journal of clinical investigation. 2010;120(3):720-734.

173.     Bence KK, Delibegovic M, Xue B, Gorgun CZ, Hotamisligil GS, Neel BG, Kahn BB. Neuronal PTP1B regulates body weight, adiposity and leptin action. NatMed. 2006;12(8):917-924.

174.     Loh K, Fukushima A, Zhang X, Galic S, Briggs D, Enriori PJ, Simonds S, Wiede F, Reichenbach A, Hauser C, Sims NA, Bence KK, Zhang S, Zhang ZY, Kahn BB, Neel BG, Andrews ZB, Cowley MA, Tiganis T. Elevated hypothalamic TCPTP in obesity contributes to cellular leptin resistance. Cell Metab. 2011;14(5):684-699.

175.     Alexander WS, Starr R, Metcalf D, Nicholson SE, Farley A, Elefanty AG, Brysha M, Kile BT, Richardson R, Baca M, Zhang JG, Willson TA, Viney EM, Sprigg NS, Rakar S, Corbin J, Mifsud S, DiRago L, Cary D, Nicola NA, Hilton DJ. Suppressors of cytokine signaling (SOCS): negative regulators of signal transduction. J LeukocBiol. 1999;66(4):588-592.

176.     Bjorbaek C, Lavery HJ, Bates SH, Olson RK, Davis SM, Flier JS, Myers MG, Jr. SOCS3 mediates feedback inhibition of the leptin receptor via Tyr985. J Biol Chem. 2000;275(51):40649-40657.

177.     Mori H, Hanada R, Hanada T, Aki D, Mashima R, Nishinakamura H, Torisu T, Chien KR, Yasukawa H, Yoshimura A. Socs3 deficiency in the brain elevates leptin sensitivity and confers resistance to diet-induced obesity. NatMed. 2004.

178.     Bjornholm M, Munzberg H, Leshan RL, Villanueva EC, Bates SH, Louis GW, Jones JC, Ishida-Takahashi R, Bjorbaek C, Myers MG, Jr. Mice lacking inhibitory leptin receptor signals are lean with normal endocrine function. J ClinInvest. 2007;117(5):1354-1360.

179.     Rui L. SH2B1 regulation of energy balance, body weight, and glucose metabolism. World journal of diabetes. 2014;5(4):511-526.

180.     Ren D, Li M, Duan C, Rui L. Identification of SH2-B as a key regulator of leptin sensitivity, energy balance, and body weight in mice. Cell Metab. 2005;2(2):95-104.

181.     Ren D, Zhou Y, Morris D, Li M, Li Z, Rui L. Neuronal SH2B1 is essential for controlling energy and glucose homeostasis. J ClinInvest. 2007;117(2):397-406.

182.     Doche ME, Bochukova EG, Su HW, Pearce LR, Keogh JM, Henning E, Cline JM, Saeed S, Dale A, Cheetham T, Barroso I, Argetsinger LS, O'Rahilly S, Rui L, Carter-Su C, Farooqi IS. Human SH2B1 mutations are associated with maladaptive behaviors and obesity. The Journal of clinical investigation. 2012;122(12):4732-4736.

183.     Nogueiras R, Lopez M, Lage R, Perez-Tilve D, Pfluger P, Mendieta-Zeron H, Sakkou M, Wiedmer P, Benoit SC, Datta R, Dong JZ, Culler M, Sleeman M, Vidal-Puig A, Horvath T, Treier M, Dieguez C, Tschop MH. Bsx, a novel hypothalamic factor linking feeding with locomotor activity, is regulated by energy availability. Endocrinology. 2008;149(6):3009-3015.

184.     Lee YS, Sasaki T, Kobayashi M, Kikuchi O, Kim HJ, Yokota-Hashimoto H, Shimpuku M, Susanti VY, Ido-Kitamura Y, Kimura K, Inoue H, Tanaka-Okamoto M, Ishizaki H, Miyoshi J, Ohya S, Tanaka Y, Kitajima S, Kitamura T. Hypothalamic ATF3 is involved in regulating glucose and energy metabolism in mice. Diabetologia. 2013;56(6):1383-1393.

185.     Wright MB, Bortolini M, Tadayyon M, Bopst M. Minireview: Challenges and Opportunities in Development of PPAR Agonists. Molecular endocrinology. 2014;28(11):1756-1768.

186.     Tordjman K, Bernal-Mizrachi C, Zemany L, Weng S, Feng C, Zhang F, Leone TC, Coleman T, Kelly DP, Semenkovich CF. PPARalpha deficiency reduces insulin resistance and atherosclerosis in apoE-null mice. The Journal of clinical investigation. 2001;107(8):1025-1034.

187.     Long L, Toda C, Jeong JK, Horvath TL, Diano S. PPARgamma ablation sensitizes proopiomelanocortin neurons to leptin during high-fat feeding. The Journal of clinical investigation. 2014;124(9):4017-4027.

188.     Ryan KK, Li B, Grayson BE, Matter EK, Woods SC, Seeley RJ. A role for CNS PPARgamma in the regulation of energy balance. Nature medicine. 2011.

189.     Lu M, Sarruf DA, Talukdar S, Sharma S, Li P, Bandyopadhyay G, Nalbandian S, Fan W, Gayen JR, Mahata SK, Webster NJ, Schwartz MW, Olefsky JM. Brain PPARgamma promotes obesity and is required for the insulin-sensitizing effect of thiazolidinediones. Nature medicine. 2011.

190.     Greenberg ME, Xu B, Lu B, Hempstead BL. New insights in the biology of BDNF synthesis and release: implications in CNS function. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2009;29(41):12764-12767.

191.     Gray J, Yeo GS, Cox JJ, Morton J, Adlam AL, Keogh JM, Yanovski JA, El Gharbawy A, Han JC, Tung YC, Hodges JR, Raymond FL, O'Rahilly S, Farooqi IS. Hyperphagia, severe obesity, impaired cognitive function, and hyperactivity associated with functional loss of one copy of the brain-derived neurotrophic factor (BDNF) gene. Diabetes. 2006;55(12):3366-3371.

192.     Han JC, Liu QR, Jones M, Levinn RL, Menzie CM, Jefferson-George KS, Adler-Wailes DC, Sanford EL, Lacbawan FL, Uhl GR, Rennert OM, Yanovski JA. Brain-derived neurotrophic factor and obesity in the WAGR syndrome. The New England journal of medicine. 2008;359(9):918-927.

193.     Gray J, Yeo G, Hung C, Keogh J, Clayton P, Banerjee K, McAulay A, O'Rahilly S, Farooqi IS. Functional characterization of human NTRK2 mutations identified in patients with severe early-onset obesity. Int J Obes (Lond). 2007;31(2):359-364.

194.     Mercader JM, Saus E, Aguera Z, Bayes M, Boni C, Carreras A, Cellini E, de Cid R, Dierssen M, Escaramis G, Fernandez-Aranda F, Forcano L, Gallego X, Gonzalez JR, Gorwood P, Hebebrand J, Hinney A, Nacmias B, Puig A, Ribases M, Ricca V, Romo L, Sorbi S, Versini A, Gratacos M, Estivill X. Association of NTRK3 and its interaction with NGF suggest an altered cross-regulation of the neurotrophin signaling pathway in eating disorders. Human molecular genetics. 2008;17(9):1234-1244.

195.     Kernie SG, Liebl DJ, Parada LF. BDNF regulates eating behavior and locomotor activity in mice. The EMBO journal. 2000;19(6):1290-1300.

196.     Unger TJ, Calderon GA, Bradley LC, Sena-Esteves M, Rios M. Selective deletion of Bdnf in the ventromedial and dorsomedial hypothalamus of adult mice results in hyperphagic behavior and obesity. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2007;27(52):14265-14274.

197.     Sharma N, Berbari NF, Yoder BK. Ciliary dysfunction in developmental abnormalities and diseases. Curr Top Dev Biol. 2008;85:371-427.

198.     Zaghloul NA, Katsanis N. Mechanistic insights into Bardet-Biedl syndrome, a model ciliopathy. The Journal of clinical investigation. 2009;119(3):428-437.

199.     Davis RE, Swiderski RE, Rahmouni K, Nishimura DY, Mullins RF, Agassandian K, Philp AR, Searby CC, Andrews MP, Thompson S, Berry CJ, Thedens DR, Yang B, Weiss RM, Cassell MD, Stone EM, Sheffield VC. A knockin mouse model of the Bardet-Biedl syndrome 1 M390R mutation has cilia defects, ventriculomegaly, retinopathy, and obesity. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(49):19422-19427.

200.     Hernandez-Hernandez V, Pravincumar P, Diaz-Font A, May-Simera H, Jenkins D, Knight M, Beales PL. Bardet-Biedl syndrome proteins control the cilia length through regulation of actin polymerization. Human molecular genetics. 2013;22(19):3858-3868.

201.     Davenport JR, Watts AJ, Roper VC, Croyle MJ, van Groen T, Wyss JM, Nagy TR, Kesterson RA, Yoder BK. Disruption of intraflagellar transport in adult mice leads to obesity and slow-onset cystic kidney disease. Current biology : CB. 2007;17(18):1586-1594.

202.     Seo S, Guo DF, Bugge K, Morgan DA, Rahmouni K, Sheffield VC. Requirement of Bardet-Biedl syndrome proteins for leptin receptor signaling. Human molecular genetics. 2009;18(7):1323-1331.

203.     Berbari NF, Pasek RC, Malarkey EB, Yazdi SM, McNair AD, Lewis WR, Nagy TR, Kesterson RA, Yoder BK. Leptin resistance is a secondary consequence of the obesity in ciliopathy mutant mice. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(19):7796-7801.

204.     Stratigopoulos G, LeDuc CA, Cremona ML, Chung WK, Leibel RL. Cut-like homeobox 1 (CUX1) regulates expression of the fat mass and obesity-associated and retinitis pigmentosa GTPase regulator-interacting protein-1-like (RPGRIP1L) genes and coordinates leptin receptor signaling. J Biol Chem. 2011;286(3):2155-2170.

205.     Yang J, Loos RJ, Powell JE, Medland SE, Speliotes EK, Chasman DI, Rose LM, Thorleifsson G, Steinthorsdottir V, Magi R, Waite L, Smith AV, Yerges-Armstrong LM, Monda KL, Hadley D, Mahajan A, Li G, Kapur K, Vitart V, Huffman JE, Wang SR, Palmer C, Esko T, Fischer K, Zhao JH, Demirkan A, Isaacs A, Feitosa MF, Luan J, Heard-Costa NL, White C, Jackson AU, Preuss M, Ziegler A, Eriksson J, Kutalik Z, Frau F, Nolte IM, Van Vliet-Ostaptchouk JV, Hottenga JJ, Jacobs KB, Verweij N, Goel A, Medina-Gomez C, Estrada K, Bragg-Gresham JL, Sanna S, Sidore C, Tyrer J, Teumer A, Prokopenko I, Mangino M, Lindgren CM, Assimes TL, Shuldiner AR, Hui J, Beilby JP, McArdle WL, Hall P, Haritunians T, Zgaga L, Kolcic I, Polasek O, Zemunik T, Oostra BA, Junttila MJ, Gronberg H, Schreiber S, Peters A, Hicks AA, Stephens J, Foad NS, Laitinen J, Pouta A, Kaakinen M, Willemsen G, Vink JM, Wild SH, Navis G, Asselbergs FW, Homuth G, John U, Iribarren C, Harris T, Launer L, Gudnason V, O'Connell JR, Boerwinkle E, Cadby G, Palmer LJ, James AL, Musk AW, Ingelsson E, Psaty BM, Beckmann JS, Waeber G, Vollenweider P, Hayward C, Wright AF, Rudan I, Groop LC, Metspalu A, Khaw KT, van Duijn CM, Borecki IB, Province MA, Wareham NJ, Tardif JC, Huikuri HV, Cupples LA, Atwood LD, Fox CS, Boehnke M, Collins FS, Mohlke KL, Erdmann J, Schunkert H, Hengstenberg C, Stark K, Lorentzon M, Ohlsson C, Cusi D, Staessen JA, Van der Klauw MM, Pramstaller PP, Kathiresan S, Jolley JD, Ripatti S, Jarvelin MR, de Geus EJ, Boomsma DI, Penninx B, Wilson JF, Campbell H, Chanock SJ, van der Harst P, Hamsten A, Watkins H, Hofman A, Witteman JC, Zillikens MC, Uitterlinden AG, Rivadeneira F, Zillikens MC, Kiemeney LA, Vermeulen SH, Abecasis GR, Schlessinger D, Schipf S, Stumvoll M, Tonjes A, Spector TD, North KE, Lettre G, McCarthy MI, Berndt SI, Heath AC, Madden PA, Nyholt DR, Montgomery GW, Martin NG, McKnight B, Strachan DP, Hill WG, Snieder H, Ridker PM, Thorsteinsdottir U, Stefansson K, Frayling TM, Hirschhorn JN, Goddard ME, Visscher PM. FTO genotype is associated with phenotypic variability of body mass index. Nature. 2012;490(7419):267-272.

206.     Stratigopoulos G, Padilla SL, LeDuc CA, Watson E, Hattersley AT, McCarthy MI, Zeltser LM, Chung WK, Leibel RL. Regulation of Fto/Ftm gene expression in mice and humans. American journal of physiology Regulatory, integrative and comparative physiology. 2008;294(4):R1185-1196.

207.     Boissel S, Reish O, Proulx K, Kawagoe-Takaki H, Sedgwick B, Yeo GS, Meyre D, Golzio C, Molinari F, Kadhom N, Etchevers HC, Saudek V, Farooqi IS, Froguel P, Lindahl T, O'Rahilly S, Munnich A, Colleaux L. Loss-of-function mutation in the dioxygenase-encoding FTO gene causes severe growth retardation and multiple malformations. Am J Hum Genet. 2009;85(1):106-111.

208.     Church C, Moir L, McMurray F, Girard C, Banks GT, Teboul L, Wells S, Bruning JC, Nolan PM, Ashcroft FM, Cox RD. Overexpression of Fto leads to increased food intake and results in obesity. Nature genetics. 2010;42(12):1086-1092.

209.     Smemo S, Tena JJ, Kim KH, Gamazon ER, Sakabe NJ, Gomez-Marin C, Aneas I, Credidio FL, Sobreira DR, Wasserman NF, Lee JH, Puviindran V, Tam D, Shen M, Son JE, Vakili NA, Sung HK, Naranjo S, Acemel RD, Manzanares M, Nagy A, Cox NJ, Hui CC, Gomez-Skarmeta JL, Nobrega MA. Obesity-associated variants within FTO form long-range functional connections with IRX3. Nature. 2014;507(7492):371-375.

210.     Buiting K. Prader-Willi syndrome and Angelman syndrome. Am J Med Genet C Semin Med Genet. 2010;154C(3):365-376.

211.     Whittington J, Holland A. Neurobehavioral phenotype in Prader-Willi syndrome. Am J Med Genet C Semin Med Genet. 2010;154C(4):438-447.

212.     Ding F, Li HH, Zhang S, Solomon NM, Camper SA, Cohen P, Francke U. SnoRNA Snord116 (Pwcr1/MBII-85) deletion causes growth deficiency and hyperphagia in mice. PloS one. 2008;3(3):e1709.

213.     Bischof JM, Stewart CL, Wevrick R. Inactivation of the mouse Magel2 gene results in growth abnormalities similar to Prader-Willi syndrome. Human molecular genetics. 2007;16(22):2713-2719.

214.     Mercer RE, Kwolek EM, Bischof JM, van Eede M, Henkelman RM, Wevrick R. Regionally reduced brain volume, altered serotonin neurochemistry, and abnormal behavior in mice null for the circadian rhythm output gene Magel2. Am J Med Genet B Neuropsychiatr Genet. 2009;150B(8):1085-1099.

215.     Muscatelli F, Abrous DN, Massacrier A, Boccaccio I, Le Moal M, Cau P, Cremer H. Disruption of the mouse Necdin gene results in hypothalamic and behavioral alterations reminiscent of the human Prader-Willi syndrome. Human molecular genetics. 2000;9(20):3101-3110.

216.     Grill HJ. Distributed neural control of energy balance: contributions from hindbrain and hypothalamus. Obesity(SilverSpring). 2006;14 Suppl 5:216S-221S.

217.     Bray GA. Afferent signals regulating food intake. The Proceedings of the Nutrition Society. 2000;59(3):373-384.

218.     Moran TH. Unraveling the obesity of OLETF rats. Physiology & behavior. 2008;94(1):71-78.

219.     Campbell JE, Drucker DJ. Pharmacology, physiology, and mechanisms of incretin hormone action. Cell Metab. 2013;17(6):819-837.

220.     Marks DL, Ling N, Cone RD. Role of the central melanocortin system in cachexia. Cancer Res. 2001;61(4):1432-1438.

221.     Yao JH, Ye SM, Burgess W, Zachary JF, Kelley KW, Johnson RW. Mice deficient in interleukin-1beta converting enzyme resist anorexia induced by central lipopolysaccharide. The American journal of physiology. 1999;277(5 Pt 2):R1435-1443.

222.     Dong ZM, Gutierrez-Ramos JC, Coxon A, Mayadas TN, Wagner DD. A new class of obesity genes encodes leukocyte adhesion receptors. ProcNatlAcadSciUSA. 1997;94:7526-7530.

223.     Hirsch E, Irikura VM, Paul SM, Hirsh D. Functions of interleukin 1 receptor antagonist in gene knockout and overproducing mice. Proceedings of the National Academy of Sciences of the United States of America. 1996;93(20):11008-11013.

224.     Hotamisligil GS, Shargill NS, Spiegelman BM. Adipose expression of tumor necrosis factor-à: Direct role in obesity-linked insulin resistance. Science. 1993;259:87-91.

225.     Hotamisligil GS. The role of TNFalpha and TNF receptors in obesity and insulin resistance. JInternMed. 1999;245(6):621-625.

226.     Lumeng CN, Saltiel AR. Inflammatory links between obesity and metabolic disease. The Journal of clinical investigation. 2011;121(6):2111-2117.

227.     Ferrante AW, Jr. The immune cells in adipose tissue. Diabetes, obesity & metabolism. 2013;15 Suppl 3:34-38.

228.     Cai D, Liu T. Hypothalamic inflammation: a double-edged sword to nutritional diseases. Annals of the New York Academy of Sciences. 2011;1243:E1-39.

229.     Ozcan L, Ergin AS, Lu A, Chung J, Sarkar S, Nie D, Myers MG, Jr., Ozcan U. Endoplasmic reticulum stress plays a central role in development of leptin resistance. Cell Metab. 2009;9(1):35-51.

230.     Kozak LP, Harper ME. Mitochondrial uncoupling proteins in energy expenditure. Annu Rev Nutr. 2000;20:339-363.

231.     Lowell BB, S-Susulic V, Hamann A, Lawitts JA, Himms-Hagen J, Boyer BB, Kozak LP, Flier JS. Development of obesity in transgenic mice after genetic ablation of brown adipose tissue. Nature. 1993;366:740.

232.     Bachman ES, Dhillon H, Zhang CY, Cinti S, Bianco AC, Kobilka BK, Lowell BB. betaAR signaling required for diet-induced thermogenesis and obesity resistance. Science. 2002;297(5582):843-845.

233.     Cummings DE, Brandon EP, Planas JV, Motamed K, Idzerda RL, McKnight GS. Genetically lean mice result from targeted disruption of the RII beta subunit of protein kinase A. Nature. 1996;382(6592):622-626.

234.     Kershaw EE, Flier JS. Adipose tissue as an endocrine organ. JClinEndocrinolMetab. 2004;89(6):2548-2556.

235.     Masuzaki H, Yamamoto H, Kenyon CJ, Elmquist JK, Morton NM, Paterson JM, Shinyama H, Sharp MG, Fleming S, Mullins JJ, Seckl JR, Flier JS. Transgenic amplification of glucocorticoid action in adipose tissue causes high blood pressure in mice. The Journal of clinical investigation. 2003;112(1):83-90.

 

Andrenal Androgens and Ageing

Introduction

Dehydroepiandrosterone (DHEA) and its active metabolite DHEA sulfate (DHEAS), are steroid hormones synthesized and excreted primarily by the zona reticularis of the adrenal cortex in response to adrenocorticotropic hormone (ACTH). They exert weak androgenic effects and are therefore considered precursor hormones that need to be transformed to potent androgens or estrogens to exert their effects. The potential clinical roles of DHEA/DHEAS have been studied extensively, as previous epidemiologic and prospective studies associated the age-related decrease of DHEA/DHEAS levels with higher prevalence of degenerative disorders and increased frailty and mortality from all causes in the elderly, attributing to adrenal androgens anti-ageing properties. But do they really suggest that they are hormones related to longevity or just another pointless alchemy against ageing? This chapter summarizes the physiology and pathophysiology of adrenal androgen synthesis, secretion and action and provides current evidence regarding their efficacy in the management of aging-related disorders.

  1. The Adrenal androgens

1a. The adrenal cortex; embryology and normal structure

The adrenal cortex is derived from the mesoderm lining the posterior abdominal wall. The fetal cortex begins its development in the 5-week-old fetus. At 2 months of gestation it is already identifiable as a separate organ and is composed of the inner fetal zone (85% of the cortex) and the outer permanent definitive zone. The anatomic relation of the fetal and definitive zones is maintained during gestation; at birth the adrenal glands are 10–20 times larger than the adult gland, relative to kilograms of body weight. After birth, the fetal zone undergoes rapid involution resulting in rapid decrease of the adrenocortical weight in the 3 months following birth. During the next 3 years, the adult adrenal cortex develops from cells of the outer layer of the cortex and differentiates into the three adult zones, the subcapsular zona glomerulosa, the zona fasciculate, which is the thickest zone (70% of the cortex), and the inner zona reticularis.

1b. Biosynthesis of Adrenal Androgens

The adrenal cortex produces many steroid hormones among which the major ones are cortisol, aldosterone and the adrenal androgens. The subcapsular zona glomerulosa produces aldosterone while the inner two zones fasciculata and reticularis appear to function as a unit and produce cortisol, androgens and small amounts of estrogens under the regulatory effect of ACTH and maybe of some other factors produced within the adrenal gland, including neurotransmitters, neuropeptides and nitric oxide. The biosynthetic pathway of the adrenal androgens is shown below (Fig. 1).

Quantitatively, the most abundantly produced adrenal androgens are dehydroepiandrosterone (DHEA) and its sulphated form dehydroepiandrosterone sulphate (DHEAS); the later is the most abundantly produced adrenal steroid. It also has a long half-life and provides a stable pool of circulating DHEA. The ovaries also synthesize DHEA; however, they lack the enzyme DHEA-sulphotransferase so that DHEAS is almost exclusively synthesized and secreted by the adrenals. DHEA is further metabolized to androstenedione [1,2], which may in turn be aromatized to estrone. Whether the adrenals may also produce small amounts of testosterone by further metabolism of androstenedione is controversial [3]. Although DHEA and DHEAS are secreted in greater quantities, androstenedione is qualitatively more important since it is more readily converted to testosterone in peripheral tissues. Roughly, the relative androgenic potency of DHEA, androstenedione, testosterone and dihydrotestosterone (DHT) is 5:10:100:300, respectively [4]. As ACTH is the main regulator of adrenal androgen production in adults, both DHEA and androstenedione exhibit circadian periodicity in concert with ACTH and cortisol and their plasma concentrations increase rapidly following ACTH administration; also they are suppressed by glucocorticoid administration. Because of its slow metabolic clearance, DHEAS does not exhibit diurnal rhythm variation.

Steroid biosynthesis in the adrenal cortex.Figure 1. Steroid biosynthesis in the adrenal cortex.

1c. Circulation of Adrenal Androgens

The adrenal androgens are secreted in an unbound state. Soon after their release in the circulation they bind to plasma proteins, chiefly to albumin (90%). Androstenedione, DHEA and DHEAS circulate weakly bound to albumin, while testosterone is bound extensively to the sex hormone binding globulin (SHBG). Bound steroids are biologically inactive; the unbound steroids are free to interact with target cells either to exert their effects or to be transformed into inactive or active metabolites.

1d. Metabolism of Adrenal Androgens; Gender-dependent Synthesis of DHEA/DHEAS

Due to lack or only minor inherent steroidogenic activity, adrenal androgens are precursor hormones (pro-hormones) that need to be transformed to potent androgens or estrogens to exert their effects [5,6]. Their transformation into active sex steroids depends upon the level of expression of the various steroidogenic and metabolizing enzymes in each cell type which allows all androgen-sensitive and estrogen-sensitive tissues to have some control over the local levels of sex steroids according to their needs [7]. Active androgens and estrogens thus synthesized exert their activity in the target cells with little diffusion, resulting in low levels in the general circulation. This intracrine mechanism serves to eliminate the exposure of other tissues to androgens or estrogens, minimizing unwanted side effects [8-11].

In males with normal gonadal function, the conversion of adrenal androgens to testosterone accounts for less than 5% of the total amount of this hormone, and thus the physiologic effect is negligible. In females of reproductive age, the adrenal contribution to total androgen production is more important; during the follicular phase, the adrenal precursors account for 2/3 of total testosterone production and 1/2 of DHT production. During midcycle, the ovarian contribution increases, and the adrenal precursors account for only 40% of testosterone production.

Apart from their peripheral conversion to more potent androgens, the adrenal androgens may be also aromatized to estrogens [5,6] or undergo degradation and inactivation (Fig 2). In more detail, DHEA is readily converted within the adrenal gland to DHEAS. DHEA secreted by the adrenal glands and the ovaries is also converted to DHEAS by the liver and the kidneys or it may be converted to Δ4-androstenedione. The adrenally produced DHEAS may be excreted without further metabolism or it may further undergo limited conversion to DHEA. Both DHEAS and DHEA may be metabolized to 7alpha- and 16alpha-hydroxylated derivatives and by 17β reduction to Δ5-Androstenediol and its sulfate. Androstenedione is converted either to testosterone or by reduction of its 4,5 double bond to etiocholanolone or androsterone, which may be further converted by 17 alpha reduction to etiocholanediol and androstanediol, respectively. Testosterone is converted to DHT in androgen-sensitive tissues by 5 alpha reduction and it in turn is mainly metabolized by 3 alpha reduction to androstanediol. The metabolites of these androgens are conjugated either as glucuronides or sulfates and excreted in the urine. Regarding aromatization to estrogens, it was shown that not only androstenedione and testosterone, but also DHEA, may be converted to estrogens in peripheral tissues such as brain, bone, breast and ovaries [7,12]; this might be of importance, especially in women during the menopausal transition (see below) [13,14].

Metabolism of adrenal androgens ; HSD3B, 3 β-hydroxysteroid dehydrogenase isozymes; HSD17B, 17 β -hydroxysteroid dehydrogenase isozymes; SRD5A, 5α -reductase isozymes; AKR1C, aldo-keto reductases 1C; CYP19, P450 aromatase.

Figure 2. Metabolism of adrenal androgens ; HSD3B, 3 β-hydroxysteroid dehydrogenase isozymes; HSD17B, 17 β -hydroxysteroid dehydrogenase isozymes; SRD5A, 5α -reductase isozymes; AKR1C, aldo-keto reductases 1C; CYP19, P450 aromatase.

1e. Age-dependent synthesis of DHEA/DHEAS

Fetal DHEA and DHEAS fall rapidly after birth and remain low until adrenarche; they then start rising again and peak during the third decade of life after which the serum levels of DHEA and DHEAS progressively decline with advancing age by around 2–5% per year [12,15], so that by menopause the DHEA level has decreased by 60% [16], and by 80-90% of the peak production by the eighth or ninth decade of life [17, 18]. This decline has been termed “adrenopause” in spite of the fact that cortisol secretion does not decline with age or may even increase [19,20]. Adrenopause is independent of menopause and occurs in both sexes as a gradual process at similar ages. A decrease in 17,20-lyase activity may be responsible for the progressive diminution of DHEA and DHEA-S with advancing age [21], although other mechanisms, such as a reduction in the mass of the zona reticularis [22] or a decrease in IGF-I and IGF-II [23] have also been proposed. Recent study by Heaney et al.[19] in accordance with previous research [24] found that older subjects exhibited lower plasma and saliva DHEA levels overall, while with increasing age, the DHEA area under the curve was attenuated and the slope of decline became less steep.

Although DHEAS concentration does not vary throughout the day, DHEA secretion exhibits a diurnal rhythm similar to that of cortisol. Studies have indicated that DHEA secretion is reduced in the morning period resulting in a flatter diurnal rhythm among the oldest old, in contrast to cortisol which remains stable or even increases in the morning [19,25]. The above diurnal rhythms of cortisol and DHEA, lead to an elevated cortisol:DHEA ratio, which is most pronounced in the morning period.

The age-related decline in DHEA/DHEAS levels shows high inter-individual variability [22]. There is a clear sex difference in DHEA/DHEAS concentrations with lower DHEAS concentrations in adult women compared to men [26], while there is also a clear genetic component predetermining circulating DHEA/DHEAS. Notably, data from the largest population-based twin study to estimate the genetic and environmental contributions of diurnal DHEAS concentrations demonstrated that salivary DHEAS is a heritable measure, with genetic effects accounting for 37%–46% of the total variance for the late morning and afternoon age-adjusted measures [27].

Since DHEA is the main source of androgens in women, its age-related decline leads to a corresponding decrease in the total androgen pool. Although there is actually no defined level of androgen below which women can be said to be deficient, the decline of DHEA in postmenopausal women would suggest they are “deficient” in both estrogens and androgens [16]. The declining circulating levels of adrenal androgens with advancing age have been related to clinical symptoms and disorders (see below).

In the last few years, the concept that adrenal androgen production gradually declines with advancing age has changed following the analysis of the longitudinal data collected in the Study of Women’s Health Across the Nation (SWAN) [28]. When the annual serum levels of DHEAS were aligned according to ovarian status [29], it was recognized that despite the overall age-related decline in DHEAS, in the majority of women (85% of those studied) the adrenal androgen production actually rose during the menopausal transition, starting in the early peri-menopause and continuing into the early post-menopause. The DHEAS rise was attributed to the adrenals and not the ovaries, as a similar rise was also observed in intact and ovariectomized women [30]; the gender-related rise of adrenal DHEAS and the time course of that rise that returns to a progressive decline following menopause, implies ovarian influences over adrenal steroidogenesis [31]. Considering previous failure to adequately attribute phenotype, symptoms, and health trajectories to the observed longitudinal changes in circulating estradiol and progesterone [32], the perimenopausal rise in adrenal androgens could potentially suggest a more important role of these hormones in the occurrence of symptoms during the menopausal transition [33]. The observational, epidemiologic, and interventional studies addressing this hypothesis are analyzed below.

1f. Biologic effects of Adrenal Androgens; Cellular and molecular actions

Role as pro-hormones

DHEA possesses pleiotropic effects. Epidemiologic and prospective studies have associated the decline of circulating levels of androgens with the development and progression of degenerative disorders. The exact mechanism of action and clinical role of DHEA and DHEAS, if any, remain unclear. Due to lack or only minor inherent steroidogenic activity, the adrenal androgens need to be transformed to potent androgens or estrogens to exert their effects on peripheral tissues. Recent data suggest additional direct actions of the adrenal androgens further to those exerted through the androgen and estrogen receptors (see below).

The principal biologic effects of the adrenal androgens typically seen during adrenarche consist mainly of pubic and axillary hair growth and the change of sweat composition that produces adult body odor [34]. During the reproductive years, in males with normal gonadal function, the adrenal androgens account for less than 5% of the daily production rate of testosterone and thus the physiologic effect is negligible. When produced in excess however, the adrenal androgens may have no clinical consequences in adult males or result in LH /FSH suppression and oligospermia/infertility. In boys, the adrenal androgen excess is associated with clinical manifestations including premature penile enlargement, early development of secondary sexual characteristics, premature closure of the epiphyseal growth plates and short final height. In females the excessive production of adrenal steroids as seen in Cushing syndrome, adrenal carcinoma, and congenital adrenal hyperplasia via peripheral conversion to testosterone and eventually to DHT result in acne, hirsutism and menstrual/fertility defects or even virilization in more severe cases.

Membrane-associated DHEA Receptors

Further to their effect via the estrogen and androgen receptors, recent data support direct actions of DHEA through specific Gi protein-coupled membrane receptors in bovine aortic and primary human umbilical vein endothelial cells (HUVECs) [35-37] through which DHEA activates the endothelial NO synthetase (eNOS) (eNOS/cGMP pathway) [38] and increases the production of nitric oxide (NO), a key modulator of vascular function, by endothelial cells. Such receptors are also seen in the kidney, heart, and liver but at lower level than that in bovine aortic endothelial cells [38] as well as in pulmonary artery smooth muscle cells (PASMCs), where DHEA inhibits voltage-dependent T type Ca-channels [39]. In systemic circulation, a plasma membrane receptor has been suggested in the anti-remodeling action of DHEA involving inhibition of the Akt/GSK-3β signaling pathway [40]. Other studies have shown inhibitory effect of DHEA on proliferation and apoptosis of endothelial and vascular smooth muscle cells independently of both estrogen and androgen receptors [41,42]. The above suggest the presence of a membrane-associated DHEA specific receptor; the molecular structure of this receptor remains to be elucidated.

Cytosolic/nuclear receptors

Steroid action involves cytosolic/nuclear hormone receptors [43]; thus, most of the studies looking at the mechanism(s) responsible for DHEA action focused on such receptors [44]. However, since DHEA can be metabolized into androgens/estrogens, it is not always easy to determine whether DHEA exerts its effects directly through the estrogen/androgen receptors or after conversion to these hormones. There is some new evidence showing that DHEA and some of its metabolites either bind to or activate nuclear receptors such as pregnane X receptor, constitutive androstanol receptor, estrogen receptor-β and peroxisome proliferators activated receptors [45-48]. Through the activation of peroxisome proliferator-activated receptor alpha for example, DHEA inhibits the activation of nuclear factor-κB and the secretion of interleukin-6 and interleukin-12, through which DHEA exerts anti-inflammatory effects [49,50]. 7-and 7-hydroxylated derivatives of DHEA also seem to have direct effects on nuclear receptors, but their physiological function is not clear [38]. Finally, DHEA inhibits apoptosis and promotes proliferation of osteoblasts in rats through MAPK signaling pathways, independently from androgens and estrogens [51]; this action could be beneficial for preservation of bone mass and reduction of fracture risk.

Endoplasmic reticulum receptor: sigma-1 receptor

More recently, it has been suggested that DHEA is an agonist of sigma-1 receptor (Sigma-1R) expressed in the endoplasmic reticulum of the heart, kidney, liver and brain [52,53]. Under physiological conditions, the sigma-1 receptor chaperones the functional inositol 1,4,5 trisphosphate receptor at the endoplasmic reticulum participating in the calcium signaling pathway [52, 54]. Animal studies have shown that via sigma-1R, but also by Akt– eNOs signaling pathway stimulation, DHEA may improve cardiac function [55] and exert vasculo-protective effects [56,57]. There is a great volume of data suggesting antioxidant properties of DHEA; overproduction of oxygen-free radicals (oxidative stress) upregulates inflammation and cellular proliferation and is believed to play a critical role in the development of cancer, atherosclerosis, and Alzheimer's disease, as well as the basic aging process [58-61]. DHEA inhibits glucose-6-phosphate dehydrogenase (G-6-PDH) [62,63] and NADPH production. The decrease in NADPH levels results in reduced oxygen-free radical production via NADPH oxidase [63].

In summary, DHEA mediates its action via transformation into androgen and estrogen derivatives acting through their specific receptors, but also via multiple signaling pathways involving specific membrane, cytosolic/nuclear and endoplasmic reticulum receptors.

  1. Potential treatment benefits. Treatment modalities

Data from epidemiologic and prospective studies indicate an inverse relation between low circulating levels of DHEA and DHEA-S and a host of ageing-associated pathologies such as sexual dysfunction, mood defects and poor sense of well-being [30,31], as well as higher risk of hospital admission [64], poor muscle strength [65,66] and mobility [67], and higher prevalence of frailty [68,69], insulin resistance, obesity, cardiovascular disease [70] and mortality from cardiovascular disease [71]. At the same time, a positive relation between higher levels of DHEA-S and better health and well-being was documented [72]. Furthermore, animal (primarily rodent) studies have suggested many beneficial effects of DHEA treatment, including improved immune function and prevention of atherosclerosis, cancer, diabetes, and obesity. Therefore, the therapeutic role of DHEA replacement as an anti-ageing factor for the prevention and/or treatment of the above conditions was studied; recent systematic reviews of the reports do not seem promising, however[73-79].

Treatment modalities

DHEA is considered as a hormone in Europe and thus becomes available only by prescription, while in the United States it is considered as a nutritional supplement and is sold over the counter without a prescription. This difference has no scientific foundation and is mostly a matter of declaration. Most DHEA supplements are made in laboratories from a substance called diosgenin, a plant sterol found in soy and wild yams. DHEA supplements were taken off the U.S. market in 1985 because of their unproven safety and effectiveness, but were reintroduced as a dietary supplement after the Dietary Supplement Health and Education Act was passed in 1994. At present, questionable over-the-counter DHEA preparations lacking pharmacokinetic and pharmacodynamic data are widely used in the United States. There is no standard dosage of DHEA replacement; some studies have used between 25 and 200 milligrams a day, or sometimes even higher amounts. DHEA in current preparations has a long half-life [44], which allows a single intake a day. Target levels of DHEA are around the middle of normal range for healthy young subjects, controlled by a blood sample 24 hr after the last intake [80].

The adrenal androgens are mainly thought to act as prohormones and exert at least part of their action via conversion to androgens and/or estrogens. Previous studies have shown that the end- products of DHEA supplementation depend on the patient’s gender, with a non-symmetrical transformation of DHEA favoring androgens in women and estrogens in men [73,81,82]. The above refer to oral administration of DHEA supplements; percutaneous administration of DHEA seems to provoke similar increases in both estrogens and androgens in the two genders [83].

  1. Low DHEA/DHEAS levels and associated comorbidities

3a. DHEA and Musculoskeletal disorders

The increasing incidence of fractures with advancing age has been related, among other factors, with the ageing-related reduced muscle mass and strength, that increase the propensity for falling [84,85]. A body of evidence exists on the effect of circulating DHEA/DHEAS on various markers of strength and physical function in older individuals. Studies in elderly individuals support a positive relation between DHEA blood levels and muscle mass [65], muscle strength [65,66] and mobility [67], as well as better self-reported [86] and objectively assessed physical function [87], and measured peak volume of oxygen consumed per minute [88] in elderly with higher DHEA/DHEAS concentrations. In this direction, higher DHEAS levels were associated with increased bone mass density (BMD) in both men [89] and post-menopausal women [90] and inversely related to risk for falls [91]. Finally, low DHEAS levels have been associated with a higher prevalence of frailty, a geriatric syndrome of loss of reserve characterized by weight loss, fatigue, weakness and vulnerability to adverse events [68, 69], and low back pain in both genders and slow rehabilitation of low-back pain in women [72,92,93].

Reports from interventional studies support a therapeutic role of DHEA replacement in ageing-associated musculoskeletal defects. For example, DHEA exerted positive effects on muscle strength, body composition [94-97] and physical performance [98], as well as on bone mass density (BMD) in both lumbar spine and the hip [70,73,95,99-103] when administered to post-menopausal women and elderly people over a period of 52 weeks. The above positive effects on musculoskeletal system were attributed to the DHEA-related increase of insulin-like growth factor-1 (IGF-1) levels [104,105] and bioavailability (decrease of insulin growth factor binding protein-1 [IGFBP-1]) [105] in both men and women and/or to the increase of androgen levels mostly in women [104-106]. Some other data also suggest aromatase activity of primary human osteoblasts converting DHEA to estrone [90], while it was shown in vitro that DHEA inhibits apoptosis and promotes proliferation of rat osteoblasts through MAPK signaling pathways, independently from androgen and estrogen effects [51]. The above support a positive effect of DHEA on bone through conversion to estrogens, but also independently from its hormonal end-products. Other studies, however, failed to show a beneficial effect of DHEA supplementation on muscle function [107-111] or on BMD [94,98,112]; of note all of these studies were conducted over a shorter period (26 weeks only). Whether these conflicting data result from DHEA’s mild/moderate effect or from great differences between study designs, such as short duration of treatment and small number of participants, is difficult to say. Overall, the effect of DHEA supplementation on BMD is small in relation to other treatments for bone loss, and no fracture data are available. Therefore, its therapeutic utility in rehabilitation and/or fracture/frailty prevention and treatment protocols for older patients remains unclear.

A recent systematic review [74] of the literature [73, 94,111,113-117] concluded that overall, the benefit of DHEA on muscle strength and physical function in older adults remains inconclusive. Some measures of muscle strength may improve, although DHEA does not appear to routinely benefit measures of physical function or performance. Therefore, consensus has not been reached. Further large clinical trials are necessary to better identify the clinical role of DHEA supplementation in this population.

3b. DHEA, Well-being and Sexual Function

If DHEA’s effects on musculoskeletal disorders are inconclusive, its utility for the management of ageing-related poor sense of well-being and sexual dysfunction is a question for top puzzle solvers. What we know so far from epidemiologic studies is that sexual function problems are common among women [118,119] and increase with increasing age [119,120]. The sex steroid hormones estrogens and androgens seem to play an important role in sexual life in women; androgens affect the arousability, pleasure, and intensity of orgasm in women and are particularly implicated in the neurovascular smooth muscle response of swelling and lubrication, whereas estrogens contribute to vulval and vaginal congestive response and affect mood and sexual responses [121]. Conditions such as menopausal symptoms, loss of libido, vulvovaginal atrophy-related sexual dysfunction and poor sense of well-being seen in menopausal and peri-menopausal women were related to the age-associated decline in sex steroids [122]. Furthermore, interventional studies in postmenopausal women with estrogens have shown much improvement on vaginal atrophy and vasomotor symptoms [122-125]; there is also much clinical evidence for the efficacy of testosterone reatment for low sexual function in women [126-131].

Given that a) the adrenal steroids are the most abundant sex steroids in post-menopausal women and provide a large reservoir of precursors for the intracellular production of androgens and estrogens in non-reproductive tissues, b) DHEA levels decline with age, c) pre- and post-menopausal women with lower sexual responsiveness have lower levels of serum DHEAS [132] and d) treatment of postmenopausal women with estrogen and testosterone have shown some improvement in sexual function, it was proposed that restoring the circulating levels of DHEA to those found in young women may improve sexual function and well-being in postmenopausal women [81]. Some early randomized trials that suffered from methodological issues, such as small number of participants, short treatment duration and supraphysiological doses, demonstrated positive effects of DHEA replacement on sexual function and well-being [70,133-136], as well as on relief of menopausal symptoms [136,138]. Similarly, women with adrenal insufficiency treated with oral DHEA replacement demonstrated significant improvement in overall well-being, as well as in frequency of sexual thoughts, sexual interest, and satisfaction [139, 140]. Other studies, however, failed to show any benefit of DHEA replacement on sexual function, well-being and menopausal symptoms in peri- and post-menopausal women [75,76,105,141,142] and women with adrenal insufficiency [143-145]. A recent review of the available data concluded that current evidence does not support the routine use of DHEA in women with adrenal insufficiency [77]. Furthermore, the more recent placebo-controlled randomized trials that are of superior design compared to the early trials, as they use validated measures of sexual function, have larger sample sizes and are of longer duration, failed to document any significant benefit of oral DHEA therapy on well-being or sexual function in women [73,75,76,78]. It has been hypothesized that the efficacy of DHEA to improve sexual function might be dependent on the route of its administration. In women, androgens and estrogens are produced from DHEA in the vagina tissue. As vaginal atrophy and dryness are common symptoms of estrogen deficiency during menopause, causing dyspareunia and sexual dysfunction [146], a possible benefit that emerged is that vaginally administered DHEA may improve the postmenopausal vaginal atrophy-related sexual dysfunction [147] without increasing the circulating levels of estrogen above the postmenopausal range [81,147-149]. Despite initial promising, beneficial effects on sexual function, again, even with intravaginally administered DHEA, a study failed to show significant benefits [78].

Apart from women, lower circulating levels of DHEA were also related to erectile dysfunction in men. A double-blind, placebo-controlled study that enrolled men with erectile dysfunction treated with per os with DHEA 50 mg daily has shown some promise for improving sexual performance in men who had low DHEA blood levels [150]. However, high-quality studies have demonstrated inconsistent results regarding DHEA supplementation for improving sexual function, libido, and erectile dysfunction. Although research in this area is promising, additional well-designed studies are required.

3c. DHEA and mood disorders

The prevalence of depression increases in cohorts of the elderly and has been independently related to high morbidity and mortality [151]. In the central nervous system, DHEA is considered a neurosteroid with a wide range of functions. Animal studies demonstrated several DHEA-modulated neurotransmitters, including dopamine, glutamate, and c-amino butyric acid [38], as well as DHEA-induced increased activity of 5-hydroxytryptamine (5-HT) neurons [152], providing the cellular basis for a potential antidepressant effect of DHEA. Furthermore, typical neuroleptic-like effects of DHEA were displayed in animal models of schizophrenia suggesting potential role of DHEA replacement in the treatment of schizophrenia [153].

Previous studies suggested a strong relation between low levels of DHEA/DHEA-S and major depression in children and adolescents [154], as well as adults and the elderly [155, 156]. On the contrary, higher DHEA-S levels were positively associated with depressive symptoms during the menopausal transition [157] and depression in patients with major depression [158, 159]; whether the elevated DHEA-S levels in the above studies represent increased adrenal activity that could explain the depressive symptoms is not clear, as cortisol was not measured. Moreover, successful treatment of depression was followed by reductions in both DHEA-S [158, 160] and DHEA levels [160], making the relation between DHEA/DHEAS and depression even more confusing.

Several interventional studies have shown that DHEA replacement may improve negative and depressive symptoms [134,135,161-163]. In women with adrenal insufficiency, in particular, oral DHEA replacement significantly improved the overall well-being, as well as scores for depression and anxiety [108]; similar results were found in the management of the negative symptoms of schizophrenia [163]. Most recent placebo-controlled randomized trials, however, failed to demonstrate a beneficial effect of DHEA therapy on mood, quality of life, perceptions of physical and emotional health, and life satisfaction in postmenopausal women [73,75,76]. Thus the therapeutic role of DHEA on mood disorders remains unclear.

3d. DHEA & psychosocial stress

It has long been suggested that long-term psychosocial stress may cause or contribute to different diseases and symptoms, including atherosclerosis [164], coronary heart disease [165] and acute coronary events [166], as well as accelerated aging [167,168]. Whether DHEA/DHEAS levels are related to psychological stress or not is still debatable. Exposure to prolonged psychosocial stress has been related to reduced [169-171] or elevated levels of DHEA/DHEA-S [172], while some other studies failed to show any clear association in any direction [173,174]. A recent study by Lennartsson et al. demonstrated that DHEA and DHEA-S levels are markedly lower in individuals that report perceived stress at work than in individuals who report no perceived stress at work [175]. Whether this is of clinical importance is not clear.

The incidence of dementia increases exponentially with increasing age in both men and women [176]. The number of elderly people nowadays is the fastest growing segment of the population, which means the related personal, social, and economic burdens are extremely high and could increase dramatically over the next few decades. Therefore, effective prevention/treatment of neurodegenerative disorders is imperative. It has been proposed that DHEA and DHEAS may exert neuroprotective effects in the brain mainly through DHEA-dependent neural stem cell stimulation, genomic activity modulation, and upregulation of androgen receptor levels [177,178], as well as via the DHEA-induced inhibition of pro-inflammatory factor production, such as tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6) [38] that are involved in the pathogenesis of the amyloid plaques of Alzheimer disease [179]. Higher serum levels of DHEAS have been related to more favorable cognitive function in older people, such as better concentration and working memory [180,181] and higher scores on the Mini Mental State Examination [182]. In this direction, low DHEA/DHEAS levels in particular brain regions were thought to play a role in the development of Parkinson disease, which is the second most common neurodegenerative disorder, just behind Alzheimer [183], while DHEA administration showed some beneficial effect in a primate model of Parkinson disease [184]. Inverse relations between DHEAS levels in saliva [185] and circulation [180] and some domains of memory impairment were also documented, supporting the hypothesis that DHEA supplementation may improve cognition in the elderly; yet solid evidence of associations between the endogenous levels of these steroids and measures of cognitive function is lacking.

No studies with DHEA replacement, either acute administration or chronic (up to 12 months) supplementation, have shown a benefit in cognitive function in healthy elderly populations [75,141, 185-189]. Furthermore, DHEA supplementation failed to show any benefit in patients with Alzheimer disease [190] and had only minimal beneficial effect on specific cognitive domains such as the verbal fluency in older women with mild to moderate cognitive impairment [191]. Remarkably, some other studies have actually shown a negative effect of DHEAS replacement on cognitive performance [186,192,193]. It should be noted however, that most studies included only small groups of patients and were up to a yearl ong, which is probably not enough time to address the potential role of DHEA / DHEAS in neurodegenerative disorders.

3f. DHEA and Metabolism

Lipids

In women, the effects of sex steroids on lipid profile differ according to the steroid treatment (estrogen or androgen) and to the route of administration. Thus oral methyltestosterone lowers high-density lipoprotein (HDL)-cholesterol [194], and oral estrogen increases HDL-cholesterol and triglycerides and lowers low-density lipoprotein (LDL)-cholesterol and total cholesterol [195,196], while transdermal estradiol and transdermal testosterone have little or no effect on lipids [194,197]. Combined oral estrogen and methyltestosterone is associated with lowering of HDL-cholesterol [198,199]. Considering that DHEA can be converted intracellularly to estrogens and androgens, the effect on the lipid profile could be mixed and may vary between individuals. Most of the recent well designed studies, addressing this issue report no association or even an adverse association (at least in women) [200,201] between plasma levels of DHEA [202,203] or DHEA administration [104,107,196,204,205] and lipid profile.

Body mass index (BMI)

Animal studies support a beneficial effect of DHEA administration on obesity [206-208]. In humans, two sets of longitudinal analyses of studies with women in menopausal transition showed that elevated DHEAS level is negatively related to BMI [30,31]. On the other hand, baseline analyses by Santoro et al [209] did not find much association between DHEAS and BMI, waist-hip ratio, or waist. Similarly, a 2-year, placebo-controlled, randomized, double-blind study involving elderly men and women with low levels of DHEAS, showed no significant effect of DHEA replacement (75 mg per day, per os) on body composition measurements [73]. Interestingly, a, recent meta-analysis of intervention studies showed that DHEA supplementation in elderly men can induce only a small positive effect on body composition which is strictly dependent on DHEA conversion into its bioactive metabolites such as androgens or estrogens [210]. Putting together these results, current data regarding DHEA effect on BMI contradict each other, and its usage in clinical practice for body weight management is not suggested or recommended at the present.

Insulin Resistance

DHEA may at least theoretically improve endothelial function [41], and ameliorate local/systemic inflammation [49,50] and oxidative stress [58-61]. These effects in association with DHEAS’s inverse relation with body mass index (BMI) [30,31,210] would most probably suggest beneficial effect of DHEA/DHEAS supplement on insulin sensitivity [211]. This hypothesis was confirmed by reports from animal studies in which DHEA replacement had a beneficial effect on insulin sensitivity [207, 208]. In humans studies, however, the results are rather inconsistent. In some studies, the lower levels of DHEA seen with aging have been associated with impaired glucose tolerance, insulin resistance and diabetes [212-215], while in another [216] exactly the opposite relation was shown as higher levels of DHEA were associated with impaired glucose tolerance and diabetes mellitus in post-menopausal women. The truth regarding DHEA/DHEAS and insulin resistance and its associated conditions gets even more complicated considering conflicting results from interventional studies with DHEA replacement. Thus, an ameliorating effect of long-term treatment with DHEA on insulin resistance was described in a group of middle-aged hypo-adrenal women treated with DHEA [217], but also in groups of elderly men [97] and postmenopausal women [97,217-220] replaced with DHEA. The DHEA dose used ranged between 25 and 100 mg/day per os and the duration of treatment varied between 3 and 12 months; in one study transdermal DHEA was used [220]. Most other interventional studies addressing this issue, failed to demonstrate any significant effect of DHEA on insulin resistance/sensitivity [73,94,105,108,111,112,143,205,221-224] and so did a recent review of the available data regarding use of DHEA in women with adrenal insufficiency [77]. Remarkably, Mortola and Yen [140] reported worsening insulin resistance with DHEA replacement in postmenopausal women; in this study however, the number of participants was small (n=6), the duration of treatment short (28 days), and the DHEA dosage supraphysiological (1600 mg/day per os). Puting together the above, the relation between DHEA and carbohydrate metabolism is still uncertain.

3g. DHEA & cardiovascular disease (CVD)

CVD represents a serious public health problem; its prevalence increases with advancing age [225]. Low androgen levels have been related to atherogenic profile in men [226,227], while data from acute coronary units have shown transient fall of the testosterone levels in the first 24 hours after myocardial infarction (MI) [228,229], which probably deprives these patients of testosterone’s pro-fibrinolytic activity [230-232] and may actually result in increased 30-day mortality rate following acute MI [233]; the above findings suggest a strong relation between sex steroid hormones and CVD morbidity and mortality. Many studies have previously documented significant inverse relations between low DHEA/DHEAS levels and key elements involved in the development of atherosclerosis and CVD, including carotid artery intima-media thickness (IMT) [234,235], oxidative stress [58-61] and endothelial dysfunction [236], independent of other coronary risk factors. Low DHEAS levels were also predictive of severe coronary atherosclerosis on coronary angiography [237], but also of earlier cardiac allograft vasculopathy development in heart transplant patients [238].

These findings are suggestive of anti-atherogenic and cardioprotective effect of DHEA/DHEAS. Numerous epidemiological studies have, therefore, looked at the specific relation between plasma levels of adrenal androgens and CVD. Most have shown that low plasma levels of DHEA/DHEA-S were clearly associated with increased incidence of atherosclerotic vascular diseases [239-243] and cardiovascular morbidity [241,244-250], independently from classic cardiovascular risk factors, as well as with increased CVD-related mortality in elderly men [18,71,251-253] but not in postmenopausal women [18,246,251-253], unless they had pre-existing coronary disease [203].

The plasma levels of DHEA were also inversely associated with the progression [254] and prognosis of heart failure [255], at least in men. The exact pathophysiologic background is still more or less unclear. Some preliminary data in patients with type 2 diabetes mellitus suggest that the adrenal androgens may increase the generation of activated protein C, an important anticoagulant protein that protects from acute coronary events [235]. Furthermore, DHEA may directly stimulate eNOS phosphorylation/activation in endothelial cells and NO production [38,256,257], which in turn induces vasodilation, and preserves myocardial perfusion [258]. DHEA may also exert anti-inflammatory actions [38,259], through which it may alleviate endothelial dysfunction, atherogenesis [260], and the acute thrombotic complications of atheroma [38,259,261-264] enhanced by systemic inflammation. The protective effects of DHEA on endothelium were also shown in several in vitro studies in which DHEA increased endothelial proliferation [41] and protected endothelial cells against apoptosis [59, 265]. Finally, DHEA can alleviate oxidative stress and inflammation in vascular smooth muscle cells (VSMCs) via ERK1/2 and NF-κB signaling pathways, although ig has no effect on their phenotype transition [266].

Other studies, however, have failed to show a significant relation between DHEA/DHEA-S and CVD. In men for example, myocardial infarction occurrence was not altered by DHEA-S levels [267,268],and acute myocardial infarctions were seen in patients with either low [269] or high [270] DHEA-S levels. Similarly in women, lower DHEA-S levels in ischemic heart disease patients versus control were observed in some studies [242,271] but not in others [272]. The reasons that account for the discrepancies among the above studies are not clear. It can be argued that smoking could be a possible confounding variable for both DHEA-S levels and CVD, as
smoking increases DHEA-S levels but also increases the incidence of adverse cardiovascular events [273,274]. The discrepancies among the above studies may also be attributed to
population variability; for example, in the study by Mazat et al. the relative risk of a 8-year mortality associated with low DHEA-S was 3.4 times higher in males under 70 years
compared to older men (odds ratio of 6.5 versus 1.9) [18]. Finally, DHEA-S was checked just once in some retrospective studies, often several years before the adverse cardiovascular events [275].

Whether exogenously administered DHEA could ameliorate key elements involved in the generation and progression of the atherosclerotic process was addressed in humans with atherosclerosis and experimental animal models. The human studies have shown a beneficial effect of DHEA on angiographic evidence of atherosclerosis [241] and improvement of vascular endothelial function [41,276]. Several animal studies have also clearly demonstrated the inhibitory effect of orally administered DHEA on atherosclerosis and plaque progression [277,278] as well as beneficial effects on ischemia–reperfusion injury in the microcirculation [279,280] and cardiac dysfunction [55,56]. Arterial stiffness, which is also considered a risk factor for CVD [281], significantly improved after DHEA replacement in both elderly men and women [282]. Whether the above findings could be translated into DHEA administration in clinical practice for the reduction of CVD morbidity and/or mortality is definitely not well documented and supported by current reports. However, since DHEA is a well-tolerated molecule and an inexpensive drug, additional large multi-centric clinical studies could address its role in the prevention and/or management of CVD.

3h. DHEA & Cerebrovascular disease

Stroke is the third-leading cause for disability worldwide [283]; therefore, early risk stratification for an optimized allocation of health care resources is imperative. The ischemic strokes that account for the great majotiry of all stroke cases (87 percent) occur as a result of acute obstruction of atherosclerotic blood vessels supplying blood to the brain [284]. Considering DHEAS has neuroprotective and antiatherosclerotic properties [248,285,289] and its synthesis has been documented in the brain [179,290,291], the role of DHEA/DHEAS in acute stroke incidence and outcome was investigated. Interestingly, in women from Nurses’ Health Study, lower DHEAS levels were associated with a greater risk of ischemic stroke [292]. In addition, it was suggested that DHEAS circulation levels may actually predict the severity and functional outcome of acute strokes [293,294]. Whether the above findings suggest baseline DHEAS levels could alter stroke management in clinical practice or whether DHEA replacement has a therapeutic potential role in stroke management need to be addressed.

3i. DHEA and pulmonary hypertension

The previously described vasorelaxant properties of DHEA in systemic circulation were also investigated in pulmonary hypertension in animal models and also in humans. Several studies have shown that DHEA replacement could effectively prevent and also reverse hypoxic pulmonary hypertension, pulmonary arterial remodeling, and right ventricular hypertrophy in rats [295-297] in a dose-dependent manner [298] and also prevent the age-related frailty induced by hypoxic pulmonary hypertension in older mice [299]. The effect of DHEA is selective to the pulmonary circulation since the systemic blood pressure was not altered. It was shown that the beneficial effects of DHEA on pulmonary hypertension were at least partly independent of its conversion to estrogen/testosterone and eNOS activation. Some of the potential molecular mechanism by which DHEA promotes pulmonary artery relaxation appear to involve K+ channel activation, upregulation of soluble guanylate cyclase [296,300,301], downregulation of hypoxia inducible factor 1a (HIF-1a) [302] and by NADPH oxidation-elicited subunit dimerization of protein kinase G 1α [303].

 

As previously discussed, DHEA may inhibit and reverse chronic hypoxia-induced pulmonary hypertension in rats. Little is known, however, about the effects of DHEA on the pulmonary circulation in humans. The levels of DHEA/DHEA-S in patients with pulmonary hypertension over time have not been determined, but the recent Multi-Ethnic Study of Atherosclerosis (MESA) - Right Ventricle  (RV) Study found that higher DHEA levels were associated with increased RV mass and stroke volume in women [304]. Another prospective study suggested a strong inverse correlation between natural DHEA/DHEA-S blood levels and the ten-year mortality in old male smokers and ex-smokers [305]. Prompted by the experimental findings in the pulmonary circulation, a recent study investigated whether DHEA can improve the clinical and hemodynamic status of patients with pulmonary hypertension associated to chronic obstructive pulmonary disease; eight patients with the disease were treated with DHEA (200mg daily orally) for 3 months. The results were very promising as DHEA treatment significantly improved the pulmonary hemodynamics and the physical performance of the patients, without worsening gas exchange, as do other pharmacological treatments of pulmonary hypertension [306].

Putting together the above evidence, there are to date experimental data to support the beneficial role of DHEA treatment in models with pulmonary hypertension, but only only few studies supporting its beneficial effect in patients with pulmonary hypertension associated with chronic obstructive pulmonary disease. Further clinical studies would probably clarify its therapeutic role in the management of pulmonary hypertension in clinical practice.

3j. DHEA and autoimmune disorders

Inflammatory bowel disease (IBD)

DHEA has anti-inflammatory properties [49,50]. Its levels appear to be low in people with ulcerative colitis and Crohn disease, irrespective of patient’s age [307,308]. A phase II small pilot trial in patients with active inflammatory bowel disease refractory to other drugs, treated with 200 mg dehydroepiandrosterone per day orally for 56 days [309] showed that DHEA may decrease the clinical activity of the disease and may even cause a remission. More studies are needed before saying for sure whether DHEA helps IBD or not.

Systemic Lupus Erythematosus (SLE)

SLE is another autoimmune disorder. A number of randomized controlled clinical studies have reported that regardless of patient’s age, taking DHEA (50-200mg/day) along with other medications improves quality of life for people with mild to moderate SLE, decreases corticosteroid requirements and reduces the frequency of flare-ups [310], though it probably does not change the overall course of their disease [311-316]. A study had actually shown DHEA replacement may increase bone mass in women with lupus [314]. A 2007 report in the Cochrane Database of Systematic Reviews [79] suggests a "modest but clinically significant impact" of DHEA replacement on health-related quality of life in the short-term for people with SLE; the impact on disease activity was inconsistent. Long- term outcomes and safety remain unstudied.

Rheumatoid arthritis (RA)

DHEA levels have been found to be low in people with rheumatoid arthritis [317,318] and get even lower with glucocorticoid therapy that is often employed in RA [319]. Considering the well-demonstrated immune-suppressive activities exerted by the adrenal androgens and their derivatives [320-322], the utility of DHEA as potential therapy for management of male and female RA patients was studied. Preliminary data from animal studies showed benefits of DHEA treatment in collagen-induced arthritis [323-325]. However, in previous carefully controlled human clinical trials, DHEA treatment produced only modest benefits [326], probably with the exception of female-treated RA patients who benefit the most by DHEA replacement [327]. The noted discrepancy in benefits from DHEA treatment between animals and humans may be related to the little endogenous DHEA in rodents relative to humans because of low levels of cytochrome P450 17α-hydroxylase [179], but also because of different DHEA metabolism between species; remarkably, in rodents DHEA has many highly oxygenated metabolites and a surprisingly complex metabolism that results in production of a multitude highly oxygenated species that may exert the beneficial effects on arthritis [328].

 

  1. DHEA and Adverse Health Outcomes

DHEA supplements are generally well tolerated in studies using oral or percutaneous administration, with daily doses ranging from 25 mg to 1,600 mg. DHEA is an important precursor for estrogen and androgen production. In women DHEA when administered per os is mainly converted to androgen metabolites. As a result, some minimal androgenic adverse effects have been reported, including mild acne, seborrhea, facial hair growth, and ankle swelling [38,44,76].

A hormonal etiology has long been suspected for breast and endometrial cancer as several risk factors for each cancer, such as obesity, nulliparity, and early menarche are hormonally related [79,329-331]. The plasma concentrations of the adrenal androgens in premenopausal women were previously associated with higher risk for development of breast cancer [332-334]. Furthermore, DHEA-S levels above a cut off limit predicted disease progression in hypoestrogenised women treated for breast cancer [335]. On the other hand, in vitro studies support an inhibitory effect of DHEA on the growth of human mammary cancer cells and the growth of chemically-induced mammary cancer in rats [63,336,337]. It was shown that the effect of DHEA in mammary tissue depends on the level of plasma estrogens. Thus, growth inhibition occurs only in the presence of high estrogen concentrations, and growth stimulation occurs in the presence of a low-level estrogen milieu [338,339]. The exact role of DHEA supplementation on breast cancer in humans has not been fully studied. A previous review of clinical, epidemiological and experimental studies suggests late promotion of breast cancer in postmenopausal women by prolonged intake of DHEA, especially if central obesity coexists, and suggests extra caution when DHEA supplements are used by obese postmenopausal women [340]. A more recent review of the medical literature for key papers investigating DHEA physiology and randomized controlled trials of the use of DHEA in postmenopausal women, however, did not find any adverse effect of DHEA supplementation [133].

Unopposed oestrogen is also known to be associated with an increased risk of endometrial carcinoma [329]. DHEA supplementation did not increase the endometrial thickness in postmenopausal women treated with 25 mg/day oral DHEA for 6 months [341] or 50 mg daily for 12 months [342]. In addition, DHEA administered percutaneously for 12 months to postmenopausal women was shown to have an estrogenic effect on the vagina without affecting the endometrium that remained atrophic [343].

In men, DHEA supplements are mainly transformed to estrogen metabolites but also to more potent androgens. As a result, concerns regarding effect of DHEA supplementation on prostate were raised, especially after the finding that about 15% of DHT present in the prostate comes from DHEA metabolism [344]. A 2-year, placebo-controlled, randomized, double blind study involving elderly men receiving DHEA did not show any adverse effects in prostate [73].

As long as long-term safety data for DHEA supplementation are lacking, the American Cancer Society advices caution in its use in people who have cancer, especially types of cancer that respond to hormones, such as certain types of breast cancer, prostate cancer, and endometrial cancer [310].

  1. Conclusion

Theoretically, supplementing a pre-hormone is extremely interesting as it would provide peripheral tissues with levels of sex steroids according to local needs and would eliminate the exposure of other tissues to androgens or estrogens, minimizing unwanted side effects. Therefore, DHEA administration is closer to ‘‘hormonal optimization’’ than hormonal supplementation. In older people, lower than normal levels of DHEA/DHEAS were previously related to ageing-associated degenerative disorders, including metabolic and cardiovascular diseases, poor physical performance, mood and memory defects, sexual dysfunction and poor sense of wellbeing. Whether this is just a statistical finding with no practical clinical meaning has been investigated by many interventional studies most of which, however, were of short duration and had small number of participants. Without exception, all recent reviews of the available data regarding DHEA replacement utility for the management of ageing-related disorders do not support its usage in clinical practice [73-79]; no significant adverse or negative side effects of DHEA were reported in clinical studies, but also no significant evidence that low levels of DHEA cause the ageing-related degenerative disorders or that taking DHEA can help prevent/treat them. Thus, current clinical modalities with DHEA supplements do not comply with evidence-based medicine. Since there are several known biochemical actions by which DHEA could ameliorate disorders affecting the elderly population and is a well-tolerated molecule and an inexpensive drug, additional large multi-centric clinical studies would probably give us a better understanding of its clinical utility in the management of ageing-related disorders. Till then, we should probably reconsider suggesting patients to start on a pro-hormone that would help them only as much as a placebo would help.

References

1. Auchus RJ. Overview of dehydroepiandrosterone biosynthesis. Semin Reprod Med 2004;22:281–8

2. Payne AH, Hales DB. Overview of steroidogenic enzymes in the pathway from cholesterol to active steroid hormones. Endocrine Rev 2004;25:947–70

3. Davis S. Androgens and female sexuality. J Gerontol B Psychol Sci Soc Sci 2000;3:32–40

4. Hemat RAS (2004). Principles Of Orthomolecularism. Urotext. p. 426

5. Labrie F. Extragonadal synthesis of sex steroids: intracrinology. Ann Endocrinol (Paris) 2003;64:95–107

6. Labrie F, Belanger A, Cusan L, Candas B. Physiological changes in dehydroepiandrosterone are not reflected by serum levels of active androgens and estrogens but of their metabolites: intracrinology. J Clin Endocrinol Metab 1997;82:2403–9

7. Labrie F. Intracrinology. Mol Cell Endocrinol 1991;78:C113–8

8. Labrie C, Belanger A, Labrie F. Androgenic activity of dehydroepiandrosterone and androstenedione in the rat ventral prostate. Endocrinology 1988;123:1412–7

9. Labrie F. Extragonadal synthesis of sex steroids: intracrinology. Ann Endocrinol (Paris) 2003;64:95–107

10. Labrie F. Adrenal androgens and intracrinology [see comment]. Semin Reprod Med 2004;22:299–309

11. Labrie F. DHEA as physiological replacement therapy at menopause [see comment]. J Endocrinol Invest 1998;21:399–401

12. Labrie F, Luu-The V, Labrie C, et al. Endocrine and intracrine sources of androgens in women: inhibition of breast cancer and other roles of androgens and their precursor dehydroepiandrosterone. Endocrine Rev 2003;24:152

13. Poulan R, Labrie F. Stimulation of cell growth by C-19 steroids of adrenal origin in the ZR-75-1 human breast cancer cell line. Cancer Res. 1985; 46:4933–4937

14. Seymour-Munn K, Adams JB. Estrogenic effects of 5-androstene-3beta, 17beta diol at physiological concentrations and its possible implications in the etiology of breast cancer.Endocrinol. 1981; 112:486–491

15. Labrie F, Belanger A, Cusan L, Gomez J-L, Candas B. Marked decline in serum concentrations of adrenal C19 sex steroid precursors and conjugated androgen metabolites during aging. J Clin Endocrinol Metab 1997;82:2396–402

16. Labrie F, Belanger A, Belanger P, et al. Androgen glucuronides, instead of testosterone, as the new markers of androgenic activity in women. J Steroid Biochem Mol Biol 2006;99:182–8

17. Baulieu EE, Thomas G, Legrain S, Lahlou N, Roger M, Debuire B, et al. Dehydroepiandrosterone (DHEA), DHEA sulfate, and aging: contribution of the DHEAge study to a sociobiomedical issue. Proc Natl Acad Sci U S A 2000;97:4279–84

18. Mazat L, Lafont S, Berr C, Debuire B, Tessier JF, Dartigues JF, et al. Prospective measurements of dehydroepiandrosterone sulfate in a cohort of elderly subjects: relationship to gender, subjective health, smoking habits, and 10-year mortality. Proc Natl Acad Sci U S A 2001;98:8145–50

19. Heaney JL, Phillips AC, Carroll D. Ageing, physical function, and the diurnal rhythms of cortisol and dehydroepiandrosterone.Psychoneuroendocrinology. 2012;37(3):341-9

20. Laughlin, G. A. and Barrett-Connor, E. Sexual dimorphism in the influence of advanced aging on adrenal hormone levels: the Rancho Bernardo Study. J Clin Endocrinol Metab 2000;85:3561-3568

 

21. Baulieu EE. Androgens and aging men. Mol Cell Endocrinol 2002;198:41–9

22. Parker, C.R. Dehydroepiandrosterone and dehydroepiandrosterone sulfate production in the human adrenal during development and aging. Steriods 1999;64, 640—647

 

23. Yen, S.S.C., Laughlin, G.A. Aging and the adrenal cortex. Exp. Gerontol. 1998;33, 897—910

24. Ahn, R.S., Lee, Y.J., Choi, J.Y., Kwon, H.B., Chun, S.I. Salivary cortisol and DHEA levels in the Korean population: age-related differences, diurnal rhythm, and correlations with serum levels. Yonsei Med. J. 2007;48, 379—38

25. Liu, C. H., Laughlin, G. A., Fischer, U. G., and Yen, S. S. Marked attenuation of ultradian and circadian rhythms of dehydroepiandrosterone in postmenopausal women: evidence for a reduced 17,20-desmolase enzymatic activity. J Clin Endocrinol Metab 1990;71:900-906

26. Orentreich, N., Brind, J. L., Rizer, R. L., and Vogelman, J. H. Age changes and sex differences in serum dehydroepiandrosterone sulfate concentrations throughout adulthood. J Clin Endocrinol Metab 1984; 59:551-555

 

27. Prom-Wormley EC, York TP, Jacobson KC, Eaves LJ, Mendoza SP, Hellhammer D, Maninger N, Levine S, Lupien S, Lyons MJ, Hauger R, Xian H, Franz CE, Kremen WSGenetic and environmental effects on diurnal dehydroepiandrosterone sulfate concentrations in middle-aged men. Psychoneuroendocrinology. 2011;36:1441-52

 

28. Lasley BL, Santoro N, Gold EB, et al. The relationship of circulating DHEAS, testosterone and estradiol to stages of the menopausal transition and ethnicity. J. Clin Endocrinol Metab. 2002; 87:3760–3767

29. World Health Organization Scientific Group. Geneva, Switzerland: World Health Organization;. Research on Menopause in the 1990s. WHO Technical Report Series 1996;no. 866: 4

30. Lasley B, Crawford S, Laughlin G, et al. Circulating Dehydroepiandrosterone Levels in Women with Bilateral Salpingo-Oophorectomy during the Menopausal Transition. Menopause. 2011; 18:494–498

31. Crawford S, Santoro N, Laughlin GA, et al. Circulating Dehydroepiandrosterone Sulfate Concentrations during the Menopausal Transition. J Clin Endocrinol Metab. 2009; 94:2945–2951

32. Randolph JF Jr, Sowers M, Gold EB, et al. Reproductive hormones in the early menopausal transition: relationship to ethnicity, body size, and menopausal status. J Clin Endocrinol Metab. 2003; 88:1516–1522

33. B. L. Lasley, S. Crawford, and D.S. McConnell.Adrenal Androgens and the Menopausal Transition. Obstet Gynecol Clin North Am. 2011 ; 38: 467–475

34. Parker, LN (1991). "Adrenarche". Endocrinology and metabolism clinics of North America 20 : 71–83

35. Liu D, Dillon JS. Dehydroepiandrosterone activates endothelial cell nitric-oxide synthase by a specific plasma membrane receptor coupled to Galpha(i2,3). J Biol Chem 2002;277:21379–21388

36. Liu D, Dillon JS. Dehydroepiandrosterone stimulates nitric oxide release in vascular endothelial cells: Evidence for a cell surface receptor. Steroids 2004;69:279–289

 

37. Simoncini T, Mannella P, Fornari L, Varone G, Caruso A, Genazzani AR. Dehydroepiandrosterone modulates endothelial nitric oxide synthesis via direct genomic and nongenomic mechanisms. Endocrinology 2003;144:3449–3455

 

38. Traish AM, Kang HP, Saad F, Guay AT. Dehydroepiandrosterone (DHEA) a precursor steroid or an active hormone in human physiology. J Sex Med 2011;8:2960–2982

39. Chevalier M, Gilbert G, Lory P, Marthan R, Quignard JF, Savineau JP. Dehydroepiandrosterone; (DHEA) inhibits voltage-gated T-type calcium channels. Biochem Pharmacol 2012;83:1530–9

40. Bonnet S, Paulin R, Sutendra G, Dromparis P, Roy M, Watson KO, et al. Dehydroepiandrosterone reverses systemic vascular remodeling through the inhibition of the Akt/GSK3-b/NFAT axis. Circulation 2009;120:1231–40

41. Williams MR, Dawood T, Ling S, Dai A, Lew R, Myles K, et al. Dehydroepiandrosterone increases endothelial cell proliferation in vitro and improves endothelial function in vivo by mechanisms independent of androgen and estrogen receptors. J Clin Endocrinol Metab 2004;89:4708–15

42. Williams MR, Ling S, Dawood T, Hashimura K, Dai A, Li H, et al. Dehydroepiandrosterone inhibits human vascular smooth muscle cell proliferation independent of ARs and ERs. J Clin Endocrinol Metab 2002;87:176–81

43. Levin ER. Integration of the extranuclear and nuclear actions of estrogen.Mol Endocrinol 2005;19:1951–9

 

44. Legrain S, Massien C, Lahlou N, Roger M, Debuire B, Diquet B, Chatellier G, Azizi M, Faucounau V, Porchet H, Forette F, Baulieu EE. Dehydroepiandrosterone replacement administration: Pharmacokinetic and pharmacodynamic studies in healthy elderly subjects. J Clin Endocrinol Metab 2000;85:3208–3217

 

45. Goodwin B, Redinbo MR, Kliewer SA. Regulation of cyp3a gene transcription by the pregnane X receptor. Annu Rev Pharmacol Toxicol 2002;42:1–23

46. Kliewer SA, Goodwin B, Willson TM. The nuclear pregnane X receptor: a key regulator of xenobiotic metabolism. Endocr Rev 2002;23:687–702

47. Moore LB, Parks DJ, Jones SA, Bledsoe RK, Consler TG, Stimmel JB, et al. Orphan nuclear receptors constitutive androstane receptor and pregnane X receptor share xenobiotic and steroid ligands. J Biol Chem 2000;275:15122–27

48. Ripp SL, Fitzpatrick JL, Peters JM, Prough RA. Induction of CYP3A expression by dehydroepiandrosterone: involvement of the pregnane X receptor. Drug Metab Dispos 2002;30:570–5

49. Straub RH, Konecna L, Hrach S, et al. Serum dehydroepiandrosterone (DHEA) and DHEA sulfate are negatively correlated with serum interleukin-6 (IL-6), and DHEA inhibits IL-6 secretion from mononuclear cells in man in vitro: possible link between endocrinosenescence and immunosenescence. J Clin Endocrinol Metab 1998; 83: 2012–7

50. Poynter ME, Daynes RA. Peroxisome proliferator-activated receptor alpha activation modulates cellular redox status, represses nuclear factor-kappaB signaling, and reduces inflammatory cytokine production in aging. J Biol Chem 1998; 273: 32 833–41

51. Wang L, Wang YD, Wang WJ, Zhu Y, Li DJ. Dehydroepiandrosterone improves murine osteoblast growth and bone tissue morphometry via mitogen-activated protein kinase signaling pathway independent of either androgen receptor or estrogen receptor. J Mol Endocrinol 2007;38:467–479

52. Maurice T, Gregoire C, Espallergues J. Neuro(active)steroids actions at the neuromodulatory sigma1 (sigma1) receptor: biochemical and physiological evidences, consequences in neuroprotection. Pharmacol Biochem Behav 2006;84:581–97

 

53. Cheng ZX, Lan DM, Wu PY, Zhu
YH, Dong Y, Ma L, et al. Neurosteroid dehydroepiandrosterone sulphate inhibits persistent sodium currents in rat medial prefrontal cortex via activation of sigma-1 receptors. Exp Neurol 2008;210:128–36

54. Tsai SY, Hayashi T, Mori T, Su TP. Sigma-1 receptor chaperones and diseases. Cent Nerv Syst Agents Med Chem 2009;9:184–9

 

55. Tagashira H, Bhuiyan S, Shioda N, Fukunaga K. Distinct cardioprotective effects of 17beta-estradiol and dehydroepiandrosterone on pressure overload-induced hypertrophy in ovariectomized female rats. Menopause 2011;18:1317–26

56. Bhuiyan MS, Tagashira H, Fukunaga K. Dehydroepiandrosterone-mediated stimulation of sigma-1 receptor activates Akt–eNOS signaling in the thoracic aorta of ovariectomized rats with abdominal aortic banding. Cardiovasc Ther 2011;29:219–30

57. Bhuiyan S, Fukunaga K. Stimulation of Sigma-1 receptor by dehydroepiandrosterone ameliorates hypertension-induced kidney hypertrophy in ovariectomized rats. Exp Biol Med (Maywood) 2010;235:356–64

58. Camporez JP, Akamine EH, Davel AP, Franci CR, Rossoni LV, Carvalho CR. Dehydroepiandrosterone protects against oxidative stress-induced endothelial dysfunction in ovariectomized rats. J Physiol 2011;589:2585–96

59. Nheu L, Nazareth L, Xu GY, Xiao FY, Luo RZ, Komesaroff P, et al. Physiological effects of androgens on human vascular endothelial and smooth muscle cells in culture. Steroids 2011;76:1590–6

60. Jacob MH, Janner Dda R, Bello-Klein A, Llesuy SF, Ribeiro MF. Dehydroepiandrosterone modulates antioxidant enzymes and Akt signaling in healthy Wistar rat hearts. J Steroid Biochem Mol Biol 2008;112:138–44

61. Jia C, Chen X, Li X, Li M, Miao C, Sun B, et al. The effect of DHEA treatment on the oxidative stress and myocardial fibrosis induced by Keshan disease pathogenic factors. J Trace Elem Med Biol 2011;25:154–9

62. Levy HR. Glucose-6-phosphate dehydrogenases. Adv Enzymol Relat Areas Mol Biol 1979;48:97–192

63. Schwartz AG, Pashko LL. Dehydroepiandrosterone, glucose-6-phosphate dehydrogenase, and longevity. Ageing Res Rev 2004;3:171–87

64. Forti P, Maltoni B, Olivelli V, Pirazzoli G.L., Ravaglia G, Zoli M. Serum Dehydroepiandrosterone Sulfate and Adverse Health Outcomes in Older Men and Women . Rejuvenation Research 2012; 15:349-58

65. Valenti G, Denti L, Maggio M, Ceda G, Volpato S, Bandinelli S, Ceresini G, Cappola A, Guralnik JM, Ferrucci L. Effect of DHEAS on skeletal muscle over the life span: The InCHIANTI study. J Gerontol Ser A, Biol Sci Med Sci 2004;59:466–472

66. Kostka T, Arsac LM, Patricot MC, Berthouze SE, Lacour JR, Bonnefoy M. Leg extensor power and dehydroepiandrosterone sulfate, insulin-like growth factor-I and testosterone in healthy active elderly people. Eur J Appl Physiol 2000;82:83–90

67. Ravaglia G, Forti P, Maioli F, Boschi F, Cicognani A, Bernardi M, Pratelli L, Pizzoferrato A, Porcu S, Gasbarrini G. Determinants of functional status in healthy Italian nonagenarians and centenarians: A comprehensive functional assessment by the instruments of geriatric practice. J Am Geriatr Soc 1997;45:1196–1202

68. Leng SX, Cappola AR, Anderson RE et al. Serum levels of insulin-like growth factor-I (IGF-I) and dehydroepiandrosterone sulfate (DHEA-S), and their relationships with serum interleukin-6, in the geriatric syndrome of frailty. Aging Clin Exp Res 2004;16:153–157

69. Voznesensky M, Walsh S, Dauser D et al. The association between dehydroepiandosterone and frailty in older men and women. Age Ageing 2009;38:401–406

70. Baulieu E, Thomas G, Legrain S, et al. Dehydroepiandrosterone (DHEA), DHEA sulfate, and aging: contribution of the DHEAge study to a sociobiomedical issue. Proc Nat Acad Natl Sci 2000;97:4279–84

71. Ohlsson C, Labrie F, Barrett-Connor E, Karlsson MK, Ljunggren O, et al. Low serum levels of dehydroepiandrosterone sulfate predict all-cause and cardiovascular mortality in elderly Swedish men. J Clin Endocrinol Metab 2010; 95: 4406–4414

72. Kroboth PD, Salek FS, Pittenger AL, Fabian TJ, Frye RF (1999) DHEA and DHEA-S: a review. J Clin Pharmacol 39: 327–348

 

73. Nair KS, Rizza RA, O’Brien P, et al. DHEA in elderly women and DHEA or testosterone in elderly men. N Engl J Med 2006;355:1647–1659

 

74. William L. Baker, Shobha Karan, and Anne M. Kenny Effect of Dehydroepiandrosterone on Muscle Strength and Physical Function in Older Adults: A Systematic Review. J Am Geriatr Soc 2011;59:997–1002

75. Kritz-Silverstein D, von Mühlen D, Gail A. Laughlin GA, Bettencourt R. Effects of dehydroepiandrosterone supplementation on cognitive function and quality of life: the DHEA and well-ness (DAWN) trial. J Am Geriat Soc 2008;56:1292–8

76. Panjari M, Bell RJ, Jane F, et al..A randomized trial of oral DHEA treatment for sexual function, well-being, and menopausal symptoms in postmenopausal women with low libido. J Sex Med 2009; 6: 2579–2590

77. Alkatib AA, Cosma M, Elamin MB, et al. A systematic review and metaanalysis of randomized placebo-controlled trials of DHEA treatment effects on quality of life in women with adrenal insufficiency.J Clin Endocrinol Metab 2009;94:3676–81

 

78. Labrie F, Archer D, Bouchard C, et al. Effect of intravaginal dehydroepiandrosterone (Prasterone) on libido and sexual dysfunction in postmenopausal women. Menopause (New York, NY) 2009;16:923–31

 

79. van
den Brandt PA
Spiegelman DYaun
SS
Adami
HO
Beeson
L
Folsom
AR
Fraser
G
Goldbohm
RA
,Graham
S
Kushi
L
Marshall
JR
Miller
AB
Rohan
T
Smith-Warner
SA
Speizer
FE
Willett
WC
Wolk
A
,Hunter
DJ
. Pooled analysis of prospective cohort studies on height, weight, and breast cancer risk. Am J Epidemiol. 2000 ;152:514-27.

 

80. Arlt W. The approach to the adult with newly diagnosed adrenal insufficiency. J Clin Endocrinol Metab 2009;94:1059–1067

81. Panjari M, Davis SR. DHEA for postmenopausal women: A review of the evidence. Maturitas 2010;66:172–179

82. Labrie F, Belanger A, Belanger P, et al. Metabolism of DHEA in postmenopausal women following percutaneous administration. J Steroid Biochem Mol Biol 2007;103:178–88

83. Labrie F, Cusan L, Gomez JL, Martel C, Be´rube´ R, Be´langer P, Chaussade V, Deloche C, Leclaire J. Changes in serum DHEA and eleven of its metabolites during 12-month percutaneous administration of DHEA. J Steroid Biochem Mol Biol 2008;110:1–9

84. Guideline for the prevention of falls in older persons. American Geriatrics Society, British Geriatrics Society, and American Academy of Orthopaedic Surgeons Panel on Falls Prevention. J Am Geriatr Soc 2001;49:664–672

85. Janssen I, Ross R. Linking age-related changes in skeletal muscle mass and composition with metabolism and disease. J Nutr Health Aging 2005;9:408–419

86. Berkman LF, Seeman TE, Albert M et al. High, usual and impaired functioning in community-dwelling older men and women: Findings from the MacArthur Foundation Research Network on Successful Aging. J Clin Epidemiol 1993;46:1129–1140

 

87. Morrison MF, Katz IR, Parmelee P et al. Dehydroepiandrosterone sulfate (DHEA-S) and psychiatric and laboratory measures of frailty in a residential care population. Am J Geriatr Psychiatry 1998;6:277–284

 

88. Abbasi A, Duthie EH, Sheldahl L et al. Association of dehydroepiandrosterone sulfate, body composition, and physical fitness in independent communitydwelling older men and women. J Am Geriatr Soc 1998;46:263–273

89. Clarke BL, Ebeling PR, Jones JD, Wahner HW, O’Fallon WM, Riggs BL, Fitzpatrick LA. Predictors of bone mineraldensity in aging healthy men varies by skeletal site. Calcified Tiss Int 2002;70:137–145

90. Nawata H, Tanaka S, Tanaka S, Takayanagi R, Sakai Y, Yanase T, Ikuyama S, Haji M. Aromatase in bone cell: Association with osteoporosis in postmenopausal women. J Steroid Biochem Mol Biol 1995;53:165–174

91. Bischoff-Ferrari HA, Orav EJ, Dawson-Hughes B. Additive benefit of higher testosterone levels and vitamin D plus calcium supplementation in regard to fall risk reduction among older men and women. Osteoporos Int 2008;19:1307–1314

92. Hasselhorn HM, Theorell T, Vingard E. Endocrine and immunologic parameters indicative of 6-month prognosis after the onset of low back pain or neck/shoulder pain. Spine (Phila Pa 1976) 2001;26: E24–29

93. Schell E, Theorell T, Hasson D, Arnetz B, Saraste H. Stress biomarkers’ associations to pain in the neck, shoulder and back in healthy media workers: 12-month prospective follow-up. Eur Spine J 2008; 17: 393–405

94. Morales AJ, Haubrich RH, Hwang JY, Asakura H, Yen SS. The effect of six months treatment with a 100 mg daily dose of dehydroepiandrosterone (DHEA) on circulating sex steroids, body composition and muscle strength in age-advanced men and women. Clin Endocrinol 1998;49: 421–432

95. Villareal DT, Holloszy JO, Kohrt WM. Effects of DHEA replacement on bone mineral density and body composition in elderly women and men. Clin Endocrinol 2000;53:561–568

96. Kalman DS CC, Swain MA, Torina GC, Shi Q. A randomized, double-blind, placebo-controlled study of 3-acetyl-7-oxo-dehydroepiandrosterone in healthy over- weight adults. Curr Ther Res 2000;61:435–442

97. Villareal DT, Holloszy JO. Effect of DHEA on abdominal fat and insulin action in elderly women and men: A randomized controlled trial. JAMA 2004;292:2243–2248

98. Kenny AM, Boxer RS, Kleppinger A, Brindisi J, Feinn R, Burleson JA. Dehydroepiandrosterone combined with exercise improves muscle strength and physical function in frail older women. J Am Geriatr Soc 2010;58:1707–1714

99. Weiss EP, Shah K, Fontana L, Lambert CP, Holloszy JO, Villareal DT. Dehydroepiandrosterone replacement therapy in older adults: 1- and 2-y effects on bone. Am J Clin Nutr 2009;89:1459–1467

100. von Muhlen D, Laughlin GA, Kritz-Silverstein D, Bergstrom J, Bettencourt R. Effect of dehydroepiandrosterone supplementation on bone mineral density, bone markers, and body composition in older adults: The DAWN trial. Osteoporosis Int 2008;19:699–707

101. Jankowski CM, Gozansky WS, Kittelson JM, Van Pelt RE, Schwartz RS, Kohrt WM. Increases in bone mineral density in response to oral dehydroepiandrosterone replacement in older adults appear to be mediated by serum estrogens. J Clin Endocrinol Metab 2008;93:4767–4773

102. Jankowski CM, Gozansky WS, Schwartz RS, Dahl DJ, Kittelson JM, Scott SM, Van Pelt RE, Kohrt WM. Effects of dehydroepiandrosterone replacement therapy on bone mineral density in older adults: A randomized, controlled trial. J Clin Endocrinol Metab 2006;91:2986–2993

103. Labrie F, Diamond P, Cusan L, Gomez JL, Belanger A, Candas B. Effect of 12-month dehydroepiandrosterone replacement therapy on bone, vagina, and endometrium in postmenopausal women. J Clin Endocrinol Metab 1997;82: 3498–3505

104. Morales AJ, Haubrich RH, Hwang JY, Asakura H, Yen SS. The effect of six months treatment with a 100 mg daily dose of dehydroepiandrosterone (DHEA) on circulating sex steroids, body composition and muscle strength in ageadvanced men and women.Clin Endocrinol 1998;49: 421–432

 

105. Morales AJ, Nolan JJ, Nelson JC, Yen SS. Effects of replacement dose of dehydroepiandrosterone in men and women of advancing age. J Clin Endocrinol Metab 1994; 78:1360–1367

 

106. Yen SS, Morales AJ, Khorram O. Replacement of DHEA in aging men and women. Potential remedial effects. Ann NY Acad Sci 1995;774:128–142

107. Jedrzejuk D, Medras M, Milewicz A, Demissie M. Dehydroepiandrosterone replacement in healthy men with age-related decline of DHEA-S: Effects on fat distribution, insulin sensitivity and lipid metabolism. Aging Male 2003; 6:151–156

108. Callies F, Fassnacht M, van Vlijmen JC, Koehler I, Huebler D, Seibel MJ, Arlt W, Allolio B. Dehydroepiandrosterone replacement in women with adrenal insufficiency: Effects on body composition, serum leptin, bone turnover, and exercise capacity. J Clin Endocrinol Metab 2001;86:1968– 1972

109. Flynn MA, Weaver-Osterholtz D, Sharpe-Timms KL, Allen S, Krause G. Dehydroepiandrosterone replacement in aging humans. J Clin Endocrinol Metab 1999;84:1527–1533

110. Percheron G, Hogrel JY, Denot-Ledunois S, Fayet G, Forette F, Baulieu EE, Fardeau M, Marini JF. Effect of 1-year oral administration of dehydroepiandrosterone to 60- to 80-year-old individuals on muscle function and cross-sectional area: A double-blind placebo-controlled trial. Arch Int Med 2003;163:720–727

111. Igwebuike A, Irving BA, Bigelow ML, Short KR, McConnell JP, Nair KS. Lack of dehydroepiandrosterone effect on a combined endurance and resistance exercise program in postmenopausal women. J Clin Endocrinol Metab 2008;93:534–538

112.Casson PR , Santoro N , Elkind,Hirsch K , Carson SA , Hornsby PJ , Abraham G , Buster JE. Postmenopausal dehydroepiandrosterone administration increases free insulin-like growth factor-I and decreases high-density lipoprotein: a six-month trial. Fertil Steril 1998;70:107–110

113. Kenny AM, Boxer RS, Kleppinger A et al. Dehydroepiandrosterone (DHEA) improves muscle strength and physical function but not bone mineral density in frail older women. J Am Geriatr Soc 2010;58:1707–1714

114. Muller M, van den Beld AW, van der Schouw YT et al. Effects of dehydroepiandrosterone and atemestane supplementation on frailty in elderly men. J Clin Endocrinol Metab 2006;91:3988–3991

115. Villareal DT, Holloszy JO. DHEA enhances effects of weight training on muscle mass and strength in elderly women and men. Am J Physiol Endocrinol Metab 2006;291:E1003–E1008

116. Percheron G, Hogrel JY, Denot-Ledunois S et al. Effect of 1-year oral administration of dehydroepiandrosterone to 60- to 80-year-old individuals on muscle function and cross-sectional area. A double-blind placebo-controlled trial. Arch Intern Med 2003;163:720–727

117. Arlt W, Callies F, Koehler I et al. Dehydroepiandrosterone supplementation inhealthy men with an age-related decline of dehydroepiandrosterone secretion. J Clin Endocrinol Metab 2001;86:4686–4692

118. Nijland E, Davis S, Laan E, Schultz WW. Female sexual satisfaction and pharmaceutical intervention: a critical review of the drug intervention studies in female sexual dysfunction. J Sex Med 2006;3:763–77

119. Shifren JL, Monz BU, Russo PA, Segreti A, Johannes CB. Sexual problems and distress in United States women: prevalence and correlates. Obstet Gynecol 2008;112:970–8

120. Hayes RD, Dennerstein L, Bennett CM, Koochaki PE, Leiblum SR, Graziottin A. Relationship between hypoactive sexual desire disorder and aging. Fertil Steril 2007;87:107–12

121. Genazzani AD, Lanzoni C, Genazzani AR. Might DHEA be considered a beneficial replacement therapy in the elderly? Drugs Aging 2007;24:173–185

122. Nappi RE, Polatti F. The use of estrogen therapy in women’s sexual functioning (CME). J Sex Med 2009;6:603–16

 

123. Bachmann GA, Leiblum SR. The impact of hormones on menopausal sexuality: a literature review. Menopause 2004;11:120–30

124. Alexander JL, Kotz K, Dennerstein L, Kutner SJ, Wallen K, Notelovitz M. The effects of postmenopausal hormone therapies on female sexual functioning: a review of double-blind, randomized controlled trials. Menopause 2004;11:749–65

125. Modelska K, Cummings S. Female sexual dysfunction in postmenopausal women: systematic review of placebo-controlled trials. Am J Obstet Gynecol 2003;188:286–93

126. Goldstat R, Briganti E, Tran J, Wolfe R, Davis S. Transdermal testosterone improves mood, well being and sexual function in premenopausal women. Menopause (New York, NY) 2003;10:390–8

127. Shifren JL, Braunstein G, Simon J, et al. Transdermal testosterone treatment in women with impaired sexual function after oophorectomy. N Eng J Med 2000;343:682

128. Braunstein GD, Sundwall DA, Katz M, et al. Safety and efficacy of a testosterone patch for the treatment of hypoactive sexual desire disorder in surgically menopausal women: a randomized, placebo-controlled trial. Arch Intern Med 2005;165:1582–9

129. Davis SR, van der Mooren MJ, van Lunsen RHW, et al. The efficacy and safety of a testosterone patch for the treatment of hypoactive sexual desire disorder in surgically menopausal women: a randomized, placebo-controlled trial. Menopause (New York, NY) 2006;13:387–96

130. Buster JE, Kingsberg SA, Aguirre O, et al. Testosterone patch for low sexual desire in surgically menopausal women: a randomized trial. Obstet Gynecol 2005;105:944

 

131. Davis S, Papalia MA, Norman RJ, et al. Safety and efficacy of a testosterone metered-dose transdermal spray for treating decreased sexual satisfaction in premenopausal women: a randomized trial. Ann Intern Med 2008;148:569–77

 

132. Davis SR, Davison SL, Donath S, Bell RJ. Circulating androgen levels and self-reported sexual function in women. JAMA 2005;294:91–96

133. Davis SR, Panjari M, Stanczyk FZ. Clinical review: DHEA replacement for postmenopausal women. J Clin Endocrinol Metab 2011;96:1642–1653

134. Schmidt PJ, Daly RC, Bloch M, et al. Dehydroepiandrosterone monotherapy in midlife-onset major and minor depression. Arch Gen Psychiatry 2005;62:154

135. Bloch M, Schmidt PJ, Danaceau MA, Adams LF, Rubinow DR. Dehydroepiandrosterone treatment of midlife dysthymia [see comment]. Biol Psychiatry 1999;45:1533–41

136. Hackbert L, Heiman JR. Acute dehydroepiandrosterone (DHEA) effects on sexual arousal in postmenopausal women. J Womens Health Gend Based Med 2002; 11: 155–162

137. Genazzani AR, Stomati M, Valentino V, Pluchino N, Pot E, Casarosa E, Merlini S, Giannini A, Luisi M. Effect of 1-year, low-dose DHEA therapy on climacteric symptoms and female sexuality. Climacteric 2011;14:661–668

138. Stomati M, Monteleone P, Casarosa E, Quirici B, Puccetti S, Bernardi F, Genazzani AD, Rovati L, Luisi M, Genazzani AR. Six-month oral dehydroepiandrosterone supplementation in early and late postmenopause. Gynecolog Endocrinol 2000;14:342–363

139. Panjari M, Davis SR. DHEA therapy for women: effect on sexual function and well being. Human Reprod Update 2007;13:239–48

140. Arlt W, Callies F, van Vlijmen JC, et al.. Dehydroepiandrosterone replacement in women with adrenal insufficiency. N Engl J Med 1999; 341: 1013–1020

141. Mortola JF, Yen SS.The effects of oral dehydroepiandrosterone on endocrine-metabolic parameters in postmenopausal women. J Clin Endocrinol Metab 1990; 71: 696–704

142. Wolf OT, Neumann O, Hellhammer DH, et al..Effects of a two-week physiological dehydroepiandrosterone substitution on cognitive performance and well-being in healthy elderly women and men. J Clin Endocrinol Metab 1997; 82: 2363–2367

143. Johannsson G, Burman P, Wiren L, et al. Low dose dehydroepiandrosterone affects behavior in hypopituitary androgen-deficient women: a placebocontrolled trial. J Clin Endocrinol Metab 2002;87:2046

144. van Thiel SW, Romijn JA, Pereira AM, et al. Effects of dehydroepiandrostenedione, superimposed on growth hormone substitution, on quality of life and insulin-like growth factor I in patients with secondary adrenal insufficiency: a randomized, placebo-controlled, cross-over trial. J Clin Endocrinol Metab 2005;90:3295–303

145. Lovas K, Gebre-Medhin G, Trovik T, et al. Replacement of dehydroepiandrosterone in adrenal failure: no benefit for subjective health status and sexuality in a 9-month randomized parallel group clinical trial. J Clin Endocrinol Metab 2003;88:1112–8

146. Avis NE, Brockwell S, Randolph Jr JF, et al. Longitudinal changes in sexual functioning as women transition through menopause: results from the Study of Women’s Health Across the Nation. Menopause 2009;16:442–52

147. Labrie F, Cusan L, Gomez JL, et al.Effect of intravaginal DHEA on serum DHEA and eleven of its metabolites in postmenopausal women. J Steroid Biochem Mol Biol 2008; 111: 178–194

148. Labrie F, Archer D, Bouchard C, et al.High internal consistency and efficacy of intravaginal DHEA for vaginal atrophy. Gynecol Endocrinol 2010; 26: 524–532

149. Ibe C, Simon JA. Vulvovaginal atrophy: current and future therapies (CME). J Sex Med 2010; 7: 1042–1050; quiz 51

150. Reiter WJ, Pycha A, Schatzl G, et al. Dehydroepiandrosterone in the treatment of erectile dysfunction: a prospective double-blind, randomized, placebo-controlled study. Urology 1999;53:590-595

151. Steffens DC. A multiplicity of approaches to characterize geriatric depression and its outcomes. Curr Opin Psychiatry 2009;22:522–526

152. Robichaud M, Debonnel G. Modulation of the firing activity of female dorsal raphe nucleus serotonergic neurons by neuroactive steroids. J Endocrinol 2004;182:11–21

 

153. Kilic
FS
1, Kulluk
D
Musmul
A
. Effects of dehydroepiandrosterone in amphetamine-induced schizophrenia models in mice.Neurosciences (Riyadh)  2014;19(2):100-5

 

154. Goodyer IM, Herbert J, Altham PM, Pearson J, Secher SM, et al. Adrenal secretion during major depression in 8- to 16-year-olds, I. Altered diurnal rhythms in salivary cortisol and dehydroepiandrosterone (DHEA) at presentation. Psychol Med 1996; 26: 245–256

155. Barrett-Connor E, von Muhlen D, Laughlin GA, Kripke A. Endogenous levels of dehydroepiandrosterone sulfate, but not other sex hormones, are associated with depressed mood in older women: The Rancho Bernardo Study. J Am Geriatr Soc 1999;47:685–691

156. Michael A, Jenaway A, Paykel ES, Herbert J. Altered salivary dehydroepiandrosterone levels in major depression in adults. Biol Psychiatry 2000;48:989–995

157. Morrison MF, Freeman EW, Lin H, and Sammel MD. Higher DHEA-S (Dehydroepiandrosterone Sulfate) Levels are Associated with Depressive Symptoms during the Menopausal Transition: Results from the PENN Ovarian Aging Study. Arch Womens Ment Health 2011; 14: 375–382

158. Takebayashi M, Kagaya A, Uchitomi Y, Kugaya A, Muraoka M, Yokota N, Horiguchi J, Yamawaki S. Plasma dehydroepiandrosterone sulfate in unipolar major depression. Short communication.J Neural Transm 1998; 105:537-42

159. Assies J, Visser I, Nicolson NA, Eggelte TA, Wekking EM, Huyser J, Lieverse R, Schene AH. Elevated salivary dehydroepiandrosterone-sulfate but normal cortisol levels in medicated depressed patients: preliminary findings. Psychiatry Res 2004 ; 128:117-22

160. Fabian TJ, Dew MA, Pollock BG, Reynolds CF 3rd, Mulsant BH, Butters MA, Zmuda MD, Linares AM, Trottini M, Kroboth PD. Endogenous concentrations of DHEA and DHEA-S decrease with remission of depression in older adults. Biol Psychiatry 2001; 50:767-74

161. Wolkowitz OM, Reus VI, Roberts E, Manfredi F, Chan T,RaumWJ,OrmistonS, JohnsonR, CanickJ, Brizendine L, Weingartner H. Dehydroepiandrosterone (DHEA) treatment of depression. Biol Psychiatry 1997;41:311–318

162. Wolkowitz OM, Reus VI, Keebler A, et al. Double-blind treatment of major depression with dehydroepiandrosterone. Am J Psychiatry 1999;156:646–649

163. Strous RD, Maayan R, Lapidus R, Stryjer R, Lustig M, Kotler M, Weizman A. Dehydroepiandrosterone augmentation in the management of negative, depressive, and anxiety symptoms in schizophrenia. Arch Gen Psychiatry 2003;60:133–141

164. Kaplan JR, Manuck SB, Clarkson TB et al. Social status, environment and atherosclerosis in cynomolgus monkeys. Arteriosclerosis 1982;2:359–368

165. Kivimaki M, Nyberg ST, Batty GD, Fransson EI, Heikkila K, et al. (2012) Job strain as a risk factor for coronary heart disease: a collaborative meta-analysis of individual participant data. Lancet 2012; 380: 1491–1497

166. MarmotM. Psychosocial factors and cardiovascular disease: epidemiological approaches. Eur. Heart J 1988;9:690–697

167. Epel ES, Blackburn EH, Lin J, Dhabhar FS, Adler NE, et al. Accelerated telomere shortening in response to life stress. Proc Natl Acad Sci USA 2004; 101: 17312–17315

168. Wolkowitz OM, Epel ES, Reus VI, Mellon SH. Depression gets old fast: do stress and depression accelerate cell aging? Depress Anxiety 2010; 27: 327–338

169. Jeckel CM, Lopes RP, Berleze MC, Luz C, Feix L, et al. Neuroendocrine and immunological correlates of chronic stress in ’strictly healthy’ populations. Neuroimmunomodulation 2010;17: 9–18

170. Izawa S, Saito K, Shirotsuki K, Sugaya N, Nomra S. Effects of prolonged stress on salivary cortisol and dehydroepiandrosterone: A study of a two-week teaching practice. Psychoneuroendocrinology 2012;37 :852-8

171. Brzoza Z, Kasperska-Zajac A, Badura-Brzoza K, Matysiakiewicz J, Hese RT, et al. Decline in dehydroepiandrosterone sulfate observed in chronic urticaria is associated with psychological distress. Psychosom Med 2008;70: 723–728

172. Lac G, Dutheil F, Brousse G, Triboulet-Kelly C, Chamoux A Saliva DHEAS changes in patients suffering from psychopathological disorders arising from bullying at work. Brain Cogn 2012; 80: 277–281

173. Du CL, Lin MC, Lu L, Tai JJ. Correlation of Occupational Stress Index with 24-hour Urine Cortisol and Serum DHEA Sulfate among City Bus Drivers: A Cross-sectional Study. Saf Health Work 2, 2011;2: 169–175

174. Kim MS, Lee YJ, Ahn RS Day-to-day differences in cortisol levels and molar cortisol-to-DHEA ratios among working individuals. Yonsei Med J 2010; 51: 212–218

175. Lennartsson A-K, Theorell T, Rockwood AL, Kushnir MM, Jonsdottir IH Perceived Stress at Work Is Associated with Lower Levels of DHEA-S. PloS ONE 2013; 8: e72460

176. Jorm AF, Jolley D. The incidence of dementia: a meta-analysis.Neurology 1998; 51:728-33

177. Lu SF, Mo Q, Hu S, Garippa C, Simon NG. Dehydroepiandrosterone upregulates neural androgen receptor level and transcriptional activity. J Neurobiol 2003;57:163–71

178. Suzuki M, Wright LS, Marwah P, Lardy HA, Svendsen CN. Mitotic and neurogenic effects of dehydroepiandrosterone (DHEA) on human neural stem cell cultures derived from the fetal cortex. Proc Natl Acad Sci USA 2004;101:3202–3207

179. Lang PO, Lapenna A, Pitts D, Aspinall R. Immunological pathogenesis of main age-related diseases and frailty: Role of immunosenescence. Eur Ger Med 2010;1:112–121

180. Davis
SR
Shah
SM
McKenzie
DP
Kulkarni
J
Davison
SL
Bell
RJ
. Dehydroepiandrosterone sulfate levels are associated with more favorable cognitive function in women. J Clin Endocrinol Metab. 2008;93:801-8

181. Hildreth
KL
Gozansky
WS
Jankowski
CM
Grigsby
J
Wolfe
P
Kohrt
WM
. Association of serum dehydroepiandrosterone sulfate and cognition in older adults: sex steroid, inflammatory, and metabolic mechanisms. Neuropsychology. 2013;27:356-63

182. Valenti G, Ferrucci L, Lauretani F, Ceresini G, Bandinelli S, Luci M, Ceda G, Maggio M, Schwartz RS. Dehydroepiandrosterone sulfate and cognitive function in the elderly: The InCHIANTI Study. J Endocrinol Invest 2009;32:766–772

183. Weill-Engerer S, David JP, Sazdovitch V, Liere P, Eychenne B, Pianos A, Schumacher M, Delacourte A, Baulieu EE, Akwa Y. Neurosteroid quantification in human brain regions: comparison between Alzheimer's and nondemented patients. J Clin Endocrinol Metab 2002; 87:5138-43

184. Bélanger N, Grégoire L, Bédard PJ, Di Paolo T DHEA improves symptomatic treatment of moderately and severely impaired MPTP monkeys. Neurobiol Aging 2006; 27:1684-93

185. van Niekerk JK, Huppert FA, Herbert J. Salivary cortisol and DHEA: association with measures of cognition and well-being in normal older men, and effects of three months of DHEA supplementation.Psychoneuroendocrinology 2001; 26:591-612

186. Wolf OT, Kudielka BM, Hellhammer DH, Hellhammer J, Kirschbaum C. Opposing effects of DHEA replacement in elderly subjects on declarative memory and attention after exposure to a laboratory stressor. Psychoneuroendocrinology 1998; 23:617-29

187. Wolf OT, Kirschbaum C. Review Actions of dehydroepiandrosterone and its sulfate in the central nervous system: effects on cognition and emotion in animals and humans. Brain Res Brain Res Rev 1999 ; 30:264-88

188. Arlt W, Callies F, Koehler I, van Vlijmen JC, Fassnacht M, Strasburger CJ, Seibel MJ, Huebler D, Ernst M, Oettel M, Reincke M, Schulte HM, Allolio B. Dehydroepiandrosterone supplementation in healthy men with an age-related decline of dehydroepiandrosterone secretion. J Clin Endocrinol Metab 2001 ; 86:4686-92

189. Grimley Evans J, Malouf R, Huppert F, van Niekerk JK. Review Dehydroepiandrosterone (DHEA) supplementation for cognitive function in healthy elderly people. Cochrane Database Syst Rev 2006 18; (4):CD006221

190. Wolkowitz OM, Kramer JH, Reus VI, Costa MM, Yaffe K, Walton P, Raskind M, Peskind E, Newhouse P, Sack D, De Souza E, Sadowsky C, Roberts E, DHEA-Alzheimer's Disease Collaborative Research. DHEA treatment of Alzheimer's disease: a randomized, double-blind, placebo-controlled study.Neurology. 2003; 60:1071-6

191. Yamada S, Akishita M, Fukai S, Ogawa S, Yamaguchi K, Matsuyama J, Kozaki K, Toba K, Ouchi Y. Effects of dehydroepiandrosterone supplementation on cognitive function and activities of daily living in older women with mild to moderate cognitive impairment. Geriatr Gerontol Int 2010;10:280–287

192. Parsons TD, Kratz KM, Thompson E, Stanczyk FZ, Buckwalter JG Dhea supplementation and cognition in postmenopausal women. Int J Neurosci 2006; 116:141-55

193. Morrison MF, Redei E, TenHave T, Parmelee P, Boyce AA, Sinha PS, Katz IR. Dehydroepiandrosterone sulfate and psychiatric measures in a frail, elderly residential care population. Biol Psychiatry 2000;47:144–150

194. Somboonporn W, Davis S , Seif M , Bell R . Testosterone for peri- and postmenopausal women. Cochrane Database Syst Rev 2005:CD004509

195. Darling GM , Johns JA , McCloud PI , Davis SR. Estrogen and progestin compared with simvastatin for hypercholesterolemia postmenopausal women. N Engl J Med 1997; 337:595–601

196. Barnhart KT, Freeman E, Grisso JA, Rader DJ, Sammel M, Kapoor S, Nestler JE. The effect of dehydroepiandrosterone supplementation to symptomatic perimenopausal women on serum endocrine profiles, lipid parameters, and healthrelated quality of life. J Clin Endocrinol Metab 1999;84:3896–3902

197. Davis SR, Goldstat R , Newman A , Berry K , Burger HG , Meredith I , Koch K. Differing effects of low dose estrogen and progestin replacement therapy and pravastatin in hypercholesterolemic postmenopausal women. Climacteric 2002;5:341–350

198. Flöter A , Nathorst-Böös J , Carlström K , von Schoultz B. Serum lipids in oophorectomized women during estrogen and testosterone replacement therapy. Maturitas 2004;47:123–129

199. Leão LM, Duarte MP , Silva DM , Bahia PR , Coeli CM , de Farias ML. Influence of methyltestosterone postmenopausal therapy on plasma lipids, inflammatory factors, glucose metabolism and visceral fat: a randomized study. Eur J Endocrinol 2006; 154:131–139

200. Srinivasan M, Irving BA, Frye RL, O'Brien P, Hartman SJ, McConnell JP, Nair KS. Effects on lipoprotein particles of long-term dehydroepiandrosterone in elderly men and women and testosterone in elderly men.J Clin Endocrinol Metab 2010 ;95:1617-25

 

201. Srinivasan M, Irving BA, Dhatariya K, Klaus KA, Hartman SJ, McConnell JP, Nair KS. Effect of dehydroepiandrosterone replacement on lipoprotein profile in hypoadrenal women.J Clin Endocrinol Metab 2009;94:761-4

 

202. Bell RJ, Davison SL, Papalia MA, McKenzie DP, Davis SR. Endogenous androgen levels and cardiovascular risk profile in women across the adult life span. Menopause 2007;14:630–638

203. Shufelt C, Bretsky P, Almeida CM, Johnson BD, Shaw LJ, Azziz R, Braunstein GD, Pepine CJ, Bittner V, Vido DA, Stanczyk FZ, Bairey Merz CN. DHEA-S levels and cardiovascular disease mortality in postmenopausal women: Results from the National Institutes of Health–National Heart, Lung, and Blood Institute (NHLBI)-sponsored Women’s Ischemia Syndrome Evaluation (WISE). J Clin Endocrinol Metab 2010;95:4985–4992

204. Christiansen JJ, Andersen NH, Sørensen KE, Pedersen EM, Bennett P, Andersen M, Christiansen JS, Jørgensen JO, Gravholt CH. Dehydroepiandrosterone substitution in female adrenal failure: no impact on endothelial function and cardiovascular parameters despite normalization of androgen status. Clin Endocrinol 2007;66:426–433

205. Panjari M, Bell RJ, Jane F, Adams J, Morrow C, Davis SR. The safety of 52 weeks of oral DHEA therapy for postmenopausal women. Maturitas 2009;63:240–245

206. Cleary MP, Zisk JF. Anti-obesity effect of two different levels of dehydroepiandrosterone in lean and obese middle-aged female Zucker rats. Int J Obes 1986;10:193–204

207. Mohan PF, Ihnen JS, Levin BE, Cleary MP. Effects of dehydroepiandrosterone treatment in rats with diet-induced obesity. J Nutr 1990;120:1103–1114

208. Hansen PA, Han DH, Nolte LA, Chen M, Holloszy JO.DHEA protects against visceral obesity and muscle insulin resistance in rats fed a high-fat diet. Am J Physiol 1997;273:R1704–R1708

209. Santoro N, Torrens J, Crawford S, et al. Correlates of circulating androgens in mid-life women: The Study of Women's Health Across the Nation. J Clin Endocrinol Metab 2005; 90:4836–4845

210. Corona G, Rastrelli G, Giagulli VA, Sila A, Sforza A, Forti G, Mannucci E, Maggi M. Dehydroepiandrosterone supplementation in elderly men: a meta-analysis study of placebo-controlled trials. J Clin Endocrinol Metab 2013 ;98:3615-26

211.
Steinberg, H. O., Chaker,
H., Leaming,
R., Johnson,
A., Brechtel,
G., Baron,
A.  Obesity/insulin resistance is associated with endothelial
dysfunction. J
Clin Invest 1996; 97: 26012610

 

212. Labrie F, Luu-The V, Belanger A, et al. Is dehydroepiandrosterone a hormone? J Endocrinol 2005;187:169–96

213. Livingstone C, Collison M. Sex steroids and insulin resistance. Clin Sci (Lond) 2002;102:151–66

214. Wellman M, Shane-McWhorter L, Orlando, Jennings JP. The role of dehydroepiandrosterone in diabetes mellitus. Pharmacotherapy 1999;19:582–91

215. Golden SH, Dobs AS, Vaidya D, Szklo M, Gapstur S, Kopp P, Liu K, Ouyang P. Endogenous sex hormones and glucose tolerance status in postmenopausal women. J Clin Endocrinol Metab 2007;92:1289–1295

216. Barrett-Connor E, Ferrara A. Dehydroepiandrosterone, dehydroepiandrosterone sulfate, obesity, waist-hip ratio, and noninsulin-dependent diabetes in postmenopausal women: The Rancho Bernardo Study. J Clin Endocrinol Metab 1996;81:59–64

217. Dhatariya K, Bigelow ML, Nair KS. Effect of dehydroepiandrosterone replacement on insulin sensitivity and lipids in hypoadrenal women. Diabetes 2005;54:765–769

218. Lasco A, Frisina N, Morabito N, et al. Metabolic effects of dehydroepiandrosterone replacement therapy in postmenopausal women. Eur J Endocrinol 2001;145:457–61

219. Gómez-Santos C, Larqué E, Granero E, Hernández-Morante J, Garaulet M. Dehydroepiandrosterone-sulphate replacement improves the human plasma fatty acid profile in plasma of obese women. Steroids 2011;76:1425–1432

220. Diamond P, Cusan L, Gomez JL, Belanger A, Labrie F. Metabolic effects of 12-month percutaneous dehydroepiandrosterone replacement therapy in postmenopausal women. J Endocrinol 1996;150(Suppl):S43–S50

221. Gebre-Medhin G, Husebye ES, Mallmin H, et al. Oral dehydroepiandrosterone (DHEA) replacement therapy in women with Addison’s disease. Clin Endocrinol 2000;52:775–80

222. Nestler JE, Barlascini CO, Clore JN, Blackard WG. Dehydroepiandrosterone reduces serum low density lipoprotein levels and body fat but does not alter insulin sensitivity in normal men. J Clin Endocrinol Metab 1988;66:57–61

223. Casson PR , Faquin LC , Stentz FB, Straughn AB , Andersen RN , Abraham GE , Buster JE 1995 Replacement of dehydroepiandrosterone enhances T-lymphocyte insulin binding in postmenopausal women. Fertil Steril 63:1027–1031

224.Basu R , DallaMan C , Campioni M , Basu A , Nair KS , Jensen MD , Khosla S , Klee G ,Toffolo G , Cobelli C , Rizza RA.Two years of treatment with dehydroepiandrosterone does not improve insulin secretion, insulin action, or postprandial glucose turnover in elderly men or women. Diabetes  2007; 56:753–766

225. Jellinger PS, Donald A. Smith DA et al. the AACE Task Force for Management of Dyslipidemia and Prevention of AtherosclerosisAmerican Association of Clinical Endocrinologists Guidelines for Management of Dyslipidemia and Prevention of Atherosclerosis. Endocrine Practice 2012:18(1)

226. Phillips GB, Pinkernell BH, Jing TY. Are major risk factors for myocardial infarction the major predictors of degree of coronary artery disease in men? Metabolism 2004; 53: 324–329

227. Freedman DS, O’Brien TR, Flanders WD, DeStefano F, Barboriak JJ. Relation of serum testosterone levels to high density lipoprotein cholesterol and other characteristics in men. Arterioscler Thromb, 1991; 11: 307–315

228. Tripathi Y, Hegde BM. Serum estradiol and testosterone levels following acute myocardial infarction in men. Indian J Physiol Pharmacol 1998; 42: 291–294

229. Pugh PJ, Channer KS, Parry H, Downes T, Jone TH. Bio-available testosterone levels fall acutely following myocardial infarction in men: Association with fibrinolytic factors. Endocr Res 2002; 28: 161–173

230. Glueck CJ, Glueck HI, Stroop D, Speirs J, Hamer T, Tracy T. Endogenous testosterone, fibrinolysis, and coronary heart disease risk in hyperlipidemic men. J Lab Clin Med 1993; 122: 412–420

231. Caron P, Bennet A, Camare R, Louvet JP, Boneu B, Sie P. Plasminogen activator inhibitor in plasma is related to testosterone in men. Metabolism 1989; 38: 1010–1015

 

232. De Pergola G, De MV, Sciaraffia M et al. Lower androgenicity is associated with higher plasma levels of prothrombotic factors irrespective of age, obesity, body fat distribution, and related metabolic parameters in men. Metabolism 1997; 46: 1287–1293

 

233. Militaru C, Donoiu I, Dracea O, Ionescu D-D. Serum testosterone and short-term mortality in men with acute myocardial infarction. Cardiol J 2010; 17: 249–253

234. Fukui M, Kitagawa Y, Nakamura N, Kadono M, Yoshida M, Hirata C, et al. Serum dehydroepiandrosterone sulfate concentration and carotid atherosclerosis in men with type 2 diabetes. Atherosclerosis 2005;181:339–44

 

235. Suzuki T, Yano Y, Sakamoto M, Uemura M, Yasuma T, Onishi Y, et al. Correlation of circulating dehydroepiandrosterone with activated protein C generation and carotid intima-media thickness in male patients with type 2 diabetes. Diabet Med 2012;29:e41–6

 

236. Akishita M, Hashimoto M, Ohike Y, Ogawa S, Iijima K, Eto M, et al. Association of plasma dehydroepiandrosterone-sulfate levels with endothelial function in postmenopausal women with coronary risk factors. Hypertens Res 2008;31: 69–74

237. Herrington DM. Dehydroepiandrosterone and coronary atherosclerosis. Ann NY Acad Sci 1995;774:271–280

238. Herrington DM, Nanjee N, Achuff SC, Cameron DE, Dobbs B, Baughman KL. Dehydroepiandrosterone and cardiac allograft vasculopathy. J Heart Lung Transplant 1996;15:88–93

239. Abbasi A, Duthie Jr EH, Sheldahl L, Wilson C, Sasse E, Rudman I, et al. Association of dehydroepiandrosterone sulfate, body composition, and physical fitness in independent community-dwelling older men and women. J Am Geriatr Soc 1998;46:263–73

240. Shono N, Kumagai S, Higaki Y, Nishizumi M, Sasaki H. The relationships of testosterone, estradiol, dehydroepiandrosterone-sulfate and sex hormonebinding globulin to lipid and glucose metabolism in healthy men. J Atheroscler Thromb 1996;3:45–51

241. Herrington DM. Dehydroepiandrosterone and coronary atherosclerosis. Ann N Y
Acad Sci 1995;774:271–80

242. Bernini GP, Moretti A, Sgro M, Argenio GF, Barlascini CO, Cristofani R, et al. Influence
of endogenous androgens on carotid wall in postmenopausal women. Menopause 2001;8:43–50

243. Bernini GP, Sgro M, Moretti A, Argenio GF, Barlascini CO, Cristofani R, et al. Endogenous androgens and carotid intimal-medial thickness in women. J Clin Endocrinol Metab 1999;84:2008–12

244. Alexandersen P, Haarbo J, Christiansen C. The relationship of natural androgens to coronary heart disease in males: a review. Atherosclerosis 1996;125:1–13

245. Feldman HA, Johannes CB, Araujo AB, Mohr BA, Longcope C, McKinlay JB. Lowdehydroepiandrosterone and ischemic heart disease in middle-aged men: prospective results from the Massachusetts Male Aging Study. Am J Epidemiol 2001;153:79–89

246. Trivedi DP, Khaw KT. Dehydroepiandrosterone sulfate and mortality in elderly men and women. J Clin Endocrinol Metab 2001;86:4171–7

247. Hak AE, Witteman JC, de Jong FH, Geerlings MI, Hofman A, Pols HA. Low levels of endogenous androgens increase the risk of atherosclerosis in elderly men: the Rotterdam study. J Clin Endocrinol Metab 2002;87:3632–9

248. Herrington DM, Gordon GB, Achuff SC, Trejo JF, Weisman HF, Kwiterovich Jr PO, et al. Plasma dehydroepiandrosterone and dehydroepiandrosterone sulfate in patients undergoing diagnostic coronary angiography. J Am Coll Cardiol 1990;16:862–70

249. Ishihara F, Hiramatsu K, Shigematsu S, Aizawa T, Niwa A, Takasu N, et al. Role of adrenal androgens in the development of arteriosclerosis as judged by pulse wave velocity and calcification of the aorta. Cardiology 1992;80:332–8

250. Mitchell LE, Sprecher DL, Borecki IB, Rice T, Laskarzewski PM, Rao DC. Evidence for an association between dehydroepiandrosterone sulfate and nonfatal, premature myocardial infarction in males. Circulation 1994;89: 89–93

251. Barrett-Connor E, Goodman-Gruen D. The epidemiology of DHEAS and cardiovascular disease. Ann N Y Acad Sci 1995;774:259–70

252. Barrett-Connor E, Goodman-Gruen D. Dehydroepiandrosterone sulfate does not predict cardiovascular death in postmenopausal women. The Rancho Bernardo Study. Circulation 1995;91:1757–60

253. Berr C, Lafont S, Debuire B, Dartigues JF, Baulieu EE. Relationships of dehydroepiandrosterone sulfate in the elderly with functional, psychological, and mental status, and short-term mortality: a French community-based study. Proc Natl Acad Sci U S A 1996;93:13410–15

254. Nakamura S, Yoshimura M, Nakayama M, Ito T, Mizuno Y, Harada E, Sakamoto T, Saito Y, Nakao K, Yasue H, Ogawa H. Possible association of heart failure status with synthetic balance between aldosterone and dehydroepiandrosterone in human heart. Circulation 2004;110:1787–1793

255. Jankowska EA, Biel B, Majda J, Szklarska A, Lopuszanska M, Medras M, Anker SD, Banasiak W, Poole-Wilson PA, Ponikowski P. Anabolic deficiency in men with chronic heart failure: prevalence and detrimental impact on survival. Circulation 2006;114:1829–1837

256. Liu D, Dillon JS. Dehydroepiandrosterone stimulates nitric oxide release in vascular endothelial cells: evidence for a cell surface receptor. Steroids 2004;69:279–89

257. Formoso G, Chen H, Kim JA, Montagnani M, Consoli A, Quon MJ. Dehydroepiandrosterone mimics acute actions of insulin to stimulate production of both nitric oxide and endothelin 1 via distinct phosphatidylinositol 3-kinaseand mitogen-activated protein kinase-dependent pathways in vascular endothelium. Mol Endocrinol 2006;20:1153–63

258. Duncker DJ, Bache RJ. Inhibition of nitric oxide production aggravates myocardial hypoperfusion during exercise in the presence of a coronary artery stenosis.Circ Res 1994; 74:629–640

259. Wang L, Hao Q, Wang YD, Wang WJ, Li DJ. Protective effects of dehydroepiandrosterone on atherosclerosis in ovariectomized rabbits via alleviating inflammatory injury in endothelial cells. Atherosclerosis 2011;214:47–57

260. Ross R. Atherosclerosis: an inflammatory disease. N Engl J Med. 1999; 340:115–126

261. Libby P, Geng Y-J, Aikawa M, et al. Macrophages and atherosclerotic plaque stability. Curr Opin Lipidol 1996; 7: 330–335

 

262. Libby P. Current concepts of the pathogenesis of the acute coronary syndromes. Circulation 2001; 104: 365–372

263. Libby P, Simon DI. Inflammation and thrombosis: the clot thickens. Circulation 2001; 103: 1718–1720

 

264.
Belosjorow S, Bolle I, Duschin A, Heusch G, Schulz R. TNFα
antibodies are as effective as ischemic preconditioning in reducing infarct size in rabbits. Am J Physiol Heart Circ Physiol 2003; 284:H927–H930

265. Liu D, Si H, Reynolds KA, Zhen W, Jia Z, Dillon JS. Dehydroepiandrosterone protects vascular endothelial cells against apoptosis through a Galphai protein-dependent activation of phosphatidylinositol 3-kinase/Akt and regulation of antiapoptotic Bcl-2 expression. Endocrinology 2007;148:3068–76

266. Chen J, Xu L,Huang C. DHEA inhibits vascular remodeling following arterial injury: a possible role in suppression of inflammation and oxidative stress derived from vascular smooth muscle cells. Mol Cell Biochem 2014;388:75-84

267. Hautanen A, Manttari M, Manninen V, Tenkanen L, Huttunen JK, Frick MH, et al. Adrenal androgens and testosterone as coronary risk factors in the Helsinki Heart Study. Atherosclerosis 1994;105:191–200

268. LaCroix AZ, Yano K, Reed DM. Dehydroepiandrosterone sulfate, incidence of myocardial infarction, and extent of atherosclerosis in men. Circulation 1992;86:1529–35

269. Zumoff B, Troxler RG, O’Connor J, Rosenfeld RS, Kream J, Levin J, et al. Abnormal hormone levels in men with coronary artery disease. Arteriosclerosis 1982;2:58–67

270. Slowinska-Srzednicka J, Zgliczynski S, Ciswicka-Sznajderman M, Srzednicki M, Soszynski P, Biernacka M, et al. Decreased plasma dehydroepiandrosterone sulfate and dihydrotestosterone concentrations in young men
after myocardial infarction. Atherosclerosis 1989;79:197–203

271. Haffner SM, Moss SE, Klein BE, Klein R. Sex hormones and DHEA-SO4 in relation to ischemic heart disease mortality in diabetic subjects. The Wisconsin Epidemiologic Study of Diabetic Retinopathy. Diabetes Care 1996;19: 1045–50

272. Golden SH, Maguire A, Ding J, Crouse JR, Cauley JA, Zacur H, et al. Endogenous postmenopausal hormones and carotid atherosclerosis: a case–control study of the atherosclerosis risk in communities cohort. Am J Epidemiol 2002;155: 437–45

273. Khaw KT, Tazuke S, Barrett-Connor E. Cigarette smoking and levels of adrenal androgens in postmenopausal women. N Engl J Med 1988;318:1705–9

274. Salvini S, Stampfer MJ, Barbieri RL, Hennekens CH. Effects of age, smoking and vitamins on plasma DHEAS levels: a cross-sectional study in
men. J Clin Endocrinol Metab 1992;74:139–43

275. Wu FC, von Eckardstein A. Androgens and coronary artery disease. Endocr Rev 2003;24:183–217

276. Kawano H, Yasue H, Kitagawa A, Hirai N, Yoshida T, Soejima H, et al.Dehydroepiandrosterone supplementation improves endothelial function and insulin sensitivity in men. J Clin Endocrinol Metab 2003;88:3190–5

277. Eich DM, Nestler JE, Johnson DE, Dworkin GH, Ko D, Wechsler AS, et al. Inhibition of accelerated coronary atherosclerosis with dehydroepiandrosterone in the heterotopic rabbit model of cardiac transplantation.
Circulation 1993;87:261–9

278. Gordon GB, Bush DE, Weisman HF. Reduction of atherosclerosis by administration of dehydroepiandrosterone. A study in the hypercholesterolemic New Zealand white rabbit with aortic intimal injury. J Clin
Invest 1988;82:712–20

279. Aragno M, Parola S, Brignardello E, Mauro A, Tamagno E, Manti R, et al. Dehydroepiandrosterone prevents oxidative injury induced by transient ischemia/reperfusion in the brain of diabetic rats.
Diabetes 2000;49: 1924–31

280. Ayhan S, Tugay C, Norton S, Araneo B, Siemionow M. Dehydroepiandrosterone protects the microcirculation of muscle flaps from ischemia-reperfusion injury by reducing the expression of adhesion molecules. Plast Reconstr Surg 2003;111:2286–94

281. Kinlay S, Creager MA, Fukumoto M, Hikita H, Fang JC, Selwyn AP, Ganz P. Endothelium-derived nitric oxide regulates arterial elasticity in human arteries in vivo. Hypertension. 2001 Nov; 38(5):1049-53.

 

282. Weiss EP, Villareal Ehsani AA, Fontana L, Holloszy JO. Dehydroepiandrosterone replacement therapy in older adults improves indices of arterial stiffness. Aging Cell 2012; 11: 876–884

283. Murray CJ, Vos T, Lozano R, Naghavi M, Flaxman AD, et al. Disability-adjusted life years (DALYs) for 291 diseases and injuries in 21 regions, 1990–2010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet 2013;380: 2197–2223

284. American stroke Association. http://www.strokeassociation.org/STROKEORG/AboutStroke/TypesofStroke/Types-of-Stroke_UCM_308531_SubHomePage.jsp

 

285. Baulieu EE, Robel P. Dehydroepiandrosterone (DHEA) and dehydroepiandrosterone sulfate (DHEAS) as neuroactive neurosteroids. Proc Natl Acad Sci U S A 1998; 95: 4089–4091

286. Phillips AC, Carroll D, Gale CR, Lord JM, Arlt W, et al. Cortisol, DHEA sulphate, their ratio, and all-cause and cause-specific mortality in the Vietnam Experience Study. Eur J Endocrinol 2010; 163: 285–292

287. Barrett-Connor E, Khaw KT, Yen SS. A prospective study of dehydroepiandrosterone sulfate, mortality, and cardiovascular disease. N Engl J Med 1986;315: 1519–1524

288. Bologa L, Sharma J, Roberts E. Dehydroepiandrosterone and its sulfated derivative reduce neuronal death and enhance astrocytic differentiation in brain cell cultures. J Neurosci Res 1987; 17: 225–234.

289. Lapchak PA, Chapman DF, Nunez SY, Zivin JA. Dehydroepiandrosterone sulfate is neuroprotective in a reversible spinal cord ischemia model: possible involvement of GABA(A) receptors. Stroke 2000; 31: 1953–1956; discussion 1957

290. Corpechot C, Robel P, Axelson M, Sjovall J, Baulieu EE. Characterization and measurement of dehydroepiandrosterone sulfate in rat brain. Proc Natl Acad Sci U S A 1981;78: 4704–4707

291. Baulieu EE. Neuroactive neurosteroids: dehydroepiandrosterone (DHEA) and DHEA sulphate. Acta Paediatr 1999; Suppl 88: 78–80

292. Jiménez MC, Sun Q, Schürks M, Chiuve S, Hu FB, Manson JE, Rexrode KM. Low dehydroepiandrosterone sulfate is associated with increased risk of ischemic stroke among women. Stroke 2013;44:1784-9

293. Pappa T, Vemmos K, Saltiki K, Mantzou E, Stamatelopoulos K, Alevizaki M. Severity and outcome of acute stroke in women: relation to adrenal sex steroid levels. Metabolism 2012;61:84-91

294. Blum CA, Mueller C, Schuetz P, Fluri F, Trummler M, Mueller B, Katan M, Christ-Crain M. Prognostic value of dehydroepiandrosterone-sulfate and other parameters of adrenal function in acute ischemic stroke. PLoS One 2013;8:e63224

295. Hampl V, Bibova J, Povysilova V, Herget J. Dehydroepiandrosterone sulphate reduces chronic hypoxic pulmonary hypertension in rats. Eur Respir J 2003;21:862–5

 

296. Bonnet S, Dumas-de-La-Roque E, Begueret H, Marthan R, Fayon M, Dos Santos P, et al. Dehydroepiandrosterone (DHEA) prevents and reverses chronic hypoxic pulmonary hypertension. Proc Natl Acad Sci USA 2003;100:9488–93

 

297. Debonneuil EH, Quillard J, Baulieu EE. Hypoxia and dehydroepiandrosterone in old age: A mouse survival study. Respir Res 2006;7:144

 

298. Oka M, Karoor V, Homma N, Nagaoka T, Sakao E, Golembeski SM, et al. Dehydroepiandrosterone upregulates soluble guanylate cyclase and inhibits hypoxic pulmonary hypertension. Cardiovasc Res 2007;74:377–87

299. Debonneuil EH, Quillard J, Baulieu EE. Hypoxia and dehydroepiandrosterone in old age: A mouse survival study. Respir Res 2006;7:144

300. Farrukh IS, Peng W, Orlinska U, Hoidal JR. Effect of dehydroepiandrosterone on hypoxic pulmonary vasoconstriction: A Ca(2+)-activated K(+)-channel opener. Am J Physiol. 1998;274:L186–95

301. Gupte SA, Li KX, Okada T, Sato K, Oka M. Inhibitors of pentose phosphate pathway cause vasodilation: Involvement of voltage-gated potassium channels. J Pharmacol Exp Ther 2002;301:299–305

 

302. Dessouroux A, Akwa Y, Baulieu EE. DHEA decreases HIF-1alpha accumulation under hypoxia in human pulmonary artery cells: Potential role in the treatment of pulmonary arterial hypertension. J Steroid Biochem Mol Biol 2008;109:81–9

 

303. Patel D,  Kandhi S,  Kelly M, Neo BH,  Wolin MS. Dehydroepiandrosterone  promotes pulmonary artery relaxation by NADPH oxidation-elicited subunit dimerization of protein kinase G 1αAm J Physiol Lung Cell Mol Physiol

304. Ventetuolo CE, Ouyang P, Bluemke DA, Tandri H, Barr RG, Bagiella E, et al. Sex hormones are associated with right ventricular structure and function: The mesa-right ventricle study. Am J Respir Crit Care Med 2011;183:659–67

 

305. Mazat L, Lafont SBerr CDebuire BTessier JFDartigues JFBaulieu EE Prospective measurements of dehydroepiandrosterone sulfate in a cohort of elderly subjects: Relationship to gender, subjective health, smoking habits, and 10-year mortality Proc Natl Acad Sci U S A  2001 3;98:8145-50

 

306. Dumas de La Roque ESavineau
JP
Metivier ACBilles MAKraemer JPDoutreleau SJougon JMarthan RMoore NFayon M,Baulieu EÉDromer C. Dehydroepiandrosterone (DHEA) improves pulmonary hypertension in chronic obstructivepulmonary disease (COPD): a pilot tudy. Ann Endocrinol (Paris) 2012 ;73:20-5

 

307. Straub RH, Vogl D, Gross V, Lang B, Schölmerich J, Andus T. Association of humoral markers of inflammation and dehydroepiandrosterone sulfate or cortisol serum levels in patients with chronic inflammatory bowel disease. Am J Gastroenterol 1998; 93: 2197–202

308. De TB, Hedman M, Befrits R. Blood and tissue dehydroepiandrosterone sulphate levels and their relationship to chronic inflammatory bowel disease. Clin Exp Rheumatol 1998; 16: 579–82

309. Andus T1, Klebl FRogler GBregenzer NSchölmerich JStraub RH. Patients with refractory Crohn's disease or ulcerative colitis respond to dehydroepiandrosterone: a pilot study. Aliment Pharmacol Ther 2003 ;17:409-14

310. American Cancer Society. http://www.cancer.org/cancer/breastcancer/).

311. Van Vollenhoven RF, Engleman EG, McGuire JL. An open study of dehydroepiandrosterone in systemic lupus erythematosus.Arthritis Rheum 1994; 37: 1305–10

312. Van Vollenhoven RF, Engleman EG, McGuire JL. Dehydroepiandrosterone in systemic lupus erythematosus. Results of a double-blind, placebo-controlled, randomized clinical trial. Arthritis Rheum 1995; 38: 1826–31

313. Van Vollenhoven RF, Morabito LM, Engleman EG, McGuire JL. Treatment of systemic lupus erythematosus with dehydroepiandrosterone: 50 patients treated up to 12 months. J Rheumatol 1998; 25: 285–9

314. Van Vollenhoven RF, Park JL, Genovese MC, West JP, McGuire JL. A double-blind, placebo-controlled, clinical trial of dehydroepiandrosterone in severe systemic lupus erythematosus. Lupus 1999; 8: 181–7

315. Mease PJ, Merril JT, Lahita RG, et al. GL701 (prasterone, dehydroepiandrosterone) improves systemic lupus erythematosus. Arthritis Rheum 2000; 43: 271

316. Chang D-M, Lan JL, Lock G, Luo SF. Dehydroepiandrosterone treatment of women with mild to moderate systemic lupus erythematosus: a multi-center, randomized, double-blind, placebo-controlled trial. Arthritis Rheum 2002; 46: 2924–7

317. Wilder RL. Adrenal and gonadal steroid hormone deficiency in the pathogenesis of rheumatoid arthritis. J Rheumatol 1996; Suppl 44: 10-12

318. Kanik KS, Chrousos GP, Schumacher HR, Crane ML, Yarboro CH, and Wilder RL. Adrenocorticotropin, glucocorticoid, and androgen secretion in patients with new onset synovitis/rheumatoid arthritis: relations with indices of inflammation. J Clin Endocrinol Metab 2000; 85: 1461-1466

319. Yukioka M, Komatsubara Y, Yukioka K, Toyosaki-Maeda T, Yonenobu K, Ochi T. Adrenocorticotropic hormone and dehydroepiandrosterone sulfate levels of rheumatoid arthritis patients treated with glucocorticoids. Mod Rheumatol 2006;16:30-5

 

320. Danenberg HD, Alpert G, Lustig S, Ben-Nathan D: Dehydroepiandrosterone protects mice from endotoxin toxicity and reduces tumor necrosis factor production. Antimicrob Agents Chemother 1992; 36:2275-2279

321. Daynes RA, Araneo BA, Ershler WB, Maloney C, Li GZ, Ryu SY: Altered regulation of IL-6 production with normal aging. Possible linkage to the age-associated decline in dehydroepiandrosterone and its sulfated derivative. J Immunol 1993; 150:5219-5230

322. Keller ET, Chang C, Ershler WB: Inhibition of NFkappaB activity through maintenance of IkappaBalpha levels contributes to dihydrotestosterone-mediated repression of the interleukin-6 promoter. J Biol Chem 1996, 271:26267-26275

323. Williams PJ, Jones RH, and Rademacher TW. Reduction in the incidence and severity of collagen-induced arthritis in DBA/1 mice, using exogenous dehydroepiandrosterone. Arthritis Rheum 1997; 40: 907-911

324. Kobayashi Y, Tagawa N, Muraoka K, Okamoto Y, and Nishida M. Participation of endogenous dehydroepiandrosterone and its sulfate in the pathology of collagen-induced arthritis in mice. Biol Pharm Bull 2003;26: 1596-1599

325. Röntzsch A, Thoss K, Petrow PK, Henzgen S, and Bräuer R. Amelioration of murine antigen-induced arthritis by dehydroepiandrosterone (DHEA). Inflamm Res 2004; 53: 189-198

326. Giltay EJ, van Schaardenburg D, Gooren LJ, and Dijkmans BA. Dehydroepiandrosterone sulfate in patients with rheumatoid arthritis. Ann N Y Acad Sci 1999; 876: 152-154

327. Cutolo M. Sex hormone adjuvant therapy in rheumatoid arthritis. Rheum Dis Clin North Am 2000; 26: 881-89

328. Marwah A, Marwah P, and Lardy H. Ergosteroids. VI. Metabolism of dehydroepiandrosterone by rat liver in vitro: a liquid chromatographic-mass spectrometric study. J Chromatogr B Analyt Technol Biomed Life Sci 2002;767: 285-299

329. Gambacciani M, Monteleone P, Sacco A, Genazzani AR. Hormone replacement therapy and endometrial, ovarian and colorectal cancer. Best Pract Res Clin Endocrinol Metab 2003; 17: 139-147

330. Hankinson SE,  Colditz GA and  Willett WC. Towards an integrated model for breast cancer etiology: The lifelong interplay of genes, lifestyle, and hormones. Breast Cancer Res 2004; 6:213-218 

 

331. Kaaks R1, Lukanova AKurzer MS. Obesity, endogenous hormones, and endometrial cancer risk: a synthetic review. Cancer Epidemiol Biomarkers Prev. 2002;11:1531-43.

 

332. Kaaks R, Berrino F, Key T, Rinaldi S, Dossus L, Biessy C, et al. Serum sex steroids in premenopausal women and breast cancer risk within the European Prospective Investigation into Cancer and Nutrition (EPIC). J Natl Cancer Inst 2005;97:755-765

333. Tworoger SS1, Missmer SAEliassen AHSpiegelman DFolkerd EDowsett MBarbieri RLHankinson SE. The association of plasma DHEA and DHEA sulfate with breast cancer risk in predominantly premenopausal women. Cancer Epidemiol Biomarkers Prev. 2006;15:967-71

334. Key T, Appleby P, Barnes I, Reeves G. Endogenous sex hormones and breast cancer in postmenopausal women: reanalysis of nine prospective studies. J Natl Cancer Inst 2002;94:606–16

335. Morris KT1, Toth-Fejel SSchmidt JFletcher WSPommier RF. High dehydroepiandrosterone-sulfate predicts breast cancer progression during new aromatase inhibitor therapy and stimulates breast cancer cell growth in tissue culture: a renewed role for adrenalectomy. Surgery 2001 ;130:947-53

336. Labrie F, Luu-The V, Labrie C, et al. Endocrine and intracrine sources of androgens in women: inhibition of breast cancer and other roles of androgens and their precursor dehydroepiandrosterone. Endocrine Rev 2003;24:152

337. Schwartz AG, Pashko LL. Cancer prevention with dehydroepiandrosterone and non-androgenic structural analogs. J Cell Biochem Supplement 1995;22:210–217

 

338. Maggiolini M, Donze O, Jeannin E, Ando S, Picard D. Adrenal androgens stimulate the proliferation of breast cancer cells as direct activators of estrogen receptor α. Cancer Res 1999;59:4864–9

 

339. Seymour-Munn K, Adams J. Estrogenic effects of 5-androstene-3β,17β-diol at physiological concentrations and its possible implication in the etiology of breast cancer. Endocrinology 1983;112:486–91

 

340. Stoll BA. Dietary supplements of dehydroepiandrosterone in relation to breast cancer risk. Eur J Clin Nutr. 1999;53:771-5

341. Genazzani AD, Stomati M, Bernardi F, Pieri M, Rovati L, Genazzani AR. Longterm low-dose dehydroepiandrosterone oral supplementation in early and late postmenopausal women modulates endocrine parameters and synthesis of neuroactive steroids. Fertil Steril 2003;80:1495–501

342. Stomati M, Monteleone P, Casarosa E, et al. Six-month oral dehydroepiandrosterone supplementation in early and late postmenopause. Gynecol Endocrinol 2000;14:342–63

343. Labrie F, Diamond P, Cusan L, Gomez JL, Belanger A, Candas B. Effect of 12-month dehydroepiandrosterone replacement therapy on bone, vagina, and endometrium in postmenopausal women. J Clin Endocrinol Metab 1997;82:3498–505

344. Arnold JT. DHEA metabolism in prostate: For better or worse? Mol Cell Endocrinol 2009;301:83–88

Adult Hypothyroidism

9.1 HISTORICAL

The full-blown expression of hypothyroidism is known as myxedema. Adult myxedema escaped serious attention until Gull described it in 1874 1 . That it was a state resembling the familiar endemic cretinism, but coming on in adult life, was what chiefly impressed Gull. Ord 2 invented the term myxedema in 1873. The disorder arising from surgical removal of the thyroid gland (cachexia strumipriva) was described in 1882 by Reverdin 3 of Geneva and in 1883 by Kocher of Berne 4 . After Gull's description, myxedma aroused enormous interest, and in 1883 the Clinical Society of London appointed a committee to study the disease and report its findings 5 . The committee's report, published in 1888, contains a significant portion of what is known today about the clinical and pathologic aspects of myxedema. It is referred to in the following discussion as the Report on Myxedema. The final conclusions of the 200-page volume are penetrating. They are as follows:

1. That myxedema is a well-defined disease.

2. That the disease affects women much more frequently than men, and that the subjects are for the most part of middle age.

3. That clinical and pathological observations, respectively, indicate in a decisive way that the one condition common to all cases is destructive change of the thyroid gland.

4. That the most common form of destructive change of the thyroid gland consists in the substitution of a delicate fibrous tissue for the proper glandular structure.

5. That the interstitial development of fibrous tissue is also observed very frequently in the skin, and, with much less frequency, in the viscera, the appearances presented by this tissue being suggestive of an irritative or inflammatory process.

6. That pathological observation, while showing cause for the changes in the skin observed during life, for the falling off the hair, and the loss of the teeth, for the increased bulk of body, as due to the excess of subcutaneous fat, affords no explanation of the affections of speech, movement, sensation, consciousness, and intellect, which form a large part of the symptoms of the disease.

7. That chemical examination of the comparatively few available cases fails to show the general existence of an excess of mucin in the tissues adequately corresponding to the amount recorded in the first observation, but that this discrepancy may be, in part, attributed to the fact that tumefaction of the integuments, although generally characteristic of myxedema, varies considerably throughout the course of the disease, and often disappears shortly before death.

8. That in experiments made upon animals, particularly on monkeys, symptoms resembling in a very close and remarkable way those of myxedema have followed complete removal of the thyroid gland, performed under antiseptic precautions, and with, as far as could be ascertained, no injury to the adjacent nerves or to the trachea.

9. That in such experimental cases a large excess of mucin has been found to be present in the skin, fibrous tissues, blood, and salivary glands; in particular the parotid gland, normally containing no mucin, has presented that substance in quantities corresponding to what would be ordinarily found in the submaxillary gland.

10. That following removal of the thyroid gland in man in an important proportion of the cases, symptoms exactly corresponding with those of myxedema subsequently develop.

11. That in a considerable number of cases the operation has not been known to have been followed by such symptoms, the apparent immunity being in many cases probably due to the presence and subsequent development of accessory thyroid glands, or to accidentally incomplete removal, or to insufficiently long observation of the patients after operation.

12. That, whereas injury to the trachea, atrophy of the trachea, injury of the recurrent laryngeal nerves, injury of the cervical sympathetic, and endemic influences, have been by various observers supposed to be the true cases of experimental or of operative myxedema (cachexia strumipriva), there is, in the first place, no evidence to show that, of the numerous and various surgical operations performed on the neck and throat, involving various organs and tissues, any, save those in which the thyroid gland has been removed, have been followed by the symptoms under consideration; that in many of the operations on man, and in most, if not all, of the experimental operations made by Professor Horsley on monkeys and other animals, this procedure avoided all injury of surrounding parts, and was perfectly antiseptic; that myxedema has followed removal of the thyroid gland in persons neither living in nor having lived in localities the seat of endemic cretinism; that, therefore, the positive evidence on this point vastly outweighs the negative; and that it appears strongly proved that myxedema is frequently produced by the removal, as well as by the pathological destruction, of the thyroid gland.

13. That whereas, according to Clause 2, in myxedema women are much more numerously affected than men, in the operative form of myxedema no important numerical difference is observed.

14. That a general review of symptoms and pathology leads to the belief that the disease described under the name of myxedema, as observed in adults, is practically the same disease as that named sporadic cretinism when affecting children; that myxedema is probably identical with cachexia strumipriva; and that a very close affinity exists between myxedema and endemic cretinism.

15. That while these several conditions appear, in the main, to depend on, or to be associated with, destruction or loss of the function of the thyroid gland, the ultimate cause of such destruction or loss is at present not evident.

9.2 DEFINITION AND EPIDEMIOLOGY OF HYPOTHYROIDISM

Hypothyroidism is traditionally defined as deficient thyroidal production of thyroid hormone. The term primary hypothyroidism indicates decreased thyroidal secretion of thyroid hormone by factors affecting the thyroid gland itself; the fall in serum concentrations of thyroid hormone causes an increased secretion of TSH resulting in elevated serum TSH concentrations. Decreased thyroidal secretion of thyroid hormone can also be caused by insufficient stimulation of the thyroid gland by TSH, due to factors directly interfering with pituitary TSH release (secondary hypothyroidism) or indirectly by diminishing hypothalamic TRH release (tertiary hypothyroidism); in clinical practice it is not always possible to discriminate between secondary and tertiary hypothyroidism, which are consequently often referred to as central hypothyroidism. In rare cases, symptoms and signs of thyroid hormone deficiency are caused by the inability of tissues to respond to thyroid hormone by mutations in the nuclear thyroid hormone receptor TRß; this condition, known as thyroid hormone resistance (see Ch. 16 ), is associated with an increased thyroidal secretion of thyroid hormones and increased thyroid hormone concentrations in serum in an attempt of the body to overcome the resistance to thyroid hormone. Mutations in other genes involved with extrathyroidal metabolism and action of thyroid hormones in target tissues may also cause a hypothyroid state. Such cases could be labelled as peripheral (extrathyroidal) hypothyroidism. It thus seems more appropriate to define hypothyroidism as thyroid hormone deficiency in target tissues, irrespective of its cause.

9.2.1. GRADES OF HYPOTHYROIDISM

Hypothyroidism is a graded phenomenon, ranging from very mild cases in which biochemical abnormalities are present but the individual hardly notices symptoms and signs of thyroid hormone deficiency, to very severe cases in which the danger exists to slide down into a life-threatening myxedema coma. In the development of primary hypothyroidism, the transition from the euthyroid to the hypothyroid state is first detected by a slightly elevated serum TSH, caused by a minor decrease in thyroidal secretion of T4 which doesn't give rise to subnormal serum T4 concentrations. The reason for maintaining T4 values within the reference range is the exquisite sensitivity of the pituitary thyrotroph for even very small decreases of serum T4, as exemplified by the log-linear relationship between serum TSH and serum FT4 1 . A further decline in T4 secretion results in serum T4 values below the lower normal limit and even higher TSH values, but serum T3 concentrations remain within the reference range. It is only in the last stage that subnormal serum T3 concentrations are found, when serum T4 has fallen to really very low values associated with markedly elevated serum TSH concentrations ( Figure 9-1 ). Hypothyroidism is thus a graded phenomenon, in which the first stage of subclinical hypothyroidism may progress via mild hypothyroidism towards overt hypothyroidism ( Table 9-1 ) 3 .

Figure 9-1. Individual and median values of thyroid function tests in patients with various grades of hypothyroidism.  Discontinuous horizontal lines represent upper limit (TSH) and lower limit (FT4,T3) of the normal reference ranges. (Reproduced with permission) (2)

Table 9-1. Grades of hypothyroidism

Grade 1 Subclinical hypothyroidism TSH + FT4 N T3 N(+)
Grade 2 Mild hypothyroidism TSH + FT4 - T3 N
Grade 3 Overt hypothyroidism TSH + FT4 - T3 -
+, above upper normal limit; N, within normal reference range; -, below lower normal limit.

Maintenance of a normal serum T3 concentration until a relatively late stage in the development of hypothyroidism obviously serves as an appropriate mechanism of the body to counteract the impact of diminishing production of T4. It is accomplished by a preferential thyroidal secretion of T3: the increased secretion of TSH enhances the synthesis of T3 more than that of T4 and stimulates thyroidal 5'-monodeiodination of T4 into T3 4,5 . It explains why sometimes a slightly elevated serum T3 is found in the early stage of development of hypothyroidism. About 80% of the daily production rate of T3 is generated in extrathyroidal tissues via the conversion of T4 into T3. The peripheral tissues also have a defense mechanism against developing hypothyroidism by increasing the overall fractional conversion rate of T4 into T3 6 .

9.2.2. EPIDEMIOLOGY OF HYPOTHYROIDISM

Thyroid hormone resistance syndromes are seldom the cause of hypothyroidism; the number of registered patients approximates one thousand (see Ch. 16 ). Central hypothyroidism is also rare; its precise prevalence is unknown, but has been estimated as 0.005% in the general population 7 . Primary hypothyroidism, in contrast, is a very prevalent disease worldwide. It can be endemic in iodine-deficient regions (see Ch. 20 ), but it is also a common disease in iodine-replete areas as evident from prevalence and incidence figures reported in a number of population-based studies 8-14 . The most extensive data has been obtained from the Whickham Survey, a study of 2779 adults randomly selected of the general population in Great Britain who were evaluated between 1972 and 1974 and again twenty years later 8,9 . Most striking are the high prevalence of thyroid microsomal (peroxidase) antibodies and of (subclinical) hypothyroidism, and the marked female preponderance ( Table 9-2 ).

Table 9-2. Prevalence and incidence of thyroid antibodies and hypothyroidism in the Whickham survey (8,9). Women Men
Prevalence • thyroglobulin antibodies• microsomal (TPO) antibodies• subclinical hypothyroidism• hypothyroidism 30 per 1000103 per 100075 per 100018 per 1000 9 per 100027 per 100028 per 10001 per 1000
Incidence • hypothyroidism 4.1 per 1000 per yr 0.6 per 1000 per yr

 

The mean incidence of spontaneous hypothyroidism in women was 3.5/1000 survivors/year, that of hypothyroidism after destructive treatment for thyrotoxicosis 0.6/1000 survivors/year; similar figures were obtained in those who had deceased during follow-up. The hazard rate (the probability to develop hypothyroidism) increased with age; the mean age at diagnosis of hypothyroidism in women was 60 years. Studies from other countries like the USA 10,11 , Japan 12 and Sweden 13 report essentially similar data.

Of particular interest are risk factors for development of hypothyroidism. In women survivors of the Whickham Survey, the risk of developing overt hypothyroidism was 4.3% per year if both raised serum TSH and thyroid antibodies were present initially, 2.6% per year if raised serum TSH was present alone, and 2.1% per year if thyroid antibodies were present alone 9 . At the time of follow-up twenty years later, hypothyroidism had developed in these three groups in 55%, 33% and 27% respectively, but only in 4% if initial serum TSH was normal and thyroid antibodies were absent. The probability of developing hypothyroidism already increases at a rise in serum TSH above 2 mU/L as shown in Figure 9-2 , in thyroid antibody positive as well as in thyroid antibody negative women; it also increases with higher titres of thyroid microsomal antibodies 9 , 15 . These data are confirmed by two other more recent large population-based longitudinal surveys with a mean follow-up of 11-13 years. A figure almost identical to figure 9.2 was obtained in an Austalian study, in which the odds of hypothyroidism increased at TSH >2.5 mU/L, being always higher in the presence of TPO antibodies 16 . Increasing TSH values within the reference range of 0.2-4.5 mU/L gradually increased the risk of future hypothyroidism in the Norwegian HUNT study: odds ratio’s were significantly higher at baseline TSH >1.5 mU/L in women and at TSH > 2.0 in men 17 .

Figure 9-2. Logit probability (log odds) for the development of hypothyroidism as a function of TSH values at first survey during a 20-year follow-up of 912 women in the Whickham Survey. (Reproduced with permission)(9).

9.3 CAUSES OF HYPOTHYROIDISM

A variety of functional or structural disorders may lead to hypothyroidism, the severity of which depends on the degree and duration of thyroid hormone deprivation. A classification according to etiology appears in Table 9-3 . The two principal categories are primary (or thyroprivic) hypothyroidism caused by an inherent inability of the thyroid gland to supply a sufficient amount of the hormone, and central (or trophoprivic) hypothyroidism due to inadequate stimulation of an intrinsically normal thyroid gland resulting from a defect at the level of the pituitary (secondary hypothyroidism) or the hypothalamus (tertiary hypothyroidism). In a third (uncommon) form of hypothyroidism, regulation and function of thyroid gland are intact. Instead, manifestations of hormone deprivation arise from a disorder in the target tissues that reduces their responsiveness to the hormone (peripheral tissue resistance to thyroid hormone) or that inactivates the hormone (in massive infantile hemangiomas).

The most common cause of hypothyroidism is destruction of the thyroid gland by disease or as a consequence of vigorous ablative therapies to control thyrotoxicosis. Primary hypothyroidism may also result from inefficient hormone synthesis caused by inherited biosynthetic defects (see Ch. 16 ), a deficient supply of iodine (see Ch. 20 ), or inhibition of hormonogenesis by various drugs and chemicals (see Ch. 5 ). In such instances, hypothyroidism is typically associated with thyroid gland enlargement (goitrous hypothyroidism).

Table 9-3. Causes of hypothyroidism
1. Central (hypothalamic/pituitary) hypothyroidism

  1. Loss of functional tissue
    1. tumors (pituitary adenoma, craniopharyngioma, meningioma, dysgerminoma, glioma, metastases)
    2. trauma (surgery, irradiation, head injury)
    3. vascular (ischemic necrosis, hemorrhage, stalk interrruption, aneurysm of internal carotid artery)
    4. infections (abcess, tuberculosis, syphilis, toxoplasmosis)
    5. infiltrative (sarcoidosis, histiocytosis, hemochromatosis)
    6. chronic lymphocytic hypophysitis
    7. congenital (pituitary hypoplasia, septooptic dysplasia, basal encephalocele)
  2. Functional defects in TSH biosynthesis and release
    1. mutations in genes encoding for TRH receptor, TSHß, pituitary transcription factors (Pit-1, PROP1, LHX3, LHX4, HESX1), or LEPr, IGSF1
    2. drugs: dopamine; glucocorticoids; bexarotene; L-T4 withdrawal
2. Primary (thyroidal) hypothyroidism

  1. Loss of functional thyroid tissue
    1. chronic autoimmune thyroiditis
    2. reversible autoimmune hypothyroidism (silent and postpartum thyroiditis, cytokine-induced thyroiditis).
    3. surgery and irradiation (131I or external irradiation)
    4. infiltrative and infectious diseases, subacute thyroiditis
    5. thyroid dysgenesis
  2. Functional defects in thyroid hormone biosynthesis and release
    1. congenital defects in thyroid hormone biosynthesis
    2. iodine deficiency and iodine excess
    3. drugs: antithyroid agents, lithium, natural and synthetic goitrogenic chemicals, tyrosine kinase inhibitors
3. "Peripheral" (extrathyroidal) hypothyroidism

  1. Consumptive hypothyroidism (massive infantile hemangioma)
  2. Mutations in genes encoding for MCT8, SECISBP2, TRα or TR β (thyroid hormone resistance)

9.3.1. CENTRAL HYPOTHYROIDISM

Hypothalamic disorders cause reduced TSH secretion by impairing the production or transport of TRH to the pituitary gland. Hypothyroidism may occur because the pituitary secretes TSH in insufficient quantities, or secretes TSH with an abnormal glycosylation pattern which reduces the biologic activity of TSH 1,2,3 . Treatment with oral TRH restores the biologic activity of TSH, suggesting that deficient hypothalamic TRH release induces both quantitative and qualitative abnormalities of TSH secretion. TSH molecules with reduced biologic activity may retain their immunologic reactivity in TSH immunoassays, explaining the sometimes observed slightly increased values of serum TSH (up to 10 mU/l) in central hypothyroidism 18, 23 .

The term central hypothyroidism is preferred because it is not always possible to distinguish between hypothalamic and pituitary causes. Central hypothyroidism is also associated with a decreased nocturnal TSH surge (due to loss of the nocturnal increase in TSH pulse amplitude under preservation of the nighttime increase in TSH pulse frequency), which further hampers maintenance of a normal thyroid function 4,5 .

Central hypothyroidism is a relatively rare condition occurring about equally in both sexes. Congenital cases of central hypothyroidism are due to structural lesions like pituitary hypoplasia, midline defects and Rathke's pouch cysts, or to functional defects in TSH biosynthesis and release like loss-of-function' mutations in genes encoding for the TRH receptor 6 , the TSH-beta subunit 7,8 , pituitary-specific transcription factors ( POU1F1 , PROP1, LHX3, LHX4 or HESX1), and LEPR or IGSF1 9 . Familial hypothyroidism due to TSHβ gene mutations follows an autosomal mode of inheritance. The β-subunit (118 aa) heterodimerizes noncovalently with the α-subunit through a segment called the seat-belt (aa 88-105). The described mutations of the TSHβ gene hamper dimerization with the α-subunit and thereby the correct secretion of the mature TSH heterodimer: Q42X and Q29X introduce a premature stop codon resulting in a truncated TSHβ subunit, G29R is a nonsense mutation preventing dimer formation, and C105V, 114X is a frameshift mutation causing disruption of one of the two disulfide bridges stabilizing the seat belt region 7,8,19,20 . Plasma TSH levels are variable, the TSH response to TRH is impaired but PRL secretion is normal, and plasma glycoprotein hormone α-subunits are high 19 . Mutations in pituitary transcription factors like POU1F1 and PROP1 are associated with deficiencies of TSH, GH and PRL 9 . Loss-of-function mutations in the membrane glycoprotein IGSF1 cause an X-linked syndrome characterized by central hypothyroidism, hypoprolactinemia, delayed puberty, macroorchidism and increased body weight; it is hypothesized that central hypothyroidism in these cases is secondary to an associated impairment in pituitary TRH signalling 32,33 . Cases of central hypothyroidism in childhood are mostly caused by craniopharyngioma (TSH deficiency in 53%) or cranial irradiation for brain tumors like dysgerminoma (TSH deficiency in 6%) or hematological malignancies 24 . Prophylactic cranial irradiation of the central nervous system in children with acute lymphoblastic leukaemia did not have an adverse effect on thyroid function within a median follow-up time of 8 years 21 .

Central hypothyroidism in adults is most frequently due to pituitary macroadenomas and pituitary surgery or irradiation 22 . The occurrence of TSH deficiency occurs usually after loss of GH and gonadotropin secretion. Return to euthyroidism is sometimes observed after selective adenomectomy 10 . Radiotherapy of brain tumors or pituitary adenomas is followed by hypothyroidism in up to 65%; the onset of hypothyroidism may be seen many years after radiotherapy 11,12 . Less common causes of adult central hypothyroidism are head injury 13, 25 , ischemic necrosis due to postpartum hemorrhage (Sheehan's syndrome), pituitary apoplexy, infiltrative diseases, and lymphocytic hypophysitis 14 . Lymphocytic hypophysitis seems to be an autoimmune disease; it occurs predominantly in women, especially during and after pregnancy, and the clinical picture is characterized by a pituitary mass and hypopituitarism 26 . A systematic review of articles published between 2000 and 2007 reported frequencies of anterior hypopituitarism in adults in the chronic phase after traumatic brain injury or subarachnoid hemorrhage (27): TSH deficiency was observed in 5.9% (95% CI 1.3-10.5) after subarachnoid hemmorrhage, and in 4.1% (95% CI 2.9-5.7) after traumatic brain injury. In prospective studies after traumatic brain injury TSH deficiency was observed in 3.9%, 6.8%, 2.1% and 4.2% at the acute phase and after 3, 6 and 12 months respectively (27).

Dopamine infusion inhibits the release of TSH, which may decrease T4 production rate by 56% 15 . Supraphysiological amounts of endogenous or exogenous glucocorticoids also dampen the release of TSH, but give seldom rise to decreased serum T4 values. The same is true for treatment with long-acting somatostatin analogs. A transient decrease of TSH secretion can be observed after withdrawal of TSH-suppressive doses of L-thyroxine, which may last up to 6 weeks 16 .

A new and novel cause of iatrogenic central hypothyroidism is from the administration of the RXR-selective ligand, bexarotene (Targretin). This medication is highly effective in cutaneous T cell lymphoma, but as reported by Sherman et al, up to 70% of patients treated with daily doses > 300 mg/m 2 had symptoms and signs of hypothyroidism. This was associated with reduction of serum TSH to an average of 0.05 mU/l, and reduction of free T4 from 12.9 pmol/l to 5.8 pmol/l 17 . A single dose of bexarotene rapidly and significantly suppresses serum TSH in healthy subjects, without an effect on serum prolactin or cortisol, suggesting a specific effect on thyrotropes (28). In vitro studies have shown that activity of the TSHβ subunit gene promoter is suppressed by 9-cis-retinoic acid and bexarotene 17 , but other studies have not confirmed this 29 . Rexinoids may further increase thyroid hormone metabolism through deiodination, sulfation and possibly glucuronidation (30,31).The condition can be appropriately treated by administration of thyroid hormone. (17)

9.3.2 CHRONIC AUTOIMMUNE THYROIDITIS

Chronic autoimmune thyroiditis may eventually cause hypothyroidism, mainly via destruction of thyrocytes (see also Ch. 7 ). In goitrous autoimmune hypothyroidism (the classical variant originally described by Hashimoto) the histology of the thyroid gland is characterized by massive lymphocytic infiltration with formation of germinal centers and oxyphilic changes of thyrocytes. In atrophic myxedema fibrosis is predominant, next to lymphocytic infiltration. The diffuse Hashimoto goiter has a peculiar firm consistency like rubber; the goiter may regress with time but can persist in many cases 1 . In some instances the patient presents with an initial transient hyperthyroid stage, called Hashitoxicosis'. The term Hashimoto's disease is generally used to indicate auto-immune destruction of thyrocytes which may eventually result in hypothyroidism although many cases remain euthyroid (see also Ch. 8 ). The serological hallmark of Hashimoto's disease is the presence of high titers of thyroid peroxidase (TPO) autoantibodies, formerly known as thyroid microsomal antibodies. The opposite of Hashimoto's disease is Graves' disease characterized by the presence of TSH receptor stimulating antibodies resulting in hyperthyroidism. The two disease entities frequently overlap, and can be viewed as the opposite ends of a continuous spectrum of autoimmune thyroid disease. Indeed, many patients with Graves' disease have TPO antibodies, and some case reports mention classical features of Graves' disease like exophthalmos and pretibial myxedema in the presence of hypothyroidism without any previous thyrotoxicosis 2 . TSH receptor blocking antibodies do occur in Hashimoto's disease, contributing to thyroid atrophy and hypothyroidism; they are more prevalent in Japanese than in Caucasian patients 3,4 . TSH receptor antibodies in Hashimoto's disease are negatively correlated to serum FT4 and thyroid size 5 .

The clinical manifestation of Hashimoto's disease with respect to thyroid function and thyroid size depends on the net effect of the various immunologic effector mechanisms involved in chronic autoimmune thyroiditis. Genetic and environmental factors may modulate the expression of the disease (6). Autoimmune hypothyroidism is associated with a number of single nucleotide polymorphisms in susceptibility genes (HLA-DR3, CTLA-4, PTPN22, Tg) (7). The prevalence of Hashimoto’s thyroiditis is higher in regions with a high ambient iodine intake than in iodine-deficient areas 8,9,10 . Smoking to a certain extent protects against the development of TPO antibodies and overt autoimmune hypothyroidism (11,12,13).

9.3.3 REVERSIBLE AUTOIMMUNE HYPOTHYROIDISM

Chronic autoimmune thyroiditis. Conventional wisdom has it that ‘once hypothyroid means always hypothyroid'. Indeed, the vast majority of patients with hypothyroidism due to chronic autoimmune thyroiditis require life-long thyroxine replacement therapy, but spontaneous recovery does occur in about 5% 1 . Return to the euthyroid state is apparently more frequent in countries like Japan, where - at a high ambient iodine intake - restriction of dietary iodine alone may induce a remission 2 .Conditions that increase the likelihood of spontaneous recovery are the presence of a goiter, a relatively high thyroidal radioiodine uptake, and a preserved increase of T3 after the administration of TRH during thyroxine treatment 3,4,5 .The spontaneous evolution from hypothyroidism back to euthyroidism has been related to the disappearance of TSH receptor blocking antibodies 6 . Changes in the titers of co-existing TSH receptor blocking and stimulating antibodies explain the sometimes observed alternating course of hypothyroidism and hyperthyroidism in the same subject 7 .

Silent thyroiditis and postpartum thyroiditis. Silent or painless thyroiditis and postpartum thyroiditis are variant forms of chronic autoimmune thyroiditis. The autoimmune reaction causes a mainly T-cell mediated destructive thyroiditis, which however is self-limiting. The characteristic course of the disease is thus first a thyrotoxic stage due to the release of stored hormone from the disrupted follicles, followed by a hypothyroid stage during the recovery towards a normal thyroid architecture; usually euthyroidism is restored within a few months (see also Ch. 8 ). In many cases the disease remains unnoticed, as clinical symptoms and signs are mostly limited. In the postpartum period it is also quite natural to attribute emerging complaints - especially if they are nonspecific in nature - to the aftermath of pregnancy and the work load of having a baby. Postpartum thyroiditis is, however, a rather common event, with an incidence of 4-6% as evident from several population-based studies 8,9 . The incidence in type I diabetes mellitus is four times higher, up to 25% 10 . Postpartum thyroiditis can be predicted to a certain extent from the presence of TPO antibodies in the serum of pregnant women in the first trimester: a titer of >100 kU/l at 2 weeks has a positive predictive value of 0.50 and a negative predictive value of 0.98 in this respect 9 . The titer of TPO antibodies decreases in the second and third trimester, and increases again in the postpartum period . Women who have experienced postpartum thyroiditis, have a 40% risk to develop again postpartum thyroiditis after a following pregnancy. About 20-30% of women with postpartum thyroiditis will develop permanent hypothyroidism within 5 years; the risk is higher in women with high titers of TPO-antibodies 11 . A subset of women with postpartum thyroiditis experience only a thyrotoxic phase; they are less at risk for later development of hypothyroidism 12 . Maternal TPO antibodies are associated with depression in the postpartum period 13 and with impaired child development 14 . A low maternal FT4 concentration during early pregnancy is also associated with impaired psychomotor development in infancy 15,16 .

Cytokine-induced thyroiditis. Cytokines are heavily involved in immune reactions (see Ch. 7 ), and it is thus not surprising that treatment with pharmacological doses of cytokines may induce autoimmune diseases in susceptible subjects. Treatment with interleukin-2 or interferon-α (IFNα) of patients with malignant tumors or hepatitis B or C is causally related to the occurrence of TPO-antibodies and the development of abnormal thyroid function 17,18,19,20 . The incidence is about 5-20%. Two types of IFNα-induced hypothyroidism have been recognized: autoimmune and non-autoimmune (21). Interferon-α induced autoimmune hypothyroidism is characterized by the presence of TPO antibodies. Elevated TPO antibodies before start of IFNα therapy increases the risk (positive predictive value 68% for the development of overt autoimmune hypothyroidism) (22), but TPO antibodies may develop de novo during IFNα treatment in 10-40% (23) Interferon-α induced non-autoimmune hypothyroidism is a destructive thyroiditis: a self-limited inflammatory disorder of the thyroid gland, characterized by an early thyrotoxic phase caused by the release of preformed thyroid hormones, and a late hypothyroid phase with complete resolution in most cases. IFNα has many immune effects including activation of immune cells, switching the immune response to Th1 pathways, downregulation of Treg cells, induction of cytokine release, and induction of MHC I expression on thyroid cells, all likely involved in the pathogenesis of IFNα induced autoimmune hypothyroidism (although it remains less well understood why IFNα can also induce Graves’ hyperthyroidism) (23). IFNα also exerts a direct toxic effect on thyrocytes, possibly involved in IFNα induced non-autoimmune hypothyroidism (26). IFNα-induced thyroiditis is most common in patients with chronic hepatitis C. It has therefore been hypothesized that hepatitis C virus could trigger thyroiditis by infecting thyroid cells. Supporting this view is the finding that the hepatitis C virus E2 protein can bind to CD81 molecules on thyroid cells and provoke IL-8 secretion (24).The prevalence of autoimmune and non-autoimmune types is about similar. Treatment is with levothyroxine, with no need to stop IFNα therapy. L-T4 replacement requirements may increase if patients are treated with a second course of interferon, or may decrease or end after completion of the IFNα course (19,25). It is recommended to screen all patients before starting IFNα therapy (TSH, TPO-Ab). If TSH is normal and TPO antibodies negative, TSH monitoring every three months is recommended until completion of IFNα treatment. If TPO antibodies are present, TSH monitoring every two months might be useful (23).

9.3.4 POSTOPERATIVE AND POSTRADIATION HYPOTHYROIDISM

Surgery. An important cause of hypothyroidism is surgical removal of the gland. Up to 40 percent of patients who undergo thyroidectomy for Graves' disease develop hypothyroidism (1). Most patients become hypothyroid in the first year after surgery; immediate postoperative hypothyroidism may resolve spontaneously by 6 months. After the first year the cumulative incidence of hypothyroidism rises by 1-2% per year. The frequency of hypothyroidism depends on the zeal of the surgeon and on other factors, such as the function of the thyroid remnant or the presence of active thyroiditis. Its occurrence correlates with the presence of antibodies to thyroid antigens. Thus, progressive destruction of residual tissue by thyroiditis may be the pathogenic mechanism. Hypothyroidism after surgical removal of multinodular goiter is less common (about 15%). Myxedema occurs almost invariably after subtotal thyroidectomy for Hashimoto's thyroiditis and after removal of lingual thyroids.

Radioiodine. A leading cause of hypothyroidism is radioactive iodine (RAI) treatment of Graves' disease. The frequency with which hypothyroidism supervenes RAI therapy is dependent on multiple factors, the principal one being the dose of RAI administered. The incidence of hypothyroidism 10 years after treatment is reported as high as 70 percent 1 . Hypothyroidism frequently develops already in the first year after treatment (with spontaneous return to euthyroidism in some patients), but it may not be manifest until years later in others. Its cumulative occurrence after the first year continues to rise with 0.5-2% annually, and it has been suggested that virtually all patients treated in this way will eventually become hypothyroid. Various treatment schedules have been devised with the hope of diminishing the incidence of RAI-induced hypothyroidism 2,3 , but in general, a lower incidence of hypothyroidism is invariably associated with a higher prevalence of persistent thyrotoxicosis that requires retreatment 3,4 . Inadvertent administration of RAI during gestation may cause neonatal hypothyroidism when given to the mother during the last two trimesters and also occasionally in the first trimester of pregnancy 5 . Hypothyroidism occurs less often (6-13 %) after 131I treatment of toxic nodular goiter 6,7 .

External irradiation. Hypothyroidism may supervene after therapeutic irradiation of the neck for any of a number of malignant diseases. It is particularly common (25-50%) after irradiation for Hodgkins' and non-Hodgkins' lymphoma, especially when the thyroid has not been shielded during mantle field irradiation and when iodine-containing X-ray contrast agents have been used prior to radiotherapy 8 . External radiotherapy for head and neck cancer (e.g. laryngeal carcinoma) carries an actuarial risk of 15% for developing overt hypothyroidism three years after treatment 10 . Elevated TSH values are even more common, with a 5-year incidence rate of 48% in another study with a median follow-up of 4,4 years 11 . Total body irradiation with subsequent bone marrow transplantation for acute leukemia or aplastic anemia may cause (subclinical) hypothyroidism in about 25%, usually occurring after one year and transient in half of the patients 9 . Probably because of radiation damage, subclinical or overt hypothyroidism is common among surviving bone marrow transplant recipients: there is a greater risk among younger patients, and need for life-long surveillance (12).

9.3.5 INFILTRATIVE AND INFECTIOUS DISEASES

The production of hypothyroidism by infiltrative disease is mentioned for completeness, despite the rarity of these conditions. Among these rare causes of primary hypothyroidism are sarcoidosis, cystinosis 1 (up to 86% in adults), progressive systemic sclerosis and amyloidosis 2 . Hypothyroidism is a frequent sequela of invasive fibrous thyroiditis of Riedel, occurring in 30-40% of the patients.

Hypothyroidism due to infectious disease is equally rare (3). Infection of the thyroid gland is somewhat more frequent in immunocompromised patients and in subjects with pre-existent thyroid abnormalities. Hypothyroidism in the recovery phase of subacute thyroiditis of De Quervain - a condition most likely related to a previous viral infection- is in contrast a very common event (see Ch. 19 ).

9.3.6 CONGENITAL HYPOTHYROIDISM

Congenital hypothyroidism can be permanent or transient in nature. Transient cases might be caused by transplacental passage of TSH receptor blocking antibodies, or iodine excess. Permanent cases are caused by either loss of functional tissue (mostly thyroid dysgenesis), by functional defects in thyroid hormone biosynthesis (‘loss of function' mutations in genes encoding for the TSH-R, NIS, Tg, TPO, DUOX2 and its maturation factor DUOXA2, or DEHAL), or by thyroid hormone resistance (TRα and TRβ1 mutations). For full discussion: see Ch. 15 and 16 .

9.3.7 IODINE DEFICIENCY AND IODINE EXCESS

Hypothyroidism caused by iodine deficiency is discussed in Ch. 20 . It is remarkable that hypothyroidism can also be caused by iodine excess, a condition described in the literature as ‘iodide-induced myxedema'. It can be explained by autoregulatory mechanisms operative in the thyroid gland. Inorganic iodide in excess of daily doses of 500-1000 µg inhibits organification of iodide; this phenomenon is known as the Wolff-Chaikoff effect. Usually an escape from the Wolff-Chaikoff effect occurs after several weeks. An unidentified iodinated product of the organification process (presumably an iodinated lipid) seems to be involved, which inhibits thyroidal iodide transport: consequently, the intrathyroidal iodine concentration falls below the level required for inhibition of organification 1 . Failure to escape from the Wolff-Chaikoff effect may produce hypothyroidism and this occurs preferentially in subjects with pre-existent subtle organification defects. Indeed patients with chronic autoimmune thyroiditis, previous subacute or postpartum thyroiditis, or previous radioiodine or surgical therapy are prone to iodide-induced hypothyroidism 2 ,3 .

Sources of iodine excess are an iodine-rich diet (e.g. seaweed ) and iodine-containing drugs like potassium iodide, some vitamin preparations, kelp tablets, topical antiseptics, radiographic contrast agents, and amiodarone. Amiodarone contains 39% of iodine by weight; large quantities of iodine are released during the biotransformation of the drug, giving rise to a 45-60 times higher iodine exposure than the optimal daily iodine intake of 150-300 µg recommended by the WHO.Amiodarone-induced hypothyroidism occurs predominantly in the first 18 months of treatment, especially in females with pre-existent thyroid antibodies 4 . Its incidence is higher in regions with a high ambient iodine intake than in areas with a lower iodine intake (22% and 5% respectively) 5 .

Mild iodide fortification of salt in Denmark increased average urinary iodide from the 45-61ug/l range up to 86-93ug/l. This cautious iodization of salt was accompanied by a moderate increase in the baseline incidence rate of overt hypothyroidism (38/100,000/yr) by 20-30%. This occurred primarily in young and middle-aged subjects with previous moderate iodine deficiency (6).

9.3.8 DRUG-INDUCED HYPOTHYROIDISM

A variety of therapeutic drugs can lead to moderate or even severe hypothyroidism (see also Ch. 9.8.3 ). The common antithyroid drugs (carbimazole, methimazole, and propylthiouracil) if given in sufficient quantity will cause hypothyroidism. This is also theoretically possible with agents that can block the uptake of iodide by the thyroid, such as perchlorate or thiocyanate, although these are rarely given. In susceptible individuals, primarily those with a history of autoimmune thyroid disease such as Hashimoto's or Graves' disease or in patients who have had either radiation or surgical trauma to the thyroid gland, large doses of iodide can cause goitrous hypothyroidism 1,2 (see also Ch. 9.3.7 ). While this is now less common, since iodides are no longer given for chronic pulmonary disease and lipid-soluble contrast agents are no longer used in diagnostic procedures, the problem may arise with patients taking iodine supplements or natural foods with high iodine content. Lithium has similar effects to those of iodide; it inhibits thyroid hormone release as well as hormone synthesis 3 . While lithium-induced hypothyroidism is more common in patients with underlying autoimmune disease, it has been reported in individuals with apparently normal thyroid glands. Long-term treatment with lithium results in goiter in about 50%, in subclinical hypothyroidism in about 20%, and in overt hypothyroidism also in 20% 4 . There are a large number of organic compounds that may impair thyroid function. These include phenol derivatives such as resorcinol, benzoic acid compounds such as para-aminosalicylic acid, the oral sulfonylurea compounds, phenylbutazone, aminoglutethimide, and a number of other agents 5 . Industrial pollution with polychlorinated biphenyls can also cause goitrous hypothyroidism 6 . In workers exposed to perchlorate, serum TSH and thyroid volume were not affected (7). Also in healthy volunteers, a 6-month exposure to perchlorate at doses up to 3 mg/day had no effect on thyroid function (8). Environmental low-level perchlorate exposure was ubiquitous in pregnant women but did not affect thyroid function (9).

Tyrosine kinase inhibitors freqently affect thyroid gland function and thyroid hormone metabolism. Imatinib and motesanib therapy has no adverse thyroid effects when a normal thyroid gland is in situ, but may require an increase in the replacement dose of L-T4 in hypothyroid patients (see section 9.8.3). In contrast, sunitinib and sorafenib therapy (applied in gastrointestinal stromal tumors and renal cell carcinoma) gives rise to primary hypothyroidism in a high proportion of patients (10). In the first study on 42 euthyroid patients treated by sunitinib, 36% had persistent hypothyroidism requiring L-T4 treatment, 17% had TSH between 5 and 7 mU/L, and 10% had suppressed TSH levels (11). In a prospective study among 59 patients, 61% were found to have a transient or permanently elevated TSH, and 27% required L-T4 replacement (12). The probability of hypothyroidism increases with each time and each cycle of treatment. Serum TSH increases at the end of the ON phase and is near the mormal range at the end of the OFF phase, leading to intermittent hypothyroidism. After several treatment cycles a permanent hypothyroidism occurs. It is uncertain if thyroid function tests return to normal after definitive withdrawal of sunitinib therapy.How sunitinib reduce thyroid gland function, is incompletely understood. Inhibition of thyroid radioiodine uptake has been observed (13) but in vitro experiments showed no effect on sodium-iodide transporters in thyroid cells (14). Impairment of thyroid peroxidase activity is shown in vitro (15), but still to be confirmed in vivo. Destructive thyroiditis has also been proposed (11,16). Recent case studies report marked reduction of thyroid volume and blood flow during sunitinib (17,18). A unifying hypothesis is that sunitinib (inhibiting vascular endothelial growth factor receptors as a major mechanism of action on tumors) causes regresion of the thyroid vascular bed resulting in impaired thyroid function(19).Vasoconstriction of thyroid vessels could reduce glandular uptake of radioiodine. Reduced thyroid perfusion could cause apoptosis of thyroid cells, resulting in thyroiditis in some patients. Sorafenib therapy in patients with metastatic renal cell carcinoma was associated with TSH elevations in 18% after 2-4 months, and one quarter of them developed thyroglobulin antibodies (20). TSH may be suppressed before the development of elevated TSH levels, suggesting destructive thyroiditis (21). Hypothyroidism would persist after sorafenib withdrawal. Sorafenib also has anti-angiogenic effects. It has been postulated that thyroid toxicity is restricted to tyrosine kinase inhibitors targeting key kinase receptors in angiogenic pathways, but not other kinase receptors (22,23).

9.3.9. CONSUMPTIVE HYPOTHYROIDISM (MASSIVE INFANTILE HEMANGIOMA)

Severe hypothyroidism has been described in a few infants with massive hemangiomas, due to high levels of activity of type 3 iodothyronine deiodinase in the hemangioma tissue 1 . Type 3 deiodinase inactivates T4 by conversion into reverse T3 (explaining the paradoxically high serum rT3 concentrations in these hypothyroid patients), and T3 by conversion into 3,3’-diiodothyronine. The high level of expression of type 3 deiodinase is likely induced by growth factors. The infants have no evidence of thyroid gland disease, and their hypothyroidism is apparently caused by an increased rate of thyroid hormone degradation in extra-thyroidal tissues outstripping the rate of thyroid hormone production: a nice example of “consumptive” hypothyroidism. This type of “peripheral” hypothyroidism has also been observed in a young adult 3 , in an athyreotic adult on levothyroxine 5 , and in a 54-yr patient with a large malignant solitary fibrous tumor expressing functional type 3 deiodinase activity 4 Surgical removal of the tumor restores euthyroidism.

9.4 PATHOLOGY OF HYPOTHYROIDISM

The characteristic pathologic finding in hypothyroidism is a peculiar mucinous nonpitting edema (myxedema), which is most obvious in the dermis but can be present in many organs. The myxedema is due to accumulation of hyaluronic acid and other glycosaminoglycans in interstitial tissue; these hydrophilic molecules attract much water 1 . The deposits of glycosaminoglycans have been related to loss of the inhibitory effects of thyroid hormone on the synthesis of hyaluronate, fibronectin and collagen by fibroblasts 2,3 .

The skin is distinctly abnormal. There is hyperkeratotic plugging of sweat glands and hair follicles. The dermis is edematous, and the collagen fibers are separated, swollen, and frayed. Extracellular material that appears eosinophilic or basophilic in hematoxylin and eosin stains, or that appears pink (metachromatic) with toluidine blue, or takes the periodic acid-Schiff (PAS) stain for mucopolysaccharides is much increased in the dermis. A sparse mononuclear cell infiltrate may be found about the blood vessels.

Skeletal muscle cells are swollen and appear grossly to be pale and edematous. Frequently microscopic examination reveals no significant abnormality. Alternatively, the normal striations are lost, and degenerative foci are seen in the cells. The fibers are separated in these degenerative foci by accumulations of a basophilic, PAS-positive homogenous infiltrate. This infiltrate may appear as a semilunar deposit under the sarcolemma.

The heart may be dilated and hypertrophied. Interstitial edema and an increase in fibrous tissue are present. The individual muscle cells may show the same changes seen in skeletal muscle. The serous cavities may all contain abnormal amounts of fluid with a normal or high protein content. The liver may appear normal or may show evidence of edema. Central congestive fibrosis in the absence of congestive heart failure has been described. The mitochondria tend to be spherical and their limiting membranes smooth, whereas those of the liver in thyrotoxicosis vary in shape and have wrinkled outer membranes 4 . The skeleton may be unusually dense on radiographic examination. In children, bone maturation is usually retarded, and typical epiphyseal dysgenesis of hypothyroidism is present 5 . The brain may show atrophy of cells, gliosis, and foci of degeneration. Deposition of mucinous material and round bodies containing glycogen (neural myxedematous bodies) has been found in the cerebellum of patients with long-standing myxedema and ataxia 6 . In uncorrected congenital hypothyroidism , the brain retains infantile characteristics. There is neuronal hypoplasia, retarded myelination, and decreased vascularity (see Ch. 15 ). The blood vessels often show prominent atherosclerosis. Whether this condition is more severe than might be anticipated on the basis of the patient's age and sex remains an unsettled question. In the intestinal tract there is an accumulation of mast cells and interstitial mucoid material, especially near the basement membrane. The smooth muscle cells may show lesions similar to those seen in skeletal muscle. The mucosa of the stomach, small bowel, and large bowel may be atrophic. The rest of the gastrointestinal tract, especially the colon, may be very dilated (myxedema megacolon). The uterus typically has a proliferative or atrophic endometrium in premenopausal women.

The kidney is grossly normal. Light and electron microscopic studies of renal biopsy samples have demonstrated thickening of the glomerular and tubular basement membranes, proliferation of the endothelial and mesangial cells, intracellular inclusions, and extracellular deposition of amorphous material with characteristics of acid mucopolysaccharides 8,9 .

In the pituitary in primary myxedema there is an increase in a class of cells that can be identified by the iron-periodic, acid Schiff, or aldehyde fuchsin staining techniques 10 . These are referred to variously as gamma cells, sparsely granulated basophils, or amphophils. Presumably they are derived from basophilic cells or chromophobes and are active in secreting TSH. Acidophilic cells are decreased. Patients who are congenitally hypothyroid and those who are hypothyroid during childhood may develop pituitary fossa enlargement. Occasionally prolonged hypothyroidism leads to sella enlargement in the adolescent and adult, and pituitary tumors have been described 11 . In these glands acidophils are virtually absent. In pituitary hypothyroidism the pituitary may be replaced by fibrous and cystic structures, granulomas, or neoplasia. Occasionally hypothyroidism due to deficient TSH secretion occurs in patients having the empty sella syndrome or because of isolated TSH or TRH deficiency. The adrenals may be normal or their cortex may be atrophied. The combination of adrenal cortical atrophy and hypothyroidism is known as Schmidt's syndrome and is thought to be of autoimmune etiology. Bloodworth found clinical evidence for hypothyroidism in 9 of 35 patients with Addison's disease; in 8 there was fibrosis of thethyroid, with atrophy in 4. The adrenal medulla appeared normal 12 . The ovaries and parathyroids have shown no definite abnormalities. The testes may show Leydig cells with involutionary nucleus and cytoplasm, hyalinization, or involution of the tubular cells, and proliferation of intertubular connective tissue in hypothyroidism with onset before puberty. Onset after maturity, in one case, led to similar changes that were restricted to the tubules.

The pancreatic islets are usually normal, although hyperplasia was present in one of our autopsied cases.

9.5 SYSTEMIC MANIFESTATIONS OF HYPOTHYROIDISM

The clinical expression of thyroid hormone deficiency varies considerably between individuals, depending on the cause, duration and severity of the hypothyroid state. Characteristically, there is a slowing of physical and mental activity, and of many organ functions.

9.5.1 ENERGY AND NUTRIENT METABOLISM

Energy metabolism. Thyroid hormone deficiency slows metabolism, resulting in a decrease of resting energy expenditure, oxygen consumption, and utilization of substrates. Reduced thermogenesis is related to the characteristic cold intolerance of hypothyroid patients. Measurement of the resting energy expenditure is rarely performed nowadays. In patients with complete athyreosis it falls between 35 and 45 percent below normal. In Addison's disease, the BMR may fall to -25 or -30 percent and, in hypopituitarism to below - 50 percent. The failure to find a metabolic rate as low as - 35 percent, when the clear-cut picture of myxedema is present, is very unusual. The effect of thyroid hormone deficiency on appetite and energy intake is not precisely known but energy expenditure certainly decreases, leading to a slight net gain in energy stores. Body weight increases on average by 10% due to an increase if body fat and retention of water and salt. An increase of adipose tissue mass results in an increase of serum leptin, which mediates a decrease in energy intake while energy disposal increases, eventually leading to a reduction in adipose tissue mass. Interactions between leptin and thyroid hormone have thus attracted much interest , especially because prolonged fasting in rodents decreases leptin and inhibits the hypothalamic-pituitary-thyroid axis resulting in a fall of serum TSH and serum T4. In hypothyroid patients, an increase, no change, or a decrease in plasma leptin concentrations has been reported (1-4,48). Whether thyroid hormone regulates leptin secretion independent of body mass index and body fat, remains controversial. In one study, leptin concentrations expressed as standard deviation scores (Z-scores) from the mean value of female controls matched for body mass index and age, were lower in hypothyroid and higher in thyrotoxic women, whereas Z-scores did not deviate from the expected values after restoration of the euthyroid state 1 . Thyroid hormone apparently modulates serum leptin only to a small extent. Ghrelin, a gastric peptide that plays a role in appetite stimulation and energy balance, is elevated in hypothyroid patients in most studies with a return to normal after L-T4 treatment (4-6). . It appears that leptin is mainly involved in thyroid hormone effects on energy homeostasis, whereas ghrelin may serve a compensatory physiological role (9).Serum adiponectin and resistin concentrations are not changed in hypothyroidism relative to controls (3,7,8,49). Serum obestatin and visfatin are increased in hypothyroidism; visfatin levels had a direct relationship with insulin resistance and body mass index (50,51).

Protein metabolism. The effect of hypothyroidism on protein metabolism is complex, and its effect on the concentration of a given protein difficult to predict. In general, both the synthesis and the degradation of proteins are reduced, but hypothyroid patients are in positive nitrogen balance. Despite both a decrease in the rate of albumin synthesis and degradation, the total exchangeable albumin pool increases in myxedema 10 . The albumin is distributed in a much larger volume, suggesting enhanced permeability of capillary walls. A synthesis of thyroid hormone-responsive proteins is clearly reduced in the hypothyroid state, whereas that of proteins such as TSH or glycosaminoglycans may be increased under the same circumstances 11,12 . Comparative studies of protein translation by hepatic ribosomes from T3-treated hypothyroid rats show that the mRNA's from some proteins are increased and others are decreased. Most of these proteins have not been identified. Treatment of myxedema is accompanied by mobilization of extracellular protein and a marked but temporary negative nitrogen balance, reflecting the mobilization of extracellular protein 13 . In a later phase there is an increase in urinary potassium and phosphorus together with nitrogen in amounts suggesting that cellular protein is also being metabolized 14 .

Carbohydrate metabolism. Glucose is absorbed from the intestine at a slower rate than normal. Fasting plasma glucose and fasting insulin levels are mostly similar to control values (8,15) although sometimes slightly lower glucose and higher insulin values than normal are reported 16,17,18 . Glycosylated hemoglobin is normal (21). The occurrence of hypoglycemia in hypothyroid patients should alert the physician to concomitant diseases (e.g. hypopituitarism). The development of hypothyroidism in patients with insulin-dependent diabetes mellitus may require lowering of the insulin dose to counteract the decreased rate of insulin degradation (22).The oral glucose tolerance test usually shows no abnormalities but a peak value that remains elevated at 2 hours can be observed (15,16,17), probably related to slow gastric motility and delayed absorption. Insulin response to an oral glucose load is variable; sometimes it is higher than in controls (15,18). When sugar is given intravenously, the glucose disappearance rate is prolonged although the peak value is normal in magnitude and in time of occurrence; the insulin response is blunted and slightly delayed 19 . Their exists fair evidence that hypothyroidism is associated with some degree of insulin resistance. The HOMA index (Homeostasis Model of Assessment) reflects the insulin resistance in the fasting state (mainly insulin resistance in the liver), while the Matsuda index reflects insulin sensitivity in the postprandrial state (mainly insulin sensitivity in the peripheral tissues). The HOMA index was found to be normal (5,8,15) or increased (18, 20) in hypothyroid patients vs euthyroid controls, whereas the Matsuda index was decreased and correlated positively with serum FT4 (15,18). The data suggest that insulin resistance might be present in some patients in the fasting state, but more frequently in the postprandrial state. Several other studies point into the same direction. In isolated monocytes derived from hypothyroid patients, insulin-stimulated rates of glucose transport are decreased due to impaired translocation of GLUT4 glucose transporters on the plasma membrane 18 . Hypothyroid patients as compared to euthyroid controls, also have lower postprandrial glucose uptake in muscles and adipose tissue 15 . Euglycemic hyperinsulinemic clamp studies in hypothyroid patients show an increase in insulin sensitivity after restoration of the euthyroid state (21).

Lipid metabolism. Biosynthesis of fatty acids and lipolysis are reduced. Changes in serum lipids are listed in Table 9-4 . The lipid changes bear in general a reciprocal relationship to the level of thyroid activity.The increased serum cholesterol in hypothyroidism may represent an alteration in a substrate steady-state level caused by a transient proportionally greater retardation in degradation than in synthesis 23,24,25 . The increase of serum cholesterol is largely accounted for by an increase of LDL-cholesterol, which is cleared less efficiently from the circulation due to a decreased T3-dependent gene expressing of the hepatic LDL-receptor. there is also evidence that the increase of LDL-cholesterol is also mediated via non-LDL receptor pathways by inducing Cyp7a1 expression and stimulating the conversion and excretion of cholesterol as bile acids 26,27,28,29 ,52,53 .

Table 9-4. Changes in serum lipids in hypothyroidism
Total cholesterolLDL-cholesterolHDL 2 -cholesterolHDL 3 -cholesterolTriglycerides increaseincreasemodest increaseno changeno change or modest increase

Interestingly, the LDL particles of hypothyroid patients are also susceptible to increased oxidizability 30 . The increase of HDL2- but not of HDL3-cholesterol 31,32,33 is due to a diminished activity of cholesteryl ester transfer protein 34,35 and hepatic lipase (which is involved in the conversion of HDL2 to HDL3). The changes in plasma LDL-and HDL-cholesterol are related to changes in free thyroxine, not to polymorphisms in LDL receptor or cholesteryl ester transfer protein genes 36 . Serum levels of apolipoprotein B and AI are increased but apolipoprotein AII levels are not. The sometimes present modest increase of serum triglycerides has been related to a decreased lipoprotein lipase activity in adipose tissue, suggesting hypertriglyceridemia in hypothyroidism is caused by a decreased clearance by adipose tissue (15). Another study however suggests the combination of observed normal lipolysis, low lipid oxidation rates and high triglyceride concentrations is compatible with increased triglyceride synthesis (37).An oral lipid tolerance test indicates postprandial lipemia (defined as an increase of triglycerides by 80% or more) is more frequent in hypothyroid patients than controls(38). Free fatty acids concentrations in serum are mostly normal, but decreased and increased values have also been reported (37,39). Lipoprotein(a) levels are also found to be normal in most studies (31,33,40) Remnant particles in serum (reflecting chylomicron and VLDL remnants) are less effectively cleared in hypothyroidism 41,42 . Taken together, the changes in plasma lipids in hypothyroidism result in an atherogenic lipid profile, although this has been disputed (43). Several studies do indicate, however, increased oxidative stress in hypothyroid patients, as evident from higher levels of serum malondialdehyde and nitric oxide and lower levels of the anti-oxidant paraoxonase in serum relative to controls (44,45).The observed increased oxidative stress is independent of body mass index (46). In subcutaneous fat biopsies of hypothyroid patients the mRNA expression of uncoupling protein-2 (UCP2) is decreased; UCP2 mRNA was related to lipid oxidation rate, basal free fatty acids, and serum T3 (47). UCP2 is a determinant of fat oxidation pathways, and may be involved in changes in metabolic pathways in thyroid disease.

9.5.2 FACIES AND INTEGUMENT

In the Report on Myxedema there is a detailed analysis of the symptoms of 109 patients described as "cretinoid," "expressionless," "heavy," "apathetic," "masklike," "vacant," "stolid," "good-tempered," "blunted," and "large-featured." The face is expression less when at rest, but it is not masklike, as in Parkinson's disease. When spoken to, the person with myxedema usually responds with a smile, which spreads after a latent period very slowly over the face. The patient is good-tempered but not entirely apathetic. The face is not vacant, as that of psychopathic patient may be. The features (except for the tongue) are not large, as in acromegaly. The face is expressionless at rest, puffy, pale, and often with a yellowish or old ivory tint. It is seldom as puffy as the classic facies of chronic renal failure. The skin of the face is parchment-like. In spite of the swelling it may be traced with fine wrinkles, particularly in pituitary myxedema. The swelling sometimes gives it a round or moonlike appearance (Fig. 9-3) .

Figure 9-3 . (A) The classic torpid facies of severe myxedema in a man. The face appears puffy, and the eyelids are edematous. The skin is thickened and dry. (B) The facies in pituitary myxedema is often characterized by skin of normal thickness, covered by fine wrinkles. Puffiness is usually less than in primary myxedema. The eyelids are often edematous. The palpebral fissure may be narrwowed because of blepharoptosis, due to diminished tone of the sympathetic nervous fibers to Müller's levator palpebral superious muscle and is the opposite of the lid retraction seen in thyrotoxicosis. The modest measurable exophthalmos seen in some patients with myxedema is presumably related to accumulation of the same mucous edema in the orbit as is seen elsewhere. It is not progressive and carries no threat to vision, as in the ophthalmopathy of Graves' disease. The tongue is usually large, occasionally to the point of clumsiness. Sometimes a patient will complain of this problem. Sometimes it is smooth, as in pernicious anemia (of course, pernicious anemia may coexist). Patients do not usually complain of soreness of the tongue, as they may in pernicious anemia. When anemia is marked, the tongue may be pale, but more often it is red, in contrast to the pallid face.

The voice is husky, low-pitched, and coarse. The speech is deliberate and slow. Often there is difficulty in articulation. Certain words are stumbled over and slurred, much as they are during alcoholic intoxication. The enlargement of the tongue, and possibly some thickness of the lips, may be responsible. The hair, both of the head and elsewhere, is dry, brittle, and sparse, and lacks shine. It varies in texture from coarse to normal. Its growth is retarded and it falls out readily. The eyebrows often are practically gone. Their disappearance begins at the lateral margin, giving rise to Queen Anne's sign. It should be noted, however, that this sign is not uncommon in elderly euthyroid women. In men the beard becomes sparse, and its rate of growth becomes greatly retarded. Haircuts are necessary only at long intervals. A shave a week is sufficient. The scalp is dry and scaly.

The skin is cool as a result of decreased metabolism as well as cutaneous vasoconstriction. It is dry due to reduced secretion by sweat and sebaceous glands.Scaling is common but rarely assumes the appearance characteristic of ichthyosis. The tissues beneath it seem thick, but usually do not pit on pressure. In the lower extremities, pitting edema is not uncommon. Subcutaneous fat may be increased, with the formation of definite fat pads, especially above the clavicles, but is conspicuously absent in the more advanced form of the disease (myxedematous cachexia).Retardation in the rate of healing of surgical wounds and of ulcerations, such as leg ulcers, has been described in myxedema. The nails are thickened and brittle. These changes are probably dependent, as are those of skin and hair, on retardation in growth. Nails require paring only at greatly lengthened intervals.

The hands and feet have a broad appearance, due to thickening of subcutaneous tissue. However, there is no bony overgrowth, so that they bear no resemblance to the extremities in acromegaly. Unusual coldness of the arms and legs is sometimes a subject of complaint. The palms are cool and dry. The characteristic skin changes are due to an increased amount of normal glycosaminoglycans and protein. The glycosaminoglycans are demonstrated by metachromasia after staining with toluidine blue. An increased concentration of glycosaminoglycans, composed principally of hyaluronic acid and chondroitin sulfuric acid, occurs in histologically similar skin lesions found in hyperthyroidism (pretibial myxedema). This excess accumulation of normal intercellular material represents not only an alteration in steady-state equilibrium but an actual increase in the synthesis and accumulation of glycosaminoglycan 1 .The glycosaminoglycans are long-chain polymers of D-glucuronic acid and N-acetyl-D-glucosamine, forming hyaluronic acid, or of L-iduronic acid and N-acetyl-D-galactosamine sulfate, forming chondroitin sulfate B. They exist free and in ionic or covalent linkage to protein. These mucoproteins comprise part of the normal nonfibrillar intercellular matrix, the ground substance holding cells together. As they are characteristically hygroscopic, they presumably hold in bound form the nonpitting water comprising the mucous edema. The total amount of exchangeable sodium is increased in myxedema despite a slight reduction in its plasma concentration 2 .

The sodium is extravascular and probably in the interstitial spaces. The diuresis seen after giving thyroid hormone to a hypothyroid subject occurs coincidentally with a decrease in tissue metachromasia and a temporary negative nitrogen balance 3 , and with this condition the extravascular sodium is mobilized and excreted. Studies with human skin fibroblasts have suggested that thyroid hormone inhibits the synthesis of hyaluronate. The mechanism for this effect has not been identified, but the thyroid hormone levels required to produce it in vitro are in the physiologic range 1 , 4 . Although similar deposits of mucopolysaccharides are found in the orbits of patients with the ophthalmopathy of Graves' disease and in the areas of localized myxedema, this striking observation has unfortunately not provided any basic understanding of the phenomenon, either in this condition or in primary myxedema (5).

In summary, in hypothyroidism the skin is dry, pale, thick, and rough with scales, and it feels cold. Hashimoto’s thyroiditis is associated with vitiligo (RR 25.6 with 95% CI 13.3-44.2 in women, and 15.8 with 95% CI 0.40-85.2 in men) (6). Thyroid autoimmunity may also be associated with chronic urticaria (7,8).

9.5.3 NERVOUS SYSTEM

Studies using 31P nuclear magnetic resonance spectroscopy of the frontal lobe of adult hypothyroid patients report reversible alterations in phosphate metabolism, suggesting impairment of mitochondrial metabolism 1 . Thyroid hormone receptors are present in human brain. These and other findings indicate the adult human brain as a thyroid hormone responsive organ, and provide a biologic basis for the very prevalent neurologic and neurobehavioral symptoms in adult hypothyroid patients 2 ( Table 9-5 ).

Table 9-5. Neurologic and psychiatric manifestations of hypothyroidism.
NEUROLOGIC SYMPTOMS AND SIGNS

  • Headache
  • Paresthesias
  • Carpal tunnel syndrome
  • Cerebellar ataxia
  • Deafness: nerve or conduction type
  • Vertigo or tinnitus
  • Delayed relaxation of deep tendon reflexes
  • Sleep apnea
  • EEG: low-amplitude theta and delta waves
  • Prolonged evoked potentials
  • CSF: elevated protein concentration

COGNITIVE FUNCTIONS

  • Reduced attention span
  • Memory deficits
  • Calculation difficulties

PSYCHIATRIC SYNDROMES

  • Myxedema madness (akinetic or agitated schizoid or affective psychoses)
  • Depression

Neurologic symptoms and signs. We are aware of no characteristic motor phenomena other than those due to weakness and to syndromes that seem to represent cerebellar dysfunction. A tendency to poor coordination was noted originally by the Myxoedema Commission. Jellinek and Kelly 3 described a series of myxedematous patients with ataxia, intention tremor, nystagmus, and dysdiadochokinesia. Ataxia has been noted in 8 percent of a large series of hypothyroid patients 4 . The delayed relaxation phase of the deep tendon reflexes is a well-known manifestation. Patients may have intention tremor, nystagmus, and an inability to make rapid alternating movements. In fact, this inability has long been used as a test for myxedema. The cause of this syndrome is not apparent, although deposition of mucinous material in the cerebellar tissue may be of pathogenetic importance. Whatever the cause is, it is important that these symptoms show a prompt and definite decrease after replacement therapy with thyroid hormone 5 . Sensory phenomena are common. Numbness, tingling, and painful paresthesias are frequent 6 and are especially common in hypothyroidism after surgery or 131I therapy. Paresthesias were present in 79 percent of one series of 109 patients. A metachromatic infiltrate has been found in the lateral femoral cutaneous nerve and sural nerve, together with axon cylinder degeneration 7 . Nerve conduction time is usually normal. Murray and Simpson 8 found that in some hypothyroid patients signs of median nerve pressure were present, apparently because of encroachment on the nerve by myxedematous infiltrates in the carpal tunnel 9,10 . A recent study reports carpal tunnel syndrome in 29% and signs of sensorimotor axonal neuropathy in 42% 22 . Deafness is a very characteristic and troublesome symptom of hypothyroidism. Both nerve and conduction deafness and combinations of the two have been reported, and vestibular abnormalities have also been demonstrated 37 . Serous otitis media is not uncommon. Two-thirds of patients complain of dizziness, vertigo, or tinnitus occasionally: these problems again suggest damage to the eighth nerve or labyrinth, or possibly to the cerebellum. Whatever type of deafness is present, there is marked improvement after thyroid therapy. Acute thyroxine depletion caused by total thyroidectomy has no deleterious effects on hearing up to 6 weeks 11 . Acquired hearing loss in association with adult-onset hypothyroidism should be distinguished from the sensorineural deafness of Pendred's syndrome. In the latter, treatment of hypothyroidism does not correct the hearing defect. Night blindness is not uncommon. It is caused by a deficiency in the pigment retinene, which is required for the adaptation to dark. Uncorrected deficiency of thyroid hormone during neonatal life causes not only more profound neurologic abnormalities but also irreversible damage (see Ch. 15 ). Hashimoto’s encephalopathy is a condition in which otherwise unexplained central nervous system dysfunction is observed in patients with Hashimoto’s disease and positive TPO-antibodies. The condition responds to glucocorticoids.A causal relation to thyroid autoimmunity is believed probable, but remains uncertain 26,27 . EEG abnormalities can be present, again depending on the severity and duration of the hypothyroidism. There may be absence of alpha waves and presence of low-amplitude theta and delta waves. Visual and auditory evoked potentials may be delayed as a consequence of abnormal cerebral cortical metabolism. Sleep apnea is not uncommon 15 . It has been difficult to assign a causal role for the myopathy versus the coexistent obesity in some of the reported cases. However, the muscular dysfunction may extend to the diaphragm and intercostal muscles, thus impairing the ventilatory mechanism.

Mental Symptoms. The mental picture in patients with overt hypothyroidism usually is one of extreme complacency. Memory is undoubtedly impaired, and attention and the desire to think are reduced . The emotional level seems definitely low, and irritability is decreased. Except in the terminal stage, reasoning power is preserved. Questions are answered intelligently, but slowly and without enthusiasm, and often with evidence of amusement. In a minority of patients, nervousness and apprehension are present. Cognitive tests of patients with moderate to severe hypothyroidism indicate difficulties in performing calculations, recent memory loss, reduced attention span, and slow reaction time 14,28 . Failing memory correlates inversely with serum T3 and T4 (23). Hypothyroidism may give rise rarely to reversible dementia, associated with reversible cerebral hypoperfusion (24). Recent studies indicate that hypothyroid-related memory deficits are not attributable to attentional deficit but rather to specific retrieval deficits (29). Hypothyroid patients showed prolongation of latencies only in the early ERP (event-related potential) components compared to controls, with speeding of sensory and cognitive processing after treatment (30).The cognitive impairment in hypothyroidism seems to be predominantly mnemonic in nature, possibly reflecting a specific defect in hippocampal memory (31). Imaging studies (functional MRI) have linked poorer memory states to specific brain areas and to reduced hippocampal volume 38,39 .

Psychiatric syndromes. The typical somnolence of severe hypothyroidism may suggest the psychiatric diagnosis of depression or dementia 16 . Patients are generally akinetic, though isolated case reports appear of patients who became hypomanic and agitated or garrulous (myxedema wit) as manifestations of this condition. Psychosis with hallucinations may occur (myxedema madness) 32 . Depression is so often associated with hypothyroidism that thyroid function tests should be performed in the evaluation of any patient presenting with this symptom. Central 5-hydroxytryptamine activity is reduced in hypothyroid patients 12 , and T3 supplementation might increase the efficacy of antidepressant drugs 13 although large randomized clinical trials in patients with major depression have produced conflicting results (32,33). At times, the depression in hypothyroidism is more severe than any of the other clinical manifestations of the disease. Because hypothyroidism is so readily treated, it is an especially important cause to eliminate in any patient with major depression. If the condition is due to hypothyroidism, it will resolve with time and appropriate treatment 17,18 . Patients hospitalized with hypothyroidism have a greater risk of readmission with depression or bipolar disorder than control patients 35 .

Cerebral blood flow, oxygen consumption, and glucose consumption have been reported to be diminished in proportion to the drop in metabolism in the rest of the body 19 , but older studies found unaltered glucose and oxygen use by the brain in either hypo- or hyperthyroid animals or humans 20 . In one study, cerebral cortical perfusion was little changed with treatment, but there was a decided fall in cerebrovascular resistance 21 . More recent studies indicate a generalized decrease in regional cerebral blood flow of 24% and in cerebral glucose metabolism of 12%, indicating that brain activity is globally reduced in severe hypothyroidism without the regional modifications usually observed in primary depression 25 . In 2009 neuropsychiatric symptoms of hypothyroid patients were studied in relation to changes in relative regional cerebral glucose metabolism after L-T4 treatment (36). Reduction of behavioral complaints during L-T4 therapy was associated with restoration of metabolic activity in brain areas that are integral to the regulation of affect and recognition.The findings suggest that thyroid hormone modulates regional glucose metabolism and psychiatric symptoms in the mature brain.. Experimental animal studies have shown that adult hypothyroidism in rats potentiates fear memory and also increases vulnerability to develop emotional memories. The findings further suggested that enhanced corticosterone signalling in the amygdala was involved in the pathophysiological mechanisms of fear memory potentiation 40 . Recent developments in brain imaging techniques thus provide novel insights in the relationship between hypothyroidism and mood disorders 41,42 .

9.5.4 CARDIOVASCULAR SYSTEM

Table 9-6. Cardiovascular manifestations of hypothyroidism.

PathophysiologySymptoms • reduced myocardial contractility• low cardiac output• increased peripheral vascular resistance• decreased blood volume• increased capillary permeability• dyspnea

• decreased exercise tolerance • angina

Signs • low pulse rate • diastolic hypertension • cardiomegaly • pericardial effusion • peripheral (non)pitting edema • low voltage ECG with conduction disturbances and nonspecific ST-T changes• prolonged systolic time intervals

Understanding of the cellular mechanisms of thyroid hormone action on the cardiovascular system has made it possible to explain to a large extent the decrease of cardiac output and cardiac contractility, the diastolic hypertension, the increased systemic vascular resistance, and the rhythm disturbances that result from hypothyroidism (Table 9-6) (37,38). Hypothyroidism decreases tissue thermogenesis by 5-8%, and increases resistance in peripheral arterioles through the direct effect of T3 on vascular smooth muscle cells. Diastolic blood pressure rises, and the afterload of the heart increases. Cardiac chronotropy and inotropy is reduced, resulting in a decrease of cardiac output to < 4.5 L/min. Thyroid hormone is an important regulator of cardiac gene expression, and many of the cardiac manifestations of hypothyroidism are associated with alterations in T3-mediated gene expression. T3 regulates positively sarcoplasmatic reticulum Ca2+-ATPase and negatively its inhibitor phospholamban, which together function in intracellular calcium cycling and thereby regulate diastolic function. The reduced expression of sarcoplasmatic reticulum Ca2+-ATPase and the increased expression of phospholamban in the hypothyroid heart explains the slowing of the isovolumic relaxation phase of diastolic function, typical for hypothyroidism.T3 regulates positively α-myosin heavy chain (the fast myosin with higher ATPase activity), and negatively β-myosin heavy chain (the slow myosin). In the hypothyroid heart, the expression of α-myosin heavy chain is decreased and of β-myosin heavy chain increased. T3 positively regulates the ion channels sodium potassium ATPase (Na+,K+-ATPase), the voltage-gated potassium channels (Kv1.5, Kv4.2, Kv4.3), whereas T3 negatively regulates the sodium-calcium exchanger (Na+/Ca2+exchanger).Together these channels coordinate the electrochemical responses of the myocardium. T3 positively regulates the β1-adrenergic receptors. The pacemaker-related genes, hyperpolarization-activated cyclic-nucleotide-gated channels 2 and 4, are transcriptionally regulated by thyroid hormone.

Systemic changes. Pulse rate and stroke volume are diminished in hypothyroidism, and cardiac output is accordingly decreased, often to one-half the normal value 1 . Myocardial contractility is reduced, but there is also a steep decline in the circulatory load, so that the circulation rarely fails until very late in the disease 2 . The speed of shortening is slowed, but the total force is not much modified. 3 . Myocardial adenyl cyclase levels are reduced 4 . The decrease in pulse rate occurs more or less in parallel with that of the metabolism. Stroke volume is reduced more than pulse rate at any given level, and is therefore the major determinant of the low cardiac output. Since the reduction in cardiac output is usually proportional to the decreased oxygen consumption by the tissues, the arteriovenous (AV) oxygen difference is normal or may be slightly increased. Slow peripheral circulation, and therefore more complete extraction of oxygen, as well as anemia, may be responsible for the increased AV oxygen difference. Myocardial oxygen consumption is decreased, usually more than blood supply to the myocardium, so that angina is infrequent. In some patients a reduction in cardiac output greater than the decline in oxygen consumption indicates specific cardiac damage from the myxedema 5 .Venous pressure is normal, but peripheral resistance is increased. Restoration of the euthyroid state normalizes peripheral vascular resistance. Changes in peripheral vascular resistance are not related to plasma adrenomedullin, but altered atrial natriuretic peptide secretion and adrenergic tone may contribute 29,39 . Central arterial stiffness is increased in hypothyroidism 30 , and arterial blood pressure is often mildly increased. It varies widely, but diastolic hypertension is usually restored to normal after treatment 6,7,31 .

Cardiomegaly. The heart in hypothyroidism has been a focus of much controversy. The term Myxodemherz was introduced by Zondek in 1918 8 . It embraced dilatation of the left and right sides of the heart, slow, indolent heart action with normal blood pressure, and lowering of the P and T waves of the electrocardiogram. Zondek found that after treatment with thyroid hormone there was a return of the dilated heart to somewhere near normal size, a more rapid pulse without change in blood pressure, and gradual return of the P and T waves to normal. These findings have been confirmed and extended. Indeed, occasional severely hypothyroid patients without underlying heart disease have congestive heart failure or low cardiac output reversed by thyroid hormone administration 7,9,10 . Therefore, congestive heart failure or impaired cardiac output relative to metabolic needs can be caused by hypothyroidism.Patients with untreated hypothyroidism are indeed at increased risk of heart failure 40 . Microscopic examination discloses myxedematous changes of the myocardial fibers.The cause of the cardiac enlargement has been disputed. Clearly, it is not due to hypertrophy alone, since it would not disappear so rapidly with treatment. One factor may be a decrease in contractility of the heart muscle. This decrease would require a lengthening of muscle fibers in order to perform the required work. Disappearance of interstitial fluid alone could account for only part of the observed shrinkage. Since the treatment of myxedema restores the hypothyroid heart to normal, there is apparently little permanent structural damage 9,10 . Cardiac glycosides will not improve the function of the heart in uncomplicated myxedema. Although the drug is efficacious if heart failure has been produced by coincident organic disease, myxedematous patients with coincident heart disease and congestive heart failure may tolerated digoxin poorly, just as they do morphine. This poor tolerance probably represents delayed metabolism, rather than myocardial sensitivity to the drug. The plasma concentration of digoxin is higher than in the normal subject at the same dose level, and smaller doses are required. When the heart in myxedema does not return to a normal size under thyroid hormone administration, hypertrophy due to some other disease is present as a complication. The return in size to normal under treatment is slow and progressive, requiring between 3 weeks and 10 months for completion.

Pericardial effusion. Gordon 11 long ago called attention to the occurrence of pericardial effusion in myxedema and explained the increase in the transverse diameter of the heart shadow on this basis. Effusion must frequently play a role in the increase in the size of the heart shadow, but it has amazingly little effect on cardiodynamics. The presence of fluid may be reflected in the right ventricular pressure contour, but tamponade, although reported, is rare 12,13 ,41 . Effusions of the pericardium, pleura, and peritoneum are common findings in hypothyroidism 14 . The protein of the effusion may be high or in the range of transudates. In 11 patients with tamponade studied, pericardial fluid protein ranged from 2.2 to 7.6 g/dl 12 . Occasionally, the fluid is high in cholesterol, with a "gold paint" appearance 13 . The hypothyroid heart responds normally to exercise 5,7 . Graettinger et al. 1 found that after exercise the low resting cardiac output increased normally with an increase of stroke volume and usually, of pulse rate. Their patients had slightly elevated resting pulmonary artery and right ventricular pressures and a diastolic dip in right ventricular pressure, all compatible with pericardial effusion. They doubt that myxedema alone can ever produce congestive heart failure, and believe that the recorded abnormalities represent not myocardial disease but pericardial effusion.

Electrocardiogram. The electrocardiogram reveals characteristic changes 5 , 7 , 10 , 15-19 . The rate is slow and the voltage is low. The T waves are flattened or inverted. Axis deviation, an increased P-R interval, and widened QRS complexes and prolonged QT interval are seen, but these signs are not diagnostic of myxedema. The pattern reverts toward normal with treatment, but the final pattern depends on the presence or absence of intrinsic myocardial disease. The rare occurrence of complete heart block complicated by Adams-Stokes attacks, with reversion to sinus rhythm after treatment with thyroid hormone, has been reported as has ventricular tachycardia 18,19 .Changes resembling those of ischemic heart disease may be found during exercise: they may indicate an intrinsic anoxia rather than organic narrowing of the coronary vessels 5 , 7 , 10 , 17 .The ECG changes have usually been attributed to the histologic changes in the myocardium. However, removal of pericardial fluid may immediately reverse the pattern toward normal suggesting that the effusion may in part be responsible for the abnormalities.The systolic time intervals are prolonged in hypothyroidism 5 , 7 , 10 , 20 . They can be measured by several techniques and have been expressed as the ratio of the pre-jection period and the left ventricular ejection time or the interval between the onset of the QRS complex of the ECG and the onset of the Korotkoff sound 21,22 . The most obvious effect of thyroid hormone deficiency on the heart is a lengthening of both systolic and early diastolic time characteristics. As evaluated by equilibrium radionuclide angiography, the time to peak emptying rate and the time to peak filling rate are longer in hypothyroid patients than in controls 23 ; the time intervals are negatively related to serum FT4 in the hypothyroid patients. The subtle decrease in early active relaxation and prolongation of contraction without major changes in global systolic function of hypothyroid patients is reversible upon thyroid hormone replacement therapy 24 .

Atherosclerosis. It is frequently suggested that accelerated atherosclerosis occurs in hypothyroidism 32 . Hypothyroidism accelerates atheromatous changes when these are induced experimentally in animals, but data in humans are not complete enough to justify this assertion. Most autopsied myxedematous subjects have severe atherosclerosis, but they are also usually 60 years or more of age. Arterial disease did not appear to be accelerated in patients rendered hypothyroid for therapy of angina pectoris or congestive heart failure 22 , but they have been observed over a relatively short period. Increased coronary arteriosclerosis is found in myxedematous patients with hypertension, but not if they are normotensive 25 (see further § 9.8.4). Nevertheless, the atherogenic profile of serum lipids and increased levels of homocysteine in hypothyroidism might well contribute to a higher prevalence of atherosclerosis in hypothyroid patients. Indeed , a 20-year follow-up study in the UK did observe a higher incidence of ischemic heart disease in subjects with subclinical hypothyroidism 35 . Another population-based study from The Netherlands found subclinical hypothyroidism to be an independent risk factor for aortic atherosclerosis and myocardial infarction; the attributable risk was comparable to that of other known risk factors for coronary artery disease 36 . A Danish nationwide register study found an excess risk of being diagnosed with cardiovascular disease, both before the diagnosis and following the diagnosis of hypothyroidism 42 .The topic is more fully discussed in § 9.10.4.

Angina pectoris. Occasionally angina pectoris is encountered in myxedema under two sets of circumstances. The less common is that in which angina or angina-like pain is present before treatment 26,27,28 . This generally indicates the presence of significant coronary artery disease since there is inadequate myocardial oxygenation despite reduced cardiac output and O2 utilization. Although improvement sometimes occurs with therapy 27 , this should not be undertaken until angiographic evaluation of the coronary arteries has been performed. Angina may also appear for the first time after therapy has been initiated, indicating that coronary flow is inadequate for resumption of normal cardiac function 26,27,28 . Again, this may indicate the presence of a structural lesion in coronary arteries.

9.5.5 RESPIRATORY SYSTEM

Dyspnea is a frequent complaint of myxedematous patients, but it is also a common symptom among well persons. Congestive heart failure of separate origin, pleural effusion, anemia, obesity, or pulmonary disease may be responsible. Some information on pulmonary function in hypothyroidism is available 1-7 . Wilson and Bedell 1 found a normal vital capacity and arterial PCO2 and PO2 in 16 patients. They also found a decreased maximal breathing capacity, decreased diffusion capacity, and decreased ventilatory response to carbon dioxide. Decreased ventilatory drive is present in about one-third of hypothyroid patients, and the response to hypoxia returns rapidly within a week after beginning therapy 6 .The severity of hypothyroidism parallels the incidence of impaired ventilatory drive. Weakness of the respiratory muscles has also been implicated as a cause of alveolar hypoventilation. Patients with myxedema may develop carbon dioxide retention, and carbon dioxide narcosis may be a cause of myxedema coma 3,4 .

Radiologic pulmonary abnormalities suggestive of fibrotic disease are associated with severe hypothyroidism, and may resolve with levothyroxine therapy (8).Myxedematous patients are more subject to respiratory infections.

Obstructive sleep apnea has been documented in hypothyroidism in about 7% and is reversible with therapy 5 , 7 . The prevalence of hypothyroidism in patients seen for snoring or obstructive sleep apnea syndrome is, however, no greater than that seen in the general population 9 . The same authors report little or no improvement in apnea symptoms upon thyroid hormone replacement therapy in the hypothyroid patients. Indeed, when patients are obese, cure rate of obstructive sleep apnea is not impressive. But replacement therapy is reported to be successful in most patients with obstructive sleep apnea, when they were grossly hypothyroid, generally nonobese and had e.g.reduction of macroglossia and goiter size (10). Obstructive sleep apnea in hypothyroidism seems to be caused by pharynx narrowing due to soft tissue infiltration by mucopolysaccharides and protein (10). Altered regulatory control of pharyngeal dilator mucles due to neuropathy may also be involved.

9.5.6 MUSCULOSKELETAL SYSTEM

Muscles. Muscle symptoms like myalgia, muscle weakness, stiffness, cramps and easy fatiguability are very prevalent in hypothyroid patients 23,24 . Weakness in one or more muscles groups is present in 38% as evident from manual muscle strength testing 22 .The symptoms are aggravated by exposure to cold. They are also prominent during the rapid onset of hypothyroidism after surgery or 131I therapy. Impairment of mitochondrial oxidative metabolism provides a biochemical substrate for these complaints, as evident from a rise in the ratio of inorganic phosphate to ATP in resting muscle and an important decrease in phosphocreatine in working hypothyroid muscle with a greater fall in intracellular pH than in controls 1,2 . Transition from fast type II to slow type I muscle fibers is involved in the change of muscle bioenergetics 3 , which is probably multifactorial. One patient with disabling muscle cramps was found to have reduced a-glucosidase activity in a muscle biopsy; after therapy with T4, the symptoms disappeared and the muscle enzyme activity returned to normal 4 . The electromyogram in myxedema may be normal or may demonstrate abnormalities distinct from those seen in myotonia or other muscle disease 5 . Polyphasic action potentials, hyperirritability, repetitive discharge, and low-voltage, short-duration motor unit potentials have been described. In the hypothyroid rat the rate of isometric relaxation is slow, and tension is less than in euthyroid or hypothyroid rat muscle at the same frequency of stimulation. Generalized muscular hypertrophy, accompanied by easy fatigue and slowness of movements, occurs in some cretins and myxedematous children or adults. It has been referred to as the Kocher-Debré-Sémélaigne syndrome in children 6 and as Hoffmann's syndrome in adults 7 . These patients do not have the classic electromyographic findings of myotonia. The myopathy of hypothyroidism is in some patients associated with weakness even though the muscles are hypertrophied.Chronic hypothyroid myopathy with increased muscular volume rarely cause entrapment syndromes 8 . Reflex contraction and relaxation time is prolonged mainly because of the intrinsic alterations in muscle contractility. Nerve conduction time may also be prolonged. Delayed reflex relaxation is characteristic and has been developed into a diagnostic test of thyroid function 9 . As with many other peripheral tissue function tests, there is considerable overlap between normal and mildly hypothyroid ranges. The rate-limiting step in muscle relaxation is the reuptake of calcium by the sarcoplasmic reticulum. In skeletal muscle, this process is dependent on the content of calcium ATPase. Recent studies have indicated that calcium ATPase activity of the fast twitch variety (SERCA-1) is markedly reduced in hypothyroidism 10 , and there is an accompanying impairment of calcium reuptake as a consequence. This occurs at a transcriptional level, since thyroid hormone response elements have been identified in the 5' flanking region of the SERCA-1 calcium ATPase gene 11 . The reduction in calcium ATPase would appear to explain one of the most obvious clinical manifestations of hypothyroidism, namely, delayed relaxation of the deep tendon reflexes. Skeletal muscle type 2 deiodinase activity is low and not influenced by hypothyroidism (25).

Table 9-7. Manifestations of hypothyroidism in the musculoskeletal system.
Clinical Symptoms and Signs• Myalgia, muscle weakness, stiffness, cramps, fatigue• Delayed reflex relaxation (e.g. prolonged Achilles tendon reflex relaxation time)• Arthralgias, joint stiffness, joint effusions• Carpal tunnel syndrome• Delayed linear bone growth in children
Laboratory• Normal ionized calcium, phosphate, and bone density• Increased serum PTH and 1,25 (OH) 2 -vitamin D 3, normal 25-OH vitamin D 3• Reduced urine calcium, hydroxyproline, serum alkaline phosphatase, osteocalcin, and IGF-1• Epiphyseal dysgenesis or delayed ossification in children

Joints. At the clinical level, patients with hypothyroidism may present with many manifestations, suggesting rheumatic disease such as arthralgias, joint stiffness, effusions, and carpal tunnel syndrome 12,13 . On the other hand, the symptoms may also suggest polymyalgia rheumatica, or primary myositis. The similarity of the symptoms of hypothyroidism to those of rheumatoid arthritis or osteoarthritis, especially when these are combined with the paresthesias of more severe hypothyroidism, should automatically lead to a consideration of hypothyroidism in any patient presenting with these symptoms. For example, in 5 to 10 percent of patients with carpal tunnel syndrome, primary hypothyroidism may be the cause, due to the accumulation of the hygroscopic glycosaminoglycan in the interstitial space with compression of the median nerve.

Bones. While calcium, phosphate, and bone density are generally normal in hypothyroidism, there is evidence of reduced bone turnover and resistance to the action of parathyroid hormone (PTH) 14-21 . Thus, serum (PTH) levels are elevated 16 . This is presumably the cause of the elevation in 1,25(OH)2-vitamin D3 19 . 25-OH-vitamin D3 levels are normal. The increase in PTH and vitamin D in turn increases calcium absorption. The reduction in glomerular filtration rate (GFR) and reduced bone turnover reduce urinary calcium and hydroxyproline levels and cause subnormal alkaline phosphatase, osteocalcin, and IGF-1 levels 15 . The alkaline phosphatase reduction is particularly important in children, in whom this enzyme is normally elevated due to bone growth. In children delayed linear growth or short stature are well-recognized signs suggesting the possibility of hypothyroidism. In addition, it is well recognized that epiphyseal dysgenesis and the delayed appearance of calcification centers are characteristic of hypothyroidism in infants and children. This subject is discussed in greater detail in Chapter 15 . Hypothyroidism is associated with enhanced susceptibility to fractures 26,27 .

9.5.7 GASTROINTESTINAL SYSTEM

The symptoms from the digestive system are essentially the expression of the slow rate at which the living machinery is turning over. Anorexia, which is common, can reasonably be interpreted as the reflection of a lowered food requirement, and constipation, which is frequently present, is the result of a lowered food intake and decreased peristaltic activity. Although two-third of patients have reported weight gain, it is of modest degree and due largely to the accumulation of fluid rather than fat. Contrary to popular belief, obesity is decidedly not a feature of hypothyroidism.

Complete achlorhydria occurs in more than half of myxedematous patients 1 . As many as 25 percent of patients with myxedema, like those with Hashimoto's thyroiditis, have circulating antibodies directed against the gastric parietal cells. This finding explains, at least in part, the frequency of achlorhydria and impaired absorption of vitamin B12. It is reported that up to 14 percent of patients with idiopathic myxedema have coincident pernicious anemia 2 .

Dysphagia or heartburn may be due to disordered esophageal motility 3 . Dyspepsia, nausea or vomiting may be due to delayed gastric emptying. Abdominal discomfort, flatulence and bloating occur in patients with small intestinal bacterial overgrowth: its prevalence (as demonstrated by a positive hydrogen glucose breath test) in hypothroid patients is rather high (54% vs 5% in a control group) (4). Bacterial overgrowth decontamination (by treatment with 1200 mg rifaximin each day for a week) is associated with improved gastrointestinal manifestations. Intestinal transit time is prolonged (5). Constipation may result from diminished motility, with the rare occurrence of fecal impaction. The syndrome of ileus may be seen occasionally 6 , and a megacolon may be evident on radiography 7,8 ; rarely pseudoobstruction develops. Intestinal absorption is slowed. Galactose and glucose tolerance curves show a delayed rise to a lower peak than normal and a delayed return to baseline. Xylose absorption is impaired 9 . Myxedematous ascites is rare 10 . Gallbladder motility is decreased, and the gallbladder may appear distended on x-ray examination 11,12 .

Symptoms or signs of disturbed liver or exocrine pancreatic function are usually not encountered, but chemical examination may suggest disease. Serum glutamine-oxaloacetic transaminase (GOT), lactate dehydrogenase (LDH), and CPK levels are elevated in patients with hypothyroidism 13,15 . The enzymes return to normal over 2 to 4 weeks during treatment. Urinary amylase levels may be increased. CEA levels are also increased and drop with therapy 14 . Serum liver enzyme activities ( alanine aminotransferase ALT and gamma-glutamyltransferase γGT) increase steadily across increasing TSH categories (also with TSH values within the reference range), ranging from mean values of 29 to 41 U/l for ALT and of 36 to 62 U/l for γGT 16 . At TSH levels >10 mU/l, ALT >40 U/l is observed in 24% and γGT >40 U/l in 30%.

Table 9-8. Gastrointestinal manifestations of hypothyroidism.

Symptoms
  • anorexia
  • gaseous distention
  • constipation
Signs
  • prolonged gastric emptying
  • prolonged intestinal transit time
  • slowed intestinal absorption
  • rarely ileus or ascites
  • gallbladder hypotonia
  • elevated liver enzymes and CEA

9.5.8 RENAL FUNCTION, WATER AND ELECTROLYTES

Hypothyroid patients tend to drink small amounts of water and to have diminished urinary output. Clinical evidence of renal failure is not often found, but laboratory examination may disclose certain departures from normal renal function; serum creatinine is raised by 10-20% , normalizing after L-T4 treatment 1,19,23 . Serum cystatin C is strongly influenced by thyroid function, and it may give erroneous results for assessing renal function in hypothyroid patients 19,23,24 .Because of decreased cardiac output and blood volume, renal blood flow is decreased, but it remains the same percentage of cardiac output. The glomerular filtration rate and effective renal plasma flow are decreased, but the filtration fraction is normal or variably altered 2,3,4 ,25 .

Hyponatremia sometimes occurs 26,27 .The response to water loading is variable. Moses et al. 5 reported that deficient diuresis after water loading is a sign of pituitary myxedema, but others, notably Crispell and co-workers 6 have found that severe primary myxedema may be associated with a delayed excretion that is not corrected by cortisone but rather by replacement with thyroid hormone. Perhaps the difference in opinion arised from interpretation of the normal response to water loading. This possibility is suggested by the data of Bleifer et al. 7 , who found a decrease in maximal diuresis in some patients with primary myxedema to below the normal lower limit of urine flow (3 ml/min), but not down to the 1 to 3 ml/min seen in panhypopituitarisn. The role of the antidiuretic hormone vasopressin (AVP) and of solute excretion in producing the decreased response to water loading was unknown. The defect was usually attributed to a decreased glomerular filtration rate, but in some patients inappropriately high levels of serum vasopressin have been demonstrated 8-12 . Since urinary hydroxycorticoid excretion is decreased, the adrenals might be incriminated as responsible for delayed water excretion. Other evidence, however, suggests (see below) that the tissue supply of adrenal cortical hormones is usually normal in myxedema. The diminished free water clearance in hypothyroidism occurs irrespective of the presence of hyponatremia. The inappropriate antiduresis in hypothyroidism was thus not fully understood, and a pure renal mechanism was hypothesized independent of vasopressin 18 . Better understanding has been obtained since the discovery of water channels, the aquaporins. Activation of the vasopressin V2 receptor on the basolateral membrane of the principal cells of the collecting ducts in the kidney by vasopressin (AVP) leads to upregulation of aquaporin-2 (AQP2). It involves shuttling the AQP2 containing vesicles from the cytoplasm to the apical membrane, leading to fast water transport across the bilipid apical membrane (20). Patients with advanced primary hypothyroidism may be hyponatriemic and fail to suppress plasma AVP with an acute water load. Advanced hypothyroidism is associated with a decrease in blood pressure, which is expected to activate baroreceptor-mediated non-osmotic AVP release, and indeed this is the case. In a rat model of hypothyroidism the increase in plasma AVP was associated with upregulation of AQP2 (21). A V2 receptor antagonist reverses the increased AQP2 and the impaired response to an acute water load. In contrast to diuretics, which enhance water and electrolyte excretion, the V2 receptor antagonists (vaptans) increase electrolyte-free water excretion.So far, the use of vaptans in severely hypothyroid patients with hyponatremia has not been reported 28 .

Occasionally, minimal proteinuria is seen. This condition could be due to congestive heart failure or to the increased capillary transudation of protein typical of hypothyroidism.

The total body sodium content is increased. The excessive sodium is presumably bound to extracellular mucopolysaccharides. In spite of reduced renal blood flow and blood volume, the sodium retention is probably not a reflection of altered renal function. In fact, salt loads are usually excreted readily and serum sodium concentrations tend to be low 13 , in contrast to other clinical situations associated with sodium retention, such as congestive heart failure 8 . The dilutional syndrome may be a result of inappropriate secretion of AVP 9-12 , but not in all patients. Thus, the dilutional syndrome in severe myxedema may be due to a resetting of the osmolar receptor, which causes water to be retained at a lower level of plasma osmotic pressure. The various changes in renal function may not return to normal at the same rate after treatment. Weight loss after therapy of hypothyroidism is mainly caused by excretion of excess body water associated with myxoedema 29 .

The serum uric acid level is elevated in hypothyroid men and postmenopausal women, apparently as a consequence of a decrease in renal blood flow characteristic of the disease 14 . No consistent changes in plasma potassium levels have been reported. Total magnesium levels may be elevated and the bound fraction and urinary excretion are reduced 15 . A modest hypocalcemia has been observed in some patients. The significance of low ANF concentrations in hypothyroidism is presently unclear 16 Plasma homocysteine concentrations are increased in hypothyroidism, related to lower folate levels and a lower creatinine clearance in thyroid hormone deficiency; restoration of the euthyroid state decreases plasma homocysteine levels into the normal range 1,17 .

In summary, the effects of hypothyroidism on the kidneys are: decreased glomerular filtration rate, decreased renal plasma flow, decreased sodium reabsorption, decreased renal ability to dilute urine, leading to increased serum creatinine and hyponatremia (22).

9.5.9 REPRODUCTIVE FUNCTION

Male gonads and reproduction 36 Alterations in androgens associated with hypothyroidism are rather complex and are due to the consequences of thyroid hormone deprivation on the production, serum transport and metabolic pathways of these steroids. Primary hypothyroidism results in a decrease in sex hormone binding globulin (SHBG) and thereby in total testosterone concentrations in serum; free testosterone is either normal or reduced (in approximately 60% of hypothyroid males) 1 . Serum estradiol is normal, but dehydroepiandrosterone (DHEA), DHEA sulfate, androstenediol and pregnenolone sulfate are decreased in serum of hypothyroid men compared with controls 2 . The metabolic clearance rate of androgens is decreased, and the conversion ratio of testosterone to androstenedione is reduced 3,4,5 . Serum LH and FSH are normal, but the LH and FSH response to GnRH is impaired. The testicles are histologically immature if hypothyroidism preceded puberty and show tubular involution if onset was after puberty 6 . In children, precocious testicular enlargement with early seminiferous tubular maturation has been reported 7 . This abnormality promptly subsides with the correction of the hypothyroid state, and is explained by spillover of the action of TRH on gonadotropes and of TSH on FSH receptors 8,9 . Libido in men may be decreased, and some men may be impotent. In one study among 14 adult hypothyroid males, a high prevalence was observed of hypoactive sexual desire, delayed ejaculation, and erectile dysfunction; there was significant improvement on L-T4 therapy 10 . In another study 84% of 44 hypothyroid patients had a score of 21 or less on the Sexual Health Inventory for Males (SHIM), compared with only 34% of 71 controls; L-T4 treatment increased SHIM scores and restored normal erectile function 11 . Little is known about the effects of hypothyroidism on human spermatogenesis and fertility. In one study reinduction of hypothyroidism did not lead to seminal changes, when compared with the same patients in the euthyroid state (12). In another small study it was reported that L-T4 treatment was associated with some improvement in sperm count and motility (13). More recently, male spermatogenesis was prospectively investigated in 25 hypothyroid men before and after L-T4 treatment by semen analysis, fructose and acid phosphatase measurements, teratozoospermia index, and acridine orange test (14). It was concluded that hypothyroidism had an adverse effect on human spermatogenesis, with sperm morphology the only parameter that was significantly affected.

Female gonads and reproduction 36 In hypothyroid women SHBG is decreased and thereby also total serum estradiol, estrone, and testosterone, but free estradiol and free testosterone are normal 1,15 . Metabolic clearance rate of estrogens is reduced, and peripheral aromatization and conversion of testosterone to Δ4-androstenedione are increased. LH and FSH are normal, but their responses to GnRH can be blunted or delayed. In children, hypothyroidism sometimes induces precocious puberty with menstruation and breast development 16 . These abnormalities disappear with correction of the hypothyroid state, and are explained by spillover of the action of TRH on gonadotropes and of TSH on FSH receptors 8,9 . The endometrium in premenopausal women is typically proliferative or, less common, atrophic. The proliferative endometrium and low urinary pregnanetriol levels suggest failure of LH production and of ovulation (17). Indeed the pulsatile gonadotropin release in the follicular phase is normal, but the ovulatory surge may not happen 18 . In adult premenopausal hypothyroid women, 77% have regular cycles and 23% irregular periods; corresponding figures in controls are 92% and 8% respectively 19 . Oligomenorrhea and hypermenorrhea/menorrhagia are the most common menstrual disturbances. Menorrhagia is probably due to estrogen breakthrough bleeding secondary to anovulation. Defects in hemostasis factors (see section 9.5.9.11) that occur in hypothyroidism may also contribute. Menstrual disturbances tend to be more severe in women with more severe hypothyroidism. Hypomenorrhea and amenorrhea are less common, at variance with the common belief that amenorrhea is the most frequent symptom. The lower frequency of menstrual abnormalities reported in more recent studies, compared with the older ones, may be attributed to delayed diagnosis and thereby more severe hypothyroidism in the earlier studies 20,21,22,23 .The amenorrhea-galactorrhea syndrome is occasionally found in adult hypothyroid women due to hyperprolactinemia; it is reversible with L-T4 treatment 24 . Although infertility may be a problem in either sex, the literature contains many reports of pregnancy in untreated myxoedematoous women 25,26 with frequent successful outcomes 27 . Studies examining the incidence of infertility in hypothyroid patients are scarce. No prospective controlled studies are available, whereas studies on the prevalence of hypothyroidism in patients presenting at specialized infertility clinics are subject to large selection bias. One study detected primary and secondary infertility in one (6.2%) of 16 overtly hypothyroid women, a frequency comparable to that found in control women 23 . Among 704 infertile women without previous thyroid disorders, 2.3% had increased serum TSH levels, comparable to that found in the general female population of reproductive age 28 . Nevertheless, a recent study shows that serum TSH levels are a significant predictor of fertilization failure in women undergoing in vitro fertilization 29 . Thus, although there is a known association between hypothyroidism and decreased fertility, hypothyroidism does not preclude the possibility to conceive. A particular study reports that 34% of hypothyroid women became pregnant without treatment: 11% of them had overt and 89% subclinical hypothyroidism 30 .

Pregnancy. When treatment has been started during pregnancy, more often than not a normal child is produced. Nevertheless, untreated hypothyroidism is associated with adverse outcomes for mother and child, as evident from many studies 31.32.33.34 . A critical review of all pertinent studies on the diagnosis and management of thyroid diseases during pregnancy and postpartum have been published in recently updated guidelines 35,37,38 . In short, when hypothyroid women become pregnant and maintain the pregnancy, they carry an increased risk for early and late obstetrical complications, such as increased prevalence of abortion, anemia, gestational hypertension, placental abruption, and postpartum hemorrhages. These complications are more frequent with overt than with subclinical hypothyroidism and, most importantly, adequate thyroxine treatment greatly decreases the risk of a poorer obstetrical outcome. Untreated maternal overt hypothyroidism is associated with adverse neonatal outcomes including premature birth, low birth weight, and neonatal respiratory distress. Though less frequent than with overt hypothyroidism, complications have also been described in newborns from mothers with subclinical hypothyroidism. Last but not least, there appears to be a significant increased risk of impairment in neuropsychological developmental indices, IQ scores, and school learning abilities in the offspring of hypothyroid mothers. The subject is more fully discussed in chapter 14 on Thyroid Dysfunction in the Pregnant Patient.

9.5.10 ENDOCRINE SYSTEM

Anterior pituitary. Thyrotroph hyperplasia caused by primary hypothyroidism may result in sellar enlargement, particularly when the condition has remained untreated for a long period of time 1,2 . Rarely, such hyperplasia may give rise to a pituitary macroadenoma that shrinks after thyroxine replacement 3,4 . The serum prolactin concentration is elevated in approximately one-third of patients with primary hypothyroidism 5 . The hyperprolactinemia is modest in degree and is rarely associated with galactorrhea 6 ,39 . When present, it subsides with thyroid hormone replacement in conjunction with the reduction in the serum prolactin level. Since thyroid hormone decreases the mRNA for pre-pro TRH in the paraventricular nuclei, it is conceivable that hypothyroidism leads to increased TRH secretion with consequent hyperprolactinemia. The growth hormone response to insulin-induced hypoglycemia is blunted in hypothyroidism 7 . Growth hormone secretion is decreased in hypothyroidism related to an increase in hypothalamic somatostatinergic tone 8 , resulting in low serum IGF-1 concentrations 9 . It may cause growth retardation in hypothyroid children. Serum IGF-2, IGFBP-1 and IGFBP-3 also fall, whereas IGFBP-2 rises; these changes are reversible upon treatment 10 . Another study reports slightly different results: IGF-1 and IGFBP-3 in hypothyroid patients indeed were lower than in healthy volunteers but did not change upon replacement therapy with levothyroxine, whereas the raised levels of IGFBP-1 in hypothyroidism decreased significantly after therapy 36 . The latest study on this issue reports substantial increases of serum IGF-1, IGFBP-3 and the acid-labile subunit (ALS) after L-T4 replacement in primary hypothyroidism (37). In patients with spontaneous autoimmune hypothyroidism due to Hashimoto’s thyroiditis who are adequately treated with levothyroxine, the distribution of IGF-I serum concentrations is similar to that of controls (38). Hashimoto’s thyroiditis can be associated with lymphocytic hypophysitis, which may cause growth hormone deficiency. The prevalence of GH deficiency in Hashimoto patients is low, in the order of 0.4% (38).

Adrenal cortex. Adrenal steroid hormone production and metabolism are considerably affected. Serum cortisol levels are normal, but the turnover time is slowed. This slowing is principally due to a decrease in the rate of cortisol oxidation as a result of reduced 11-β -hydroxysteroid dehydrogenase activity 11 . Conjugation with glucuronic acid in the liver is normal 12 . Reflecting these alterations, urinary 17-hydroxycorticoid (as well as 17-ketosteroid) excretion is reduced 11 , 13 . The turnover rate of aldosterone is also decreased in hypothyroidism 11 . This reduction is probably due to an alteration in steroid reductases that tends to diminish the proportion of androsterone formed and reciprocally increases the level of the etiocholanolone metabolite 14 . The serum concentration of aldosterone is low or normal 15 . Renin activity is also often reduced, as is the sensitivity to angiotensin II 16 . Adrenal responsiveness to adrenocorticotrophic hormone (ACTH) may be reduced, or the response may be delayed until the second and third days of the standard ACTH test, with an actual augmentation of the total response 17 . The adrenal glands often atrophy. Pituitary responsiveness to the metyrapone test has been variable. Normal but delayed peak response 18 , impaired response 19 , or even lack of response 19 has been reported. Grossly impaired responses to the stimulation with lysine-8-vasopressin and a delayed increase in serum cortisol levels after insulin-induced hypoglycemia have also been observed 20,21,22 . A general picture of adrenal function in the hypothyroid patient who is not under stress seems clear. Adrenal steroid metabolism and production decrease. The decreased production is accomplished automatically by the pituitary through decreased ACTH secretion. The result is a normal concentration of free cortisol in the serum. Presumably, sufficient hormone is produced for the reduced needs of the hypothyroid subject. Whether steroid production can be augmented sufficiently in times of stress is not clear, but the provocative test results suggest that these patients usually have a mildly impaired hypothalamic-pituitary adrenal axis 23,24 . The specific association of primary autoimmune hypothyroidism and primary autoimmune adrenal insufficiency is called Schmidt’s syndrome (25). It may be part of polyglandular autoimmune syndromes (26,27). Untreated primary adrenal insufficiency may slightly increase serum TSH, which returns to normal upon glucocorticoid replacement (28).

Sympathoadrenal system. The plasma concentration of norepinephrine in hypothyroid humans is elevated and returns to normal with L-T4 treatment (29,30). The epinephrine concentration is normal. Excretion of catecholamines in the urine is normal, but a decrease in urinary dopamine excretion has been described (31). The circulatory response to injected epinephrine decreases in hypothyroidism but returns rapidly to normal after small doses of levothyroxine (32,33,34). The increased central sympathetic output seems to be compensatory to a reduced response to catecholamines in target tissues (35). Mechanisms involved include a decreased number of β1-adrenergic receptors in the heart. The physiological and clinical implications of the interactions between the sympathoadrenal system (including the sympathetic nervous system and the adrenal medulla) and thyroid hormones have been reviewed, with specific attention of its value in cold adaptation and in states needing high-energy output 40 .

Adipose tissue: see § 9.5.1. Gonads: see § 9.5.9. Parathyroid glands: see § 9.5.6. Posterior Pituitary: see § 9.5.8.

9.5.11 HEMATOPOIETIC SYSTEM

Erythrocytes. In hypothyroidism, plasma volume and red blood cell (RBC) mass are both diminished, and blood volume is decreased. Anemia of mild degree is commonly present, and the hemoglobin level may be as low as 8 to 9 g/dl. In two reports on a large series of patients with hypothyroidism from various causes, the incidence of anemia ranged from 32 percent (1) to as high as 84 percent (2). The anemia may be a result of a specific depression of marrow that lacks thyroid hormone (3), or may be due to blood loss from menorrhagia, to decreased absorption because of gastric achlorhydria, to coincident true Addisonian pernicious anemia, to a decreased absorption of vitamin B12 (which has been found to occur in certain patients with myxedema as a result of diminished intrinsic factor), or to diminished production of erythropoietin by the kidney. The erythropoietic effect of thyroid hormone is mediated through erythropoietin (4). This substance increases RBC production by stimulating the erythroid differentiation of the bone marrow, and its secretion by the kidney appears to be related to the oxygen tension of the tissue. Anemia caused by hypothyroidism per se may be normocytic or macrocytic and respond to thyroid therapy. If iron deficiency develops from menorrhagia, a hypochromic and microcytic anemia may occur. This condition usually responds to iron alone, but may respond optimally only to combined iron and thyroid hormone (5). Hypothyroidism per se causes diminished blood cell formation probably as a response to decreased oxygen demand (6). Plasma and RBC iron turnover are decreased, and the bone marrow is frequently hypoplastic. The relationship between hypothyroidism and pernicious anemia has been well established. Patients have been reported who developed pernicious anemia while hypothyroid, and who lost their need for parenteral vitamin B12 when hypothyroidism was treated. It is also known that some hypothyroid patients absorb oral vitamin B12 poorly, and the defect is sometimes corrected by intrinsic factor (7,8). After thyroid therapy, the absorption defect may disappear or may persist (8). The incidence of pernicious anemia is higher than normal in myxedematous persons (5,8). In Tudhope and Wilson's series of 73 patients with spontaneous primary hypothyroidism, 12.3 percent had true Addison's anemia that responded to vitamin B12 (8). They believe that macrocytic anemia in hypothyroidism should not be accepted as a manifestation of thyroid hormone lack per se, but that it is due instead to the increased coincidence of Addison's anemia. Half of the patients with Addisonian anemia have serum antibodies against the thyroid gland and half of the patients with Hashimoto's thyroiditis have antibodies against gastric cell cytoplasm, parietal cells or intrinsic factor. Megaloblastic anemia due to folic acid deficiency has also been demonstrated in hypothyroidism. Reduced intestinal absorption secondary to hypothyroidism may be responsible for this deficiency, as suggested by the changes observed in a patient given tritiated pteroylglutamate before and after treatment with thyroid hormone (9). Also, a peculiar RBC abnormality has been described in patients with untreated hypothyroidism (10): a small number of irregularly contracted RBCs resembling burr cells are present. The significance of this condition, which may be reversed by the administration of thyroid hormone, is unknown.The erythrocyte sedimentation rate may be elevated in uncomplicated hypothyroidism 11 .

Leucocytes and thrombocytes. Granulocyte, lymphocyte and platelet counts are usually normal in hypothyroidism. Leukopenia might indicate associated vitamin B12 or folic acid deficiency. Hypothyroidism is associated with enhancement of phagocytosis, increased levels of reactive oxygen species (ROS) and increased expression of pro-inflammatory molecules like interleukin-1β in monocytes and macrophages 20 . Mean platelet volume is positively correlated with serum TSH in healthy subjects and in subclinical hypothyroidism, whereas the increase in overt hypothyroidism was insignificant 21,22,23 .

Hemostasis. Hypothyroid patients may have bleeding symptoms such as easy bruising, menorrhagia, or prolonged bleeding after tooth extraction. A systematic review concludes that coagulation tests indicate a hypocoagulable state in overt hypothyroidism (in contrast, hyperthyroidism is associated with a hypercoagulable state, predisposing to thrombosis) (12). Observed defects in general hemostatic tests are prolonged bleeding time, prolonged activated partial thromboplastin time, prolonged prothrombin time, and prolonged clotting time (12,13). Coagulation tests in overt hypothyroidism reveal low or normal factor VIII activity, low von Willebrand factor antigen and activity, low or normal fibrinogen, and low ristocetin induced platelet agglutination (12,14). Fibrinolytic tests in overt hypothyroidism indicate a hyperfibrinolytic condition with reduced TAFIa (activated thrombin-activatable fibrinolysis inhibitor) dependent prolongation of clot-lysis time, despite unaltered TAFI levels (15). Acquired von Willebrand’s syndrome may be the main factor responsible for a bleeding diathesis in overtly hypothyroid patients with a prevalence of 33% (being moderately severe in 9% and mild in 23%) (16,24,25 Even if bleeding episodes are mainly mild and mucocutaneous, blood transfusion, drug administration, or surgical procedure may sometimes be required (16). Desmopressin rapidly reduces these abnormalities (17), and may be of value for the acute treatment of bleeding or as cover for surgery. Usually the clinical relevance of these abnormalities is limited, as illustrated by no excess blood loss or bleeding complications during and after surgery in a large series of hypothyroid patients (18). In patients with moderate hypothyroidism a hypofibrinolytic state has been found, which carries a risk of developing thrombosis (19). However, the concept of a hypercoagulable state in subclinical hypothyroidism is not supported by a systematic review of the existing literature (12,26,27).

9.6 COURSE OF THE DISEASE

Although technologically dated, one of the most charming and clear descriptions of a typical case of myxedema is that given by William M. Ord 1 in Allbutt's System of Medicine, published first in 1897. It is as follows:

THE PICTURE OF THE DISEASE

"Thirty years ago the writer of this article had occasion to investigate the case of a lady suffering from myxedema in a most definite form, and therefore offering complete opportunity of studying the symptoms and the relations of the disease. The patient, a lady of thirty-five, who had several children, presented an appearance suggestive of Bright's disease, yet, although she was greatly swollen on the whole of her body, on careful examination the swelling did not appear to be due to an ordinary dropsy. There was nowhere any pitting on pressure, and there was no albuminuria in the slightest amount. The diagnosis of chronic Bright's disease without albuminuria at first suggested itself, but on further examination many symptoms not known to be related with Bright's disease came under the eye. The face, very much swollen in all parts, was particularly swollen in the eyelids, upper and lower, in the lips, and in the alae nasi. There was a flush, very limited, over the malar bones, contrasting with a complete pallor over the orbital regions. The eyebrows were greatly raised by the effort to keep the lids apart. The skin of the face, and indeed of the whole body, was completely dry, rough and harsh to the touch; not exactly doughy, but giving a sensation of the loss of all elasticity or resilience. The hair was scanty, had no proper gloss, and was much broken. In the absence of all signs of visceral disease the condition of the nervous system was such as to attract much attention. The physiognomy was singularly placid at most times, less frequently heavy, with signs of somnolence, very rarely alert. In interviews the patient was imperturbably garrulous to a degree that could not fail to attract attention. For many minutes she would talk without cessation until obliged to stop and take a good breath. What she said was not altogether relevant, but it had to be said. All interrupting questions were disregarded. If, at the end of a small pause, she was asked to put her tongue, she ignored the request, but at the end of a varying time, when her breath became short, she would put out her tongue for a long time. She dealt in the same way with questions put to her in respect of the points raised by her statements. Her letters were frequent, voluminous, and, as regarded handwriting, very good. Her speech was slow and laboured. There was some difficulty in it, evidently due to the swelling of the lips, but was more than this: the words hung in a way that indicated nervous as well as physical difficulty, and inflexions of the voice were wanting. The tones of the voice were leathery, and suggested rather those of an automaton. The proper timbre was quite lost. Doubtless this was in part, again, due to obvious thickenings in the fauces and the larynx; but it did not in any way resemble the character of voice observed in ordinary swellings of those parts. Her temper was singularly equable, she was the most tender and solicitous of mothers, and in a long course of years during which she was under the writer's observation no word of unkindness or suspicion fell from her lips. Lethargy was an impressive part of her mental condition. Memory was slow, but correct. She thought slowly, performed all movements slowly, and was slow in receiving impressions. Her toilet, and she was no fashionable person, occupied hours. Her household duties could never be overtaken, and she had to seek assistance. Her gait presented a distinct ataxic quality. As her bulky body moved across a room, there occurred at each step forward a quiver running from the legs upwards, such as may be seen in people under the influence of great emotion, as in Lady Macbeth. This appeared to be due to a want of complete concert in the action of the flexors and extensors of the body, the flexors acting for the most part in advance. The interval between the action of the two sets of muscles was at some times extreme enough to determine falls, not in any way produced by obstacles. She fell forward on her knees, and, as a result, she sustained fracture of the patella on one side, and the patellar tendon on the other. Similar conditions existing in the head and neck produced excessive distress. From time to time the head would fall forward in spite of all voluntary effort to prevent it. The chin would then rest on the upper part of the sternum, as is seen in cretins. Sometimes by prolonged exertion of the will, sometimes with the assistance of the hands, the head would be raised, not always to good effect; for unless great care were exercised the head would fall backwards with a suddenness that was alarming. There was no obvious defect of the sense of touch, but it must be admitted that the speed of the reception of tactile sensations was not noted. After the establishment of the disease she bore two children; on both occasions severe postpartum haemorrhage occurred. She had no other haemorrhages. The first impression was, as I said above, that the case was one of Bright's disease without albuminuria. The urine was examined regularly for years without detection of albumin, and there were no such changes in the heart and arteries as belong to Bright's disease. After ten years, however, albumin appeared in the urine, and the patient died ultimately with symptoms of contracting granular kidney. A postmortem examination could not be obtained, and therefore the condition of the thyroid gland and of the kidneys cannot be recorded."

ONSET OF THE DISEASE

The onset of naturally occurring hypothyroidism is insidious. The patient is often unaware of it, as may be friends and relations. As the gland is gradually replaced by fibrous tissue, lymphocytic infiltration, or both, the serum hormone levels and metabolic rate begin slowly to fall. The first symptoms may be a decrease in sweating and dislike of cold. They may be present alone for a period of years before dramatic events occur. One of our patients gave a story of marked hypersensitivity to cold for 12 years, at the end of which time the picture of full-blown myxedema developed. Sometimes the presenting symptom may be a demand for a warmer room or more clothing. Sometimes a mere decrease in activity due to listlessness, lack of energy, or fatigue, is the first change noted. In other patients, mental dullness or drowsiness may be observed. We have also seen the opposite change, namely, nervousness and irritability, or even peevishness in the exceptional case.

Progressive constipation or increase in menstrual flow may occasionally be the first event. So, too, may any of the following: deafness, falling hair, thick speech, dizziness, puffiness of the face, headache, pallor, weight gain, or fatigue. When hypothyroidism occurs more suddenly, as after surgical thyroidectomy or RAI therapy, the symptoms may not be so insidious, and indeed may be quite upsetting to the patient. Musculoskeletal symptoms such as frequent cramps may be distressing, and acute depression or acute anxiety may appear. Thus, the clinical course may be much influenced by the cause of the hypothyroid state. Obvious symptoms and signs usually appear as the thyroxine (T4) level falls below normal. Of these symptoms, nonpitting edema, from which myxedema derives its name, is pathognomonic. It is a specific thyroprival sign, and when it develops, the disease is in the full-blown state. There may be little apparent change in the patient's appearance or condition for several years. During such a period the patient may be well off subjectively. The increased sensitivity to cold can be met by maintaining the living area at an unduly warm temperature. The decreased energy makes the person content to do little or nothing. The myxedematous state is characterized by an amazing placidity. The terminal stage may be called myxedematous cachexia.

MYXEDEMATOUS CACHEXIA

Myxedematous cachexia is characterized by an intensification of all symptoms and signs. There is great thickening of the tongue, thickness, dryness and coarseness of the skin, thickening and brittleness of the nails, falling and brittleness of the hair, progressive decrease in activity and responsiveness, and a closer and closer approach to a purely vegetative existence. Although the mucous edema persists - and indeed tends to increase - body fat may disappear, so that actual wasting takes place. After this stage has persisted for an indefinite period of months or even years, death takes place because of intercurrent infection, congestive heart failure, or both. The final symptom is coma, which may last for days. In the untreated patient, the length of time between the first symptoms and death may be as long as 15 years. It is, fortunately, seldom nowadays that one witnesses the natural termination of the disease. It is seen only when the patient is already moribund when he or she comes to the physician, the diagnosis previously having been overlooked or where severe myxedema is present in association with another serious illness. In the Report on Myxoedema, which was published before the discovery of the cure of the disease, the duration is given as 10 years or more. The evolution of the symptoms of myxedema is slowly progressive. If one compares patients with myxedema of 3, 6 or 12 years' duration, although all may have classic symptoms and identical thyroid function test results, the clinical picture will be more intense at 6 years than at 3 and still more at 12. The mental manifestations, and the integumentary changes in particular, intensify as the years pass. Such severe manifestations of hypothyroidism are rarely seen in the current era. Patients and their friends and relatives are often strangely unaware of evidence of myxedema. Often patients are identified during treatment for some entirely unrelated disorder. Myxedema has been called a "consultant's diagnosis", because the changes that appear as the disease develops are so subtle and gradual that they are often overlooked by the patient's family physician. This fact is becoming less true with the ready availability of objective diagnostic tests.

9.7 DIAGNOSIS OF HYPOTHYROIDISM

Evaluation of a patient suspected of hypothyroidism starts with obtaining conclusive evidence that thyroid hormone deficiency is absent or present. Clinical examination suffices to provide a definitive answer in very severe cases of thyroid hormone deficiency, but is less accurate in mild cases. Biochemical proof of thyroid hormone deficiency is thus required in the vast majority of patients. If hypothyroidism is demonstrated, the next question to be answered is which disease entity has caused the hypothyroid state (nosological diagnosis). Delineation of the cause of hypothyroidism is relevant for identification of patients with potentially reversible hypothyroidism; it migh also give a clue for the existence of other conditions associated with a specific cause. The diagnostic process is schematically represented in Table 9-9 .

Table 9-9. Schematic approach of the patient suspected of hypothyroidism.

Stage 1 Is hypothyroidism present?A. Clinical assessment:B. Biochemical assessment: TSH and FT4 assays
Stage 2 If hypothyroidism is present, what is the cause?A. Clinical assessment: history, goiterB. Biochemical assessment: TPO antibodies; sometimes thyroid ultrasound

9.7.1 CLINICAL EVALUATION (STAGE 1A)

Table 9-10 lists the relative frequency of symptoms and signs accumulated by Lerman in a study of 77 myxedematous patients in one thyroid clinic and by Murray in a study of 100 patients with primary hypothyroidism, 15 pituitary patients, and 100 normal control subjects. This analysis identifies the cardinal manifestations of the disease. It also discloses that a certain number of manifestations are occasionally found in overt myxedema that are somewhat more suggestive of hyperthyroidism than of hypothyroidism. Under this heading may be listed dyspnea, nervousness, palpitations, precordial pain, loss of weight, and emotional instability. These symptoms are also found in normal control subjects in nearly the same frequency.

Many symptoms typical of primary hypothyroidism are not commonly found in pituitary hypothyroidism - for example, coarse skin, thick tongue, coarseness of hair, peripheral edema, hoarseness, and paresthesias.

Table 9-10. Incidence of symptoms and signs in hypothyroidism

Lermans’s Series

Murray’s Series

Symptoms and Signs

: Percent of 77
Cases of Primary
Hypothyroidism

Percent of 100
Cases of Primary
Hypothyroidism

Percent of 15
Cases of Pituitary
Hypothyroidism

Percent of 100
Normal Control
Subjects

Weakness
Dry skin
Coarse skin
Lethargy
Slow speech
Edema of eyelids
Sensation of cold
Decreased sweating
Cold skin
Thick tongue
Edema of face
Coarseness of hair
Cardiac enlargement (on x-ray film)
Pallor of skin
Impaired memory
Constipation
Gain in weight
Loss of hair
Pallor of lips
Dyspea
Peripheral edema
Hoarseness
Anorexia
Nervousness
Menorrhagia
a
Deafness
Palpitations
Poor heart sounds
Precordial pain
Poor vision
Fundus oculi changes
Dysmenorrhea
Los of Weight
Atrophic tongue
Emotional instability
Choking sensation
Fineness of hair
Cyanosis
Dysphagia
Brittle nails
Depression
Muscle weakness
Muscle pain
Joint pain
Paresthesia
Heat intolerance
Slow cerebration
Slow movements
Exophthalmos
Sparse eyebrows
99
97
97
91
91
90
89
89
83
82
79
76
68
67
66
61
59
57
57
55
55
52
45
35
32
30
31
30
25
24
20
18
13
12
11
9
9
7
3
--
--
--
--
--
--
--
--
--
--
--
98
79
70
85
56
86
95
68
80
60
95
75
--
50
65
54
76
41
50
72
57
74
40
51
33
40
23
--
16
--
--
--
9
--
--
--
--
--
--
41
60
61
36
29
56
2
49
73
11
81
100
47
7
80
67
40
93
80
60
20
53
40
--
87
67
33
47
13
--
73
0
33
--
53
--
26
13
--
7
--
--
--
26
--
--
--
--
--
--
13
73
73
13
26
13
0
67
60
0
80
21
26
10
17
7
28
39
17
33
17
27
43
--
14
31
10
36
21
--
52
2
18
15
42
--
15
20
--
9
--
--
--
23
--
--
--
--
--
--
20
41
21
17
24
15
12
9
14
4
58
aPremenopausal patients

 

The diagnosis of severe hypothyroidism is relatively straightforward on clinical grounds. All of the manifestations mentioned in the above discussion are present, and laboratory testing merely confirms the high index of clinical suspicion. However, severe hypothyroidism has become increasingly rare due to physicians' raised level of consciousness about the relatively high prevalence of this disease in women and the ease of making a laboratory diagnosis. Rather it is the more sublte or unusual presentations of hypothyroidism that may present difficulties 1 . Since laboratory confirmation of hypothyroidism is straightforward, the critical factor in successful diagnosis is maintaining a high degree of suspicion. If the diagnosis is not suspected in a patient with some of the typical manifestations of hypothyroidism at the first encounter, it may be several months before the physician reconsiders this explanation for the patient's complaints. Thus, hypothyroidism may be more readily diagnosed by a consultant who has not seen the patient before, since both the patient and the regular physician may have assumed that the many nonspecific symptoms are insignificant or at least unrelated to a specific organic disease.

There are certain symptoms or signs that should, irrespective of other factors, lead to a biochemical evaluation for possible hypothyroidism. In the child or adolescent, growth retardation is one of these. The presence of an enlarged thyroid should trigger a similar response. However, more subtle, less specific complaints, including depression or other organic mental syndromes, muscle cramps, paresthesias, carpal tunnel syndrome, hoarse voice, elevated cholesterol, pericardial effusion, arthritis, yellow skin (carotenemia), hyperkeratosis of the palms or soles, or menorraghia, can be manifestations of hypothyroidism. In addition, certain constellations of autoimmune disease occur in concert with hypothyroidism, including primary adrenal insufficiency, type I diabetes, and pernicious anemia. The presence of any of these should lead to search for primary thyroid dysfunction.

Statistical methods have been applied to the clinical diagnosis of hypothyroidism, based on the frequency of symptoms and signs in patients and controls. Well-known is the Billewicz score, composed of points given in a weighted manner for the presence or absence of 17 symptoms and signs 2 . Application of this score increases the pretest likelihood of hypothyroidism by 15-19% 3 . A newly developed clinical score is, however, easier to perform and more sensitive 4 ( Table 9-11 ).

Table 9-11 Accuracy of 12 symptoms and signs in the diagnosis of primary hypothyroidism 122 .

sensitivity

(%)

specificity

(%)

positive predictive value (%)

negative predictive value (%)

score if present

Symptoms

impairment of hearing

diminished sweating

constipation

paraesthesia

hoarseness

weight increase

dry skin

22

54

48

52

34

54

76

98

86

85

83

88

78

64

90

80

76

75

73

71

68

53

65

62

63

57

63

73

1

1

1

1

1

1

1

Physical signs

slow movements

periorbital puffiness

delayed ankle reflex

coarse skin

cold skin

 

36

60

77

60

50

 

99

96

94

81

80

 

97

94

92

76

71

 

61

71

80

67

62

 

1

1

1

1

1

Sum of all symptoms and signs present†

12§

 Add 1 point in women younger than 55 yr
§ Hypothyroid, 6 points; intermediate, 3-5 points; euthyroid, 2 points.

The positive predictive value of this new score for hypothyroidism is 96.9% at a score of 6 points or more; the negative predictive value for the exclusion of hypothyroidism is 94.2% at a score of 2 points or less. 62% of all overt hypothyroid and 24% of subclinical hypothyroid patients are classified as clinically hypothyroid by the new score, as opposed to 42% and 6% respectively by the Billewicz score Figure 9-4 . The diagnostic accuracy of these clinical scores is thus very low. In view of their poor performance, they should not be used for the diagnosis of hypothyroidism 12 .

Figure 9-4. Assessment of hypothyroidism by a clinical score, composed of 12 symptoms and signs as listed in Table 3 (Reproduced with persmission(4)).

Age and smoking have been recognized as modifiers of the clinical expression of thyroid hormone deficiency. Elderly patients have a smaller number of clinical signs than younger patients 5 . Smokers have more severe manifestations of hypothyroidism than nonsmokers 6 .

9.7.2 LABORATORY EVALUATION (STAGE 1B)

The assay of TSH in serum has proven to be the best single test for the exclusion or detection of hypothyroidism 12 . Using the flow-chart of Figure 9-5 the following results can be obtained:

Figure 9-5. Flow-diagram for the biochemical diagnosis of hypothyroidism.

(Figure 9-5) . Flow-diagram for the biochemical diagnosis of hypothyroidism.

1. TSH normal. Euthyroidism is almost certain, as primary hypothyroidism is excluded. However, two conditions will not be recognized. The first is the existence of central hypothyroidism. As isolated TSH deficiency is very rare, clinical examination of the patient will usually provide sufficient clues (symptoms and signs of a pituitary mass, of hypopituitarism, or of overproduction of pituitary hormones) to warrant further evaluation by a FT4 assay.The second is thyroid hormone resistance due to TRα mutations; this recently discovered and probably rare disease is characterized by low normal to slightyly low FT4 and high normal to slightly high T3 13,14 .

2. TSH elevated, FT4 decreases. This classical combination of test results indicates primary hypothyroidism. Test results are sometimes due to central hypothyroidism or nonthyroidal illness when TSH is slightly elevated (5-15 mU/l).

3. TSH elevated, FT4 normal. Test results indicate most often subclinical hypothyroidism, sometimes nonthyroidal illness.

4. TSH elevated, FT4 increased. A rarely encountered combination of test results, indicating either thyroid hormone resistance due to mutations in TRβ or TSH producing pituitary adenoma.

5. TSH decreased, FT4 decreased. Central hypothyroidism accounts for these test results, which, however, also can be observed in severe nonthyroidal illness and after recently instituted treatment for thyrotoxicosis (131I, surgery, antithyroid drugs) or recent discontinuation of excessive thyroid hormone medication.

6. TSH decreased, FT4 increased or normal. Hypothyroidism is excluded. Results indicate overt thyrotoxicosis or subclinical hyperthyroidism respectively.

The reference interval of serum TSH is about 0.4-4.0 mU/L. However, the lower normal limit is lower in pregnancy, and the upper normal limit increases with advancing age, reaching levels of 6.3 mU/L at the age of ≥80 yr 16. . Serum T3 should not be done for the diagnosis of hypothyroidism 12 .

9.7.3 NOSOLOGICAL DIAGNOSIS (STAGE 2)

The cause of the hypothyroid condition is in general easily established. Most informative are a careful clinical examination and determination of TPO antibodies in serum. Particularly relevant questions in the history taking are: family history of thyroid disease? recent delivery? previous thyroid surgery or 131I therapy? use of antithyroid drugs? exposure to iodine excess? Symptoms and signs of a pituitary mass or of hypopituitarism suggest the presence of central hypothyroidism. Physical examination may reveal a goiter (like the characteristic firm rubbery' goiter in goitrous Hashimoto's hypothyroidism), but many if not most hypothyroid patients have no palpable thyroid gland. High titers of TPO antibodies indicate chronic autoimmune thyroiditis, the most prevalent cause of hypothyroidism. Thyroid ultrasound can be helpful: the finding of a non-homogeneous hypoechogenic pattern indicates chronic autoimmune thyroiditis, which may be observed also in the absence of thyroid antibodies in serum 15 . Although most cases of hypothyroidism are permanent and require life-long treatment with thyroxine, a substantial minority is transient in nature due to the natural course of the underlying disease entity. Elimination of the causal factor is possible only in a few patients in whom the hypothyroid state is induced by antithyroid drugs or iodine excess. Table 9-12 provides the physician with possible clues for assessing the likelihood of reversible hypothyroidism in a particular patient. In selected cases further evaluation by thyroidal radioiodine uptake studies might be useful.

Table 9-12. Reversible causes of hypothyroidism

Etiology Frequency of reversibility Clues for potential reversibility
• chronic autoimmune thyroiditis about 5% 7 goiter 8 ; preserved thyroidal radioiodine uptake 9 ; preserved T3 response to TRH during thyroxine treatment 10
• postpartum thyroiditis up to 80% recent delivery; relatively low titers of TPO antibodies
• subacute thyroiditis almost 100% recent painful goiter
• postoperative and postradioiodinehypothyroidism not unusual thyroidectomy or 131 I therapy in previous 6 months
• iodine-induced myxedema high exposure to iodine excess; preserved thyroidal radioiodine uptake 11
• drug-induced hypothyroidism high exposure to antithyroid drugs or goitrogenic chemicals

9.8 TREATMENT OF HYPOTHYROIDISM

9.8.1 PHARMACOLOGY OF THYROID HORMONE REPLACEMENT PREPARATIONS

Levothyroxine.

L-thyroxine is prescribed as the sodium salt in order to enhance its absorption, which occurs along the entire small intestine 1,2 . Intestinal absorption of oral T4 is on average 80% 3 , and is greater in the fasting than in the fed state. Absorption is apparrently more complete and less eratic if the daily dose is taken in the fasting state 4 . Serum T4 concentrations peak 2 to 4 hours after an oral dose and remain above normal for approximately 6 hours in patients receiving daily replacement therapy 5,6 . The gradual conversion of T4 into T3 in various tissues increases serum T3 concentrations so slowly after thyroxine absorption that with daily levothyroxine administration, no significant changes in circulating free T3 are detectable. In North America, levothyroxine tablets are available in tablet strengths of 25, 50, 75, 88, 100, 112, 125, 137, 150, 175, 200 and 300 µg. The long half-life of thyroxine of about 7 days allows treatment with a singly daily tablet. Omission of an occasional tablet is of little relevance. If a patient misses a pill one day, they may take two the next. If they miss for two days, they may take three the next (7). Generic and brand-name levothyroxine preparations are mostly bioequivalent (8), but altered bioavailability has been reported due to changes in the formulation of preparations (9). Levothyroxine has a narrow therapeutic index with the potential for putting patients at risk for iatrogenic hyperthyroidism or hypothyroidism at doses only 25% less or greater than optimal, based on patient’s serum TSH (10). Available data suggest 15-29% of patients receiv.e inadequate doses of L-T4 and 18-24% receive excessive doses, based on serum TSH levels outside the reference range (11,12,46). It re-emphasizes the question whether the different marketed L-T4 formulations are mutually exchangeable: this would require bioequivalence between products. The FDA set criteria for testing bioequivalence of levothyroxine sodium tablets (13), and moved to approve generic L-T4 preparations as equivalent to branded preparations. This led to a joint statement by the American Thyroid Association, the Endocrine Society and the American Association of Clinical Endocrinologists in which they oppose the stand taken by the FDA (14). For pharmacokinetic studies designed to measure the bioavailability of L-T4 formulations, the FDA recommends that a single dose be administered to healthy subjects at a strength several times the normal therapeutic dose. With correction for the endogenous T4 pool, the method could distinguish the doses that differed by 25% and 33%, but not dosage strenghts that differed by 12.5% (15). The way the FDA proposes to measure the results of small changes in the T4 content of commercial preparations thus seems not very sensitive. Therefore it has been concluded that the type of T4 absorption studies the FDA uses as evidence, as opposed to serum TSH which thyroidologists rely on, is inappropriate (16). According to the FDA definition of bioequivalence, particular test formulations have found to be bioequivalent to the reference formulation of levothyroxine in healthy volunteers (17). However, a pharmacodynamic TSH-measurement bioequivalence protocol, using normal L-T4 replacement dosing in athyreotic volunteers, is likely to be more sensitive and safer than current FDA Guidance based on T4 pharmacokinetics (18). In the meantime the allowable potency range for L-T4 tablets has been tightened to 95-105%, a significant improvement of the original FDA-approved 90-110% range. A study comparing pH-dissolution profiles of selected commercial levothyroxine preparations, shows a considerable decrease in dissolution of L-T4 tablets with increase in pH, with differences in T4 dissolvement between the various preparations (19). Differential dissolution of T4 products can impact oral absorption and bioavailability of L-T4, and may result in bioequivalence problems. A survey among physicians in the USA managing patients with thyroid diseases that required the use of L-T4 preparations, indicated that 177 of 198 reports of adverse events associated with changes in TSH values, were associated with a change in the source of L-T4 (20). The exchanges were done by the patient’s pharmacy without the clinician’s knowledge in 92%. Fifty-four of the 198 reported cases resulted in serious adverse events, and 52 of these 54 cases were associated with a substitution of one L-T4 preparation with another. Against this background it is recommendable to continue to use the same L-T4 preparation once the appropriate L-T4 dose has been established.(27). The pharmacist should inform the clinician if a switch is being made to another L-T4 preparation.

Liothyronine.

After oral administration of L-triiodothyronine sodium (which is more readily absorbed than T4) peak levels of serum T3 are observed within 2 to 4 hours 21 .The serum T3 concentration may reach elevated values after a single dose of 50 µg or even 25 µg, sometimes associated with cardiac symptoms like palpitations 22 . The half-life of T3 is short (approximately one day) which requires several gifts per day.. Preparations of L-T3 can be useful in the short-term management of patients with thyroid cancer to shorten the period of hypothyroidism required for diagnosis and treatment of remaining tumor tissue with 131I, and in myxedema coma; they are not recommended for long-term replacement therapy in hypothyroidism. Pharmacodynamic equivalence of levothyroxine and liothyronine is achieved at a dose ratio of about 3:1 27

Desiccated thyroid.

Desiccated thyroid is prepared from porcine or bovine thyroid glands. In former days desiccated thyroid was standardized by the organic iodine content, which did not distinguish between iodotyrosines and iodothyronines 23 . Current guidelines stipulate that one grain (65 mg) of desiccated thyroid contains about 44 µg T4 and 9 µg of T3; the hormones are in the form of thyroglobulin 24,25 . In our experience, the biologic potency of a 1-grain desiccated thyroid tablet is about 75 to 88 µg T4. Because of the relatively high ratio of T3 to T4 in desiccated thyroid, patients receiving an amount of this medication adequate to normalize serum TSH generally have serum T4 concentrations in the lower half of the normal range. Serum T3 concentrations will vary in such patients, depending on the interval between ingestion of the medication and the time of blood sampling. The time course of the absorption of T3 is similar whether it is contained in thyroglobulin or free in the tablet, with peak levels approximately 2 to 4 hours after oral administration 21 .A recent randomized clinical trial compared levothyroxine replacement with desiccated thyroid extract (DTE, Armour Thyroid, of which each grain of 65 mg contained 38 μg L-T4 and 9 μg L-T3) 29 .Use of DTE relative to levothyroxine was associated with modest weight loss and greater patient preference; Serum T3 was higher and serum FT4 was lower during treatment with DTE than during levothyroxine treatment. Although DTE can provide satisfactory replacement therapy if TSH levels are maintained in the normal range, it is not recommended in current guidelines for treatment of hypothyroidism 27 .

Combinations of T3 and T4.

Liotrix, the only combination preparation currently available in the United States, contains 50 µg T4 and 12.5 µg T3/1 grain equivalent, but is biologically equivalent to a 65 mg (1 grain) tablet of desiccated thyroid. Recent studies in thyroidectomized rats have demonstrated that restoration of the euthyroid state in all tissues can only be restored by the combination of T4 and T3, and not by T4 alone 26 . This finding has aroused new interest in combinations of T3 and T4 in hypothyroid patients who are dissatisfied with the outcome of levothyroxine replacement therapy 33-37 (see 9.8.2).However, a meta-analysis of 11 randomized clinical trials found no evidence for superiority of L-T4 + L-T3 combination therapy over levothyroxine monotherapy: bodily pain, depression, anxiety,fatique and quality of life were not different between the two treatment modalities 30 . Some patients nevertheless do prefer combination therapy 31 , and it is hypothesized that particular genetic polymorphisms in thyroid hormone transporters and deiodinases are linked to this preference. Guidelines do not recommend combination therapy, and levothyroxine remains the standard treatment modality for hypothyroidism 27 . In patients with persistent complaints despite adequate levothyroxine replacement as evident from normal serum TSH levels, one may consider a trial of L-T4 + L-T3. This should be offered after exclusion of other conditions which might be responsible for the persistent complaints. Detailed instructions on selection of patients for combination therapy, calculation of levothyroxine and liothyronine doses, and monitoring are given in a recent guideline on this issue by the European Thyroid Association 32 ,

9.8.2 REPLACEMENT WITH THYROXINE

Of the available thyroid hormone replacement preparations, thyroxine is presently recommended as the drug of choice in view of its long half-life, ready quantitation in the blood, ease of absorption, and the availability of multiple tablet strenghts 1-4 .All guidelines state levothyroxine sodium is the standard treatment of hypothyroidism 51 .

Institution of therapy.

The rapidity with which normal thyroid hormone levels should be restored depends on a number of factors, including the age of patient, the duration and severity of the hypothyroidism, and the presence or absence of other disorders, particularly those of the cardiovascular system. Most patients under the age of 60 can immediately begin a complete replacement dose of 1.6 to 1.8 µg levothyroxine/kg ideal body weight (about 0.7 to 0.8 µg/1b). Requirements for children and infants are discussed separately and are higher than those for adults between the ages of 20 and 70. A randomized double-blind trial comparing a full starting dose (1.6 μg/kg L-T4) with a low starting dose (25 μg L-T4, increased every 4 weeks) in patients with newly diagnosed cardiac asymptomatic hypothyroidism observed that the full starting dose was safe and more convenient and cost-effective than a low starting dose regimen (47). Preference for gradual levothyroxine replacement and conservative dosage titration is, however, expressed widely (48). The standard advice is to take L-T4 tablets in the morning half an hour before breakfast. Interestingly, L-T4 taken at bedtime is associated with higher FT4 and T3 and lower TSH concentrations in serum compared to the same L-T4 dose taken in the morning, most likely due to better absorption of L-T4 during the night (49). Patients might be adviced to take their L-T4 tablets always either in the morning or in the late evening. Drinking espresso together while ingesting the L-T4 tablets, reduces intestinal absorpton of levothyroxine (50). The cause of hypothyroidism also influences replacement in that patients with total thyroidectomy or severe primary hypothyroidism have slightly higher requirements than do patients who become hypothyroid after radioiodine or surgical treatment for Graves' disease 5 . The latter group may have some residual thyroid function that is autonomous, and thus a complete replacement dose is excessive. For most women, a complete replacement dose will be between 100 and 150 µg per day and, for most men, between 125 and 200 µg per day. Pretreatment serum TSH predicts to a certain extent the daily maintenance dose of levothyroxine in patients with primary hypothyroidism (Figure 9-6) 6 . Individual L-T4 requirements are dependent on lean body mass. Age- and gender-related differences in L-T4 needs reflect different proportions of lean mass over the total body weight. An estimate of lean mass may be helpful to shorten the time  required to attain a stable dose of L-T4, particularly in subjects with high body mass index values that may be due either to increased muscular mass or obesity (7).

Figure 9-6. Relationship between the optimal daily dose of levothyroxine sodium and the mean pretreatment serum TSH concentration in patients with primary hypothyroidism. Simple linear regressions are shown for two subgroups calculated according to the daily dose of L-T4 divided at the median dose of 125 µg; the intercept of these two correlation lines occurs at the TSH concentration of 36 mU/l. (Reproduced with permission) 6 .

Full replacement doses should not be administered initially to patients over the age of 60, to patients who have a history of coronary artery disease, or to patients with long-standing severe hypothyroidism. While levothyroxine improves cardiac function in patients with hypothyroidism and increases cardiac output and decreases systemic vascular resistance and end-diastolic volume, it also increases myocardial oxygen consumption. Thus, while patients with coronary artery disease and angina may benefit from reversal of their hypothyroid state, to avoid precipitating acute myocardial ischemia, the dose should be titrated, starting with 25µg a day and increased by increments of 25 µg at 8-week-intervals until serum TSH falls to normal or symptoms of angina worsen or appear. A similar slow approach is prudent in patients with long-standing, severe hypothyroidism, also because occasionally psychosis or agitation occurs during the initial phase of replacement in such cases 8,9,10 .

Monitoring.

In the patient given what is thought to be a complete replacement dose of levothyroxine, a TSH and free T4 should be measured about 2 months after therapy begins to establish that the estimated dose is appropriate for the patient. At that time, serum TSH may be still elevated, indicating the need for a modest increase in dose, or TSH may be suppressed, indicating that a reduction is in order. This is usually done in 12.5- to 25-µg increments, depending on the patient 11 . These studies should be repeated again in 2 months to titrate proper dosage. After proper dosage has been achieved, the test should be repeated yet again after the patient has been euthyroid for approximately 6 months. This is because in certain patients, normalization of thyroxine clearance may require more than 8 weeks, and a dose of levothyroxine that is adequate when the patient is metabolizing thyroxine more slowly may be inadequate when the patient is euthyroid. This dose should be continued and monitored on an annual basis. In patients with severe primary hypothyroidism, few adjustments will be required after the initial titration until the eight decade. However, patients with Graves' disease who have had radioactive iodine may require dosage adjustments up to as long as 5 to 10 years after treatment is begun. A similar course may be followed by patients who have had subtotal thyroidectomy for Graves' disease due to the slow deterioration of residual thyroid function. Therapy should be monitored with TSH measurements and estimates of free T4. As the goal of levothyroxine therapy is to normalize the thyroid status of the patient, and as serum TSH provides the most sensitive and readily quantification of thyroid status in the patient with primary hypothyroidism, one aims at TSH values in the low normal range. Serum FT4 concentrations will generally be above the middle of the normal range or slightly elevated if serum TSH concentrations are normalized, but serum T3 concentrations (predominantly derived from T4-5'-monodeiodination) will be in the low or midnormal range 12 . In 1,811 athyreotic patients under levothyroxine replacement, FT4 was higher than the upper normal limit in 7.2%, and FT3 was lower than the lower normal limit in 15.2%; the FT3/FT4 ratio was abnormally low in 29.6% (see Figure 9-7) 52 .

figure9-7-1

Figure 9-7. Frequency distribution curves of serum FT4, FT3 and FT3/FT4 ratio in 1,811 athyreotic patients under levothyroxine replacement. Shaded areas indicate normal range (2.5 and 97.5 percentiles) as calculated from 3,875 controls, with vertical lines indicating the median (reproduced with permission) 52 .

In patients with central hypothyroidism one should rely primarily on serum FT4, aiming at values in the mid-normal range 13,53 . The required replacement dose will frequently suppress serum TSH values to below 0.1 mU/l 14 .

Clinical response.

In general, serum thyroxine normalizes before serum TSH, and both may normalize before the disappearance of all of the symptoms of hypothyroidism. In the severely hypothyroid patient with long-standing disease, a number of profound alterations may occur as the hypothyroid state is corrected. Thus, loss of weight, primarily due to mobilization of interstitial fluid as the glycosaminoglycans are degraded, is prominent. The moon facies, coarse nasal voice, puffy fingers, deafness, and sleep apnea will all diminish. Many of nonspecific symptoms, such as fatigue or cold intolerance, will eventually reverse as well. Hair and skin abnormalities take longer to improve. Despite weight loss due to fluid loss, the obese patient should not expect more than a 10-pound weight change, particularly if serum TSH values are only modestly elevated. Virtually all of the weight loss in hypothyroidism is associated with mobilization of fluid, and significant decreases in body fat rarely occur. 54 While metabolic rate increases, in general, appetite increases as well, and a new equilibrium is established.

Treatment failures.

There are few compliant patients whose symptoms and signs do not resolve after thyroid hormone administration. Patients with thyroid hormone resistance sometimes present in this fashion. In patients whose symptoms do not improve with levothyroxine therapy, one should establish that they are taking and absorbing the medication and that it is effective in reducing TSH. The most common cause of treatment failure is poor compliance with ingestion of thyroxine tablets.Non-adherence to treatment can be assessed by a thyroxine absorption test 55 . Compliance might be enhanced by the (supervised) administration of thyroxine once weekly 14 . A slightly larger dose than 7 times the normal daily dose may be required; a singly weekly gift of 1000 µg T4 orally seems to be effective and well tolerated.

Dissatisfaction with treatment outcome

Whereas the vast majority of hypothyroid patients are satisfied with T4 replacement therapy, some are not. A Dutch study reports impaired cognitive functioning in T4 replaced hypothyroid patients relative to a reference population, as evident from worse scores on tests of cognitive motor speed, attention span, and learning and memory tasks (15). An English study reports a higher proportion of distressed subjects in T4-replaced hypothyroid patients than in controls, as evident from general health questionnaires (32.3 vs 25.6%) and thyroid symptom questionnaires (46.8 vs 35.0%)(p<0.01 after correction for age, sex, chronic diseases and chronic medication)(16). The significant difference with controls remained when only hypothyroid patients with TSH values between 0.4 and 4.0 mU/L were analyzed. Subtracting the proportion of distressed subjects in controls from that in the T4-replaced hypothyroid patients leaves us with an excess of 10% of distressed subjects among T4-replaced patients. How can we explain the dissatisfaction, assuming associated autoimmune diseases have been ruled out? It could be that simply being aware of having a chronic disease requiring lifelong treatment and regular control visits makes the patients feel unhappy and less healthy. A more specific explanation would be that L-T4 replacement therapy fails to mimic precisely the thyroidal secretion rates of T4 and T3 and the serum FT4 and FT3 concentrations of healthy subjects (17). A combination of L-T4 and L-T3 replacement therapy might reflect better physiological conditions. A number of randomized clinical trials ( RCTs) comparing T4 monotherapy with T4/T3 combination therapy have been performed to clarify this issue, also provoked by animal data indicating that restoration of the euthyroid state in all tissues of thyroidectomized rats is accomplished only by the combination of T4 and T3, and not by T4 alone. However, a meta-analysis published in 2006 of 11 randomized controlled trials with in total 1216 patients concluded that T4/T3 combination therapy used as replacement therapy for hypothyroid patients provided no advantage when compared with standard T4 monotherapy (18). Despite the outcome of the meta-analysis, the issue of potential benefit of T4/T3 combination therapy cannot be considered as closed in view of further developments: 1. A re-appraisal of the RCTs. The RCTs in the meta-analysis are heterogeneous with respect to the cause of hypothyroidism (thyroidectomy for thyroid cancer, 131I therapy for Graves’ hyperthyroidism, or spontaneous autoimmune hypothyroidism) and the study design (parallel or crossover), but this did not affect outcome (17). In 7 of the 11 RCTs (as well as in a more recent RCT)(19) 50 μg of the daily T4 dose was replaced by a fixed T3 dose (ranging from 10 to 25 μg T3) in patients randomized to combined T4/T3 treatment, giving rise to a wide variation between patients in each RCT in the ratio of the administered T4 to T3 dose (ranging from 20:1 to 1:1 by weight). This is a far cry from mimicking the ratio of T4 to T3 secretion by the human thyroid gland under physiological conditions, which is close to 13:1 by weight (20). Only four trials in the meta-analysis used a variable T3 dose in order to reach the same ratio of administered T4 to T3 dose in all study subjects; the fixed T4 to T3 ratio’s by weight in these studies were 5:1, 10:1, 15:1 and 19:1. Nevertheless, in these four trials, combination therapy was also judged not to be better than monotherapy with T4. To obtain TSH values similar to controls during T4 monotherapy, serum FT4 concentrations higher than controls are needed, whereas serum FT3 values are similar to those in controls (12) (see also figure 9.7) Indeed, the serum FT4 to FT3 ratio in patients randomized to receive T4 monotherapy in the meta-analysis ranged from 4.0 to 6.7, higher than the value of 3.3 observed in controls (21). The serum FT4 to FT3 ratio during combination therapy ranged from 2.2 to 4.8; in only two of the RCTs the ratio’s (3.3. and 3.4 respectively) were close to control values, but both studies still failed to demonstrate superiority of combination therapy over monotherapy (17). Applied outcome measurements in the various RCTs are a number of questionnaires on health-related quality-of-life, cognition, mood and thyroid symptoms, about the same in most but not all studies. The most recent RCT reports significantly better outcome of combination therapy in quality-of-life and depression scales (19), in contrast to all previous RCTs except two early biased ones (17) and one in which the beneft at 3 months was lost at 12 months (22). In this recent RCT when asking the patients themselves, 49% preferred combination therapy, 15% preferred monotherapy, and 36% had no preference (19). That 15% felt better in the period in which the same T4 dose was used as before entering the study, indicates a strong Hawthorne effect also observed in a previous study (23): patients feel better just by participating in a trial. If true, what might explain the preference for combination treatment? It could be loss of body weight: patients at the end of the combination therapy were on average 1.7 kg lighter than after monotherapy (19), and a similar loss of 1.7 kg was observed during combination therapy in another RCT in which patients also preferred the combination (23). In summary, each of the RCTs can be criticized and none is perfect. But these trials are very demanding to perform. If still further RCTs are required (17), special attention should be given to sample size calculation, T4 to T3 ratio’s in combination therapy, and dynamic monitoring of TSH in order to maintain a normal level by adjusting study medications if needed (24). 2. Over- or under-treatment in T4 monotherapy . Serum T3 levels during T4 treatment are mostly in the normal range, but the median is lower than in controls and 15.2% are below the lower normal limit (see figure 9.7) 52 .It is thus obvious that the capacity to generate T3 from exogenous levothyroxine is insufficient in many patients.Serum T3 concentrations similar to those prior to the development of hypothyroidism can be obtained according to a study in euthyroid patients before total thyroidectomy and after surgery when the patients were fully replaced with T4 (25). Nevertheless, the concentration ratio’s of FT4 to FT3 will still be nonphysiological, which may affect deiodinase activity and nuclear T3 receptor occupancy in target tissues. Another study with a double-blind randomized cross-over design investigated T4-replaced patients who were asked to continue with their usual T4 dose, or take 25 μg less or more (26). Mean T4 doses in each of the 6-week study periods were 100, 125 and 150 μg daily. It resulted in expected changes in serum FT4, TSH and cholesterol, but no changes were observed in scores of well-being, cognitive function, and quality-of-life and thyroid symptom questionnaires. It is thus unlikely that slight over- or under-treatment with levothyroxine provides a reasonable explabation for continuous dissatisfaction with T4 monotherapy.It appears more likely that the modality of levothyroxine replacement itself is involved (see 9.8.1). 3. Mode of T3 administration . During T4/T3 combination therapy, the T3 dose is given once or twice daily. It results in wide peak-to-trough variation in serum FT3, e.g FT3 increased by 42% in the first 4 h after T3 but did not change after T4 (27). A slow-release formula of T3 might circumvent the marked changes in serum FT3, and proof of principle of such a preparation has been obtained in a recent study (in which the serum FT4 to FT3 ratio was lower during T4 + slow-release T3 than during T4 monotherapy, but still higher than in controls)(28). Serum FT3 has –in contrast to FT4- a circadian rhythm: the FT3 acrophase occurs in early morning hours around 0300 h, about 90 minutes after the TSH acrophase (29). If one’s goal is to replicate the circadian T3 rhythm and to maintain a physiological ratio of serum FT4 to FT3 throughout 24 h in hypothyroid patients, replacement should provide constant FT4 levels and an early morning rise in serum FT3. This goal possibly can be reached by the administration of levothyroxine once daily in combination with a single nighttime dosing of a sustained-release T3 preparation. 4. Gene polymorphisms. Genetic polymorphisms in deiodinases and thyroid hormone transporters may not only affect serum thyroid hormone concentrations but also the biological availability of thyroid hormone in particular tissues (30). Single nucleotide polymorphisms (SNPs) in the gene encoding for deiodinase type 1 influence the serum FT4 to FT3 ratio, but do not have any association with psychological well-being in patients on thyroid hormone replacement (31,32). An early report did not find an association between the Thr92Ala polymorphism in the deiodinase type 2 gene and well-being, neurocognition or preference for T4/T3 combination therapy (33), but a study with a much larger sample size observed associations between the CC genotype of the D2 Thr92Ala polymorphism and worse baseline scores for general health and greater improvement on T4/T3 combination therapy (32). Lastly, several polymorphisms in the brain-specific thyroid hormone transporter OATP1C1 are associated with fatigue and depression in hypothyroid patients on levothyroxine, but not with neurocognitive functioning or preference for combination therapy (34). Taken together, it might well be that hypothyroid patients dissatisfied with levothyroxine monotherapy are frequent carriers of these polymorphisms, and might have a better response to T4/T3 combination therapy (17,35).One could envisage RCTs restricted to patients who are carriers of such polymorphisms.

Potential adverse effects of treatment.

Life-long treatment with thyroxine when properly monitored with annual assessments, seems to be free of complications. Long-term morbidity and mortality have reported to be normal 1-4 . Thyroxine treatment in TSH-suppressive doses, however, might give reason for some concern as it has been associated with detrimental effects on the heart and the bones. A TSH value of <0.1 mU/l has been identified as a risk factor for the development of atrial fibrillation 36 . Long-term levothyroxine therapy in TSH-suppressive doses may cause left ventricular hypertrophy 37 and increases the risk of ischemic heart disease in patients under the age of 65 years 38 . TSH-suppressive doses of levothyroxine have been associated with bone loss in some but not all studies. A recent extensive meta-analysis concluded that indeed bone mineral density was reduced in hypothyroid patients with a suppressed TSH due to excessive levothyroxine therapy, but only in postmenopausal women 39 . No or a minimal excess of bone fractures, however, has been observed in patients on levothyroxine even if TSH is suppressed 40,41,42,43,,44 A recent population-based study of all patients in Tayside, Scotland  taking L-T4 replacement therapy (n = 17,684) considered fatal and nonfatal endpoints for cardiovascular disease, dysrhythmias, and fractures (45). Patients were categorized as having a suppressed TSH (≤ 0.03 mU/L), low TSH (0.04-0.4 mU/L), normal TSH (0.4-4.0 mU/L), or raised TSH (>4.0 mU/L). Cardiovascular disease, dysrhythmias, and fractures were increased in patients with a high TSH (adjusted hazards ratio 1.95, 1.80, and 1.83 respectively) and in patients with a suppressed TSH (1.37, 1.6, and 2.02 respectively) when compared to patients with a TSH in the laboratory reference range. Patients with a low TSH did not have an increased risk of any of these outcomes. Thus it may be safe for patients treated with T4 to have a low but not suppressed serum TSH.

9.8.3 SITUATIONS REQUIRING DOSE ADJUSTMENT

Table 9-13 lists a number of circumstances in which dosage requirements of levothyroxine may change in compliant patients.

Nonspecific absorption of L-T4 by dietary fibers decreases the bioavailability of T4 and necessitates a higher dose of L-T4 in patients with high intake of dietary fiber (whole-wheat bread, granola, bran) (1). A similar phenomenon may operate with the use of soy protein supplement (2).The timing of the L-T4 dose should be adjusted to take this into account. The association between dietary fiber and levothyroxine malabsorption could not be reproduced in healthy volunteers taking calcium polycarbophil or psyllium (3). Patients with impaired gastric acid secretion require a 22% to 34% higher than usual dose of L-T4 to suppress serum TSH (4), suggesting that normal gastric acid secretion is necessary for effective absorption of L-T4. Most likely this is due to suboptimal dissolution of L-T4 tablets in an environment with higher pH than usual (see section 9.8.1). The L-T4 requirement in autoimmune hypothyroidism is about 18% higher in parietal cell antibodies (PCA)-postive patients than in PCA-negative patients, and a significant positive correlation was found between L-T4 requirement and serum PCA levels (5). In patients with multinodular goiter on suppressive therapy, initiation of the proton-pump inhibitor omeprazole required a 37% increase in L-T4 dose to suppress TSH (4). Likewise, in patients with primary hypothyroidism commencing lansoprazole treatment, 19% of patients required dosage increases in levothyroxine (6). However, pharmacokinetic studies in healthy volunteers who were given L-T4 alone or L-T4 together with proton-pump inhibitors, did not observe significant differences in T4 absorption (7,8). The discrepancy between studies may be caused by differences in the duration of proton-pump treatment: one week in the healthy volunteers (7,8) but up to 6 months in the patients (6). Patients who develop clinical malabsorptive disorders like coeliac disease may require an increase in L-T4 dosage 9,10 . In view of the frequent occurrence of coeliac disease in patients with autoimmune thyroid disease, many authors suggest screening for coeliac disease with anti-gliadin antibodies in patients with hypothyroidism who require higher than expected doses of L-T4. Malabsorption may also occur in patients with short bowel syndrome 11-14 , and in particular cases of lactose intolerance (15) or chronic intestinal Giardiasis (16). Reduced drug absorption requiring higher doses pf levothyroxine may occur after bariatric surgery: it is observed rather frequently after jejunoileal bypass, less often after gastric bypass/gastroplasty, and rarely after biliopancreatic diversion (78.79). Oral liquid levothyroxine formula may be better absorbed compared to levothyroxine tablets following bariatric surgery (80). Interestingly, levothyroxine pharmacokinetics may improve after bariatric surgey: it occurred after sleeve gastrectomy and biliopancreatic diversion with long limps but not after Roux-en-Y gastric bypass (81). Bile acid sequestering agents bind –at least in vitro- large amounts of levothyroxine, and also interfere with the entero-hepatic circulation of thyroid hormone in which T4 and T3 conjugates are excreted in bile and partially deconjugated in the intestine with the release of small amounts of T4 and T3 for reabsorption. Treatment with colestipol or colestyramine may cause a slight increase in TSH in levothyroxine-treated patients, but not in normal subjects (17,18,19). The newer bile acid sequestrant colesevelam reduced absorption of levothyroxine by 96% in healthy subjects (20). An interval of at least 4-5 hours separating thyroxine and bile acid sequestrants is recommended to attain near-normal absorption of L-T4 (21). There are a number of other agents that bind L-T4 and thereby decrease the absorption of L-T4. As a result, serum TSH increases in some but not all levothyroxine-treated patients. The effect of these drug interactions is in general modest, and can be avoided largely by taking L-T4 and the other drug several hours apart (21); sometimes the L-T4 dose has to be increased. Sucralfate binds L-T4 in vitro (22), decreases absorption of L-T4 in healthy volunteers, and may cause resistant hypothyroidism in particular cases (23) but overall its interaction with levothyroxine in L-T4 treated patients seems to be limited (24). Aluminium may complex with levothyroxine, and a dose-related adsorption of levothyroxine with aluminium hydroxide has been demonstrated in vitro (25,26). Aluminium hydroxide treatment increases serum TSH to elevated levels in L-T4 replaced patients (25,26,27).Ferrous sulphate increases serum TSH to slightly elevated levels in hypothyroid patients on stable L-T4 replacement 28,29,30 . The basis of this interaction may be the formation of an insoluble complex by binding of Fe3+ to three T4 molecules. Calcium carbonate adsorbs L-T4 significantly in vitro (31), and it decreases the bioavailability of L-T4 in healthy volunteers (32). Calcium carbonate therapy in L-T4 replaced patients increases serum TSH, sometimes to above the normal range (31,33). Calcium carbonate is also used as a phosphate binder in chronic renal failure. Other phosphate binders may also adsorb L-T4, and both sevelamer and lanthanum carbonate reduce bioavailability of L-T4 in healthy subjects (34,35). The use of sevelamer in hemodialysis patients on levothyroxine is associated with significant increases of L-T4 dose after 6 months of therapy (36). Raloxifene and orlistat have also been reported to interfere with absorption of levothyroxine, but additional studies are required to confirm these observations 21,37,38,39 .

Estrogens increase the serum concentration of thyroxine-binding-globulin (TBG) through increased sialylation of TBG thereby slowing its clearance from the circulation by the liver (40). The effect of estrogens on TBG is dependent on the route of administration, the dose and the chemical structure of the estrogen (41). Transdermal administration of estradiol causes minimal changes in TBG, in contrast to oral estradiol which raises serum TBG by 50-70% (41). The contrast is due to high estrogen levels in the portal vessels and first-pass metabolism in the liver after oral administration. With regard to chemical structure of estrogens, ethinylestradiol in view of its limited liver metabolism causes a rise of serum TBG when administered either orally or transvaginally (42,43). Subjects with normal thyroid glands adapt quickly to the estrogen-induced changes in TBG, and reach a new steady state with no changes in serum FT4 and TSH (41,44). In contrast, estrogen therapy in hypothyroid subjects on stable levothyroxine replacement causes a decrease in serum FT4 and an increase in serum TSH; the effect is dose-dependent and is usually seen within 6 weeks after initiation of estrogens and reaches its peak at 12 weeks (45). Thyroid function tests therefore should be obtained 8-12 weeks after starting estrogen use, and the L-T4 dose adjusted accordingly. Selective estrogen receptor modifiers (SERM) may have similar effects: tamoxifen increases serum TBG by 24% and droloxifene by 41% (46,47). One would expect an increase in required L-T4 dose in hypothyroid patients initiating SERM therapy, but so far this has not been documented. An interesting case report describes the effect of loss of estrogens in a hypothyroid premenopausal women on levothyroxine: chemical castration by the GnRH analogue Goserelin caused an increase of serum FT4 and a decrease of serum TSH (48).

There is an increase in thyroxine requirement in pregnant patients with primary hypothyroidism, probably related to increased lean body mass and increased serum TBG 49,50 . In a review of four series comprising a total of 108 women, serum TSH increased in 58% and the mean L-T4 dose increased from 117 µg to 150 µg 51 . In a prospective study among women with hypothyroidism who were planning pregnancy, an increase in the L-T4 dose was necessary during 17 out of 20 pregnancies; the mean L-T4 requirement increased 47% during the first half of pregnancy (median onset of increase at gestational week 8) and plateaued by week 16, and the increased dose was required until delivery (52). The authors recommended that women should increase their L-T4 dose by the equivalent of two daily doses each week as soon as pregnancy is confirmed. Alternatively, the daily L-T4 dose might be increased by 25 to 50 μg,with thyroid function tests six weeks later (53). A recent study randomized 48 women with treated hypothyroidism seeking pregnancy in two groups: to increase L-T4 by either two tablets per week or three tablets per week (54). TSH suppression below 0.5 mU/L occurred less often at two tablets per week increase than at three tablets per week increase. A two-tablet per week increase in L-T4 initiated at confirmation of pregnancy significantly reduced the risk of maternal hypothyroidism and mimicked normal physiology. Monitoring TSH every four weeks through midgestation and less often thereafter was recommended. The L-T4 dose may be reinstituted at its pregestational level immediately after delivery. Timely adjustment of the thyroxine dose in early gestation might be relevant for infant development. Children at the age of 7-9 years have a lower intelligence quotient if their mother was hypothyroid (elevated TSH) during pregnancy (55).But also children of healthy women with normal TSH but FT4 levels below the 10th percentile (<10.4 pmol/l) at 12 weeks gestation have lower scores on a psychomotor developmental scale at 10 months of age, compared to children of mothers with higher FT4 values 56 ; psychomotor development was not related to maternal FT4 at 32 weeks gestation. A recent population-based study seems to confirm these findings: in pregnant women with normal TSH levels at 13 week gestation (> 2.5 mU/L), maternal hypothyroxinemia (FT4 below the 10 th percentile of 11.7 pmol/l) was associated with cognitive delay of children at 18 and 30 months of age 57 . The findings support the argument of Morreale de Escobar et al that not elevated maternal prenatal TSH levels but maternal hypothyroxinemia (low FT4 levels) is the principal factor leading to poor development of children 58 . For, the fetal brain is dependent on local deiodination of maternal T4 into T3 until the end of the first trimester when the hypothalamus-pituitary-thyroid axis of the fetus becomes functional. The data raise the issue of screening pregnant women for thyroid function disorders in the first trimester.

Several antiepileptic and tuberculostatic drugs induce mixed function oxygenases responsible for hepatic drug oxidation, which accelerates thyroxine clearance via pathways that do not lead to T3 production 59,60,61,62 . Under these circumstances, L-T4 dosage must be increased to compensate for this. Lastly, scattered reports indicate increased L-T4 requirements during treatment with amiodarone, sertraline or chloroquine 63,64,65 . The responsible mechanism is not well understood, although in the case of amiodarone it might have to do with inhibition of T4 transport and T4 deiodination into T3. A more recent study among hypothyroid patients on levothyroxine replacement, however, did not reveal changes in thyroid function tests upon randomization for treatment with sertraline or fluoxetine (both selective serotonin reuptake inhibitors used fro major depression) (83). Tyrosine kinase inhibitors can alter thyroid hormone regulation by mechanisms that apparently are specific to each molecule. Regular assessment of thyroid function is recommended before and during treatment with tyrosine kinase inhibitors (66).Sunitinib induces frequently hypothyroidism in euthyroid patients, but no data are available in levothyroxine-replaced hypothyroid subjects. Imatinib and motesanib do not affect serum TSH in euthyroid subjects with an intact thyroid gland. Imatinib in L-T4 treated hypothyroid patients increases serum TSH in all patients (towards five times the upper normal limit), and FT4 values are reduced by about 60% but remain within the normal range (67). The required L-T4 dose increased by 210% (68). The effect appears rapidly after initiation of therapy and is reversible, since TSH normalized after discontinuation of imatinib. Whereas changes in TBG or in deiodination are not observed, it has been hypothesized that these effects of imatinib are due to stimulation of T4 and T3 clearance by the induction of uridine diphosphate-glucuronosyltransferases. Motesanib in L-T4 treated hypothyroid patients is associated with TSH concentrations ten times higher than baseline on at least one occasion in 50% of patients. Hypothyroidism or TSH above the upper normal limit occurred in 22% of differentiated thyroid cancer and in 61% of medullary thyroid cancer, frequently necessitating a higher replacement dose of L-T4 (69,70). Sorafenib induces hypothyroidism in euthroid patients with intact thyroid glands in 18%. In L-T4 replaced hypothyroid patients sorafenib is associated with an increased TSH in 33% of patients (71). In another study sorafenib decreased serum FT4 and T3 by 11% and 18% respectively, whereas TSH levels increased (requiring a slight increase of L-T4 dose of 10%). The ratio’s of serum T3 to T4 and of T3 to reverse T3 decreased by 18% and 22% respectively, compatible with increased type 3 deiodination (72).In vitro studies demonstrate that several tyrosine kinase inhibitors inhibit cellular uptake of T3 and T4 mediated by the MCT8 thyroid hormone transporter; such a mechanism could also be operative in vivo (83)..

There are fewer conditions in which L-T4 dose requirements decrease. When discussing the effects of androgens on thyroid function, one should consider that synthetic androgens have variable degrees of aromatisation to estrogens in adipose tissue and liver (41). The 17β testosterone esters (given intramuscularly) still undergo significant hepatic metabolism. The 17α alkylated derivatives (given orally) are generally resistant to hepatic metabolism. In euthyroid subjects with an intact thyroid gland, administration of testosterone esters is associated with a decrease in TBG with 14% (44), whereas a nonaromatisable androgen caused a decrease in TBG with 50% (73); serum FT4 and TSH do not change. In hypothyroid patients on a stable dose of L-T4, fluoxymesterone (a non-aromatisable androgen used in breast cancer) caused a dramatic decrease in serum TBG associated with a rise in serum FT4 and a fall in TSH within 4 weeks after initiating therapy; the L-T4 dose had to be decreased by 25%-50% in order to maintain euthyroidism 73 . In patients over the age of 70, levothyroxine requirements are reduced about 25 percent, related to the decrease of lean body mass with age 74,75.76 . Hypothyroid patients with end-stage renal insufficiency need lower doses of T4 after renal transplantation 77 .

Table 9-13. Conditions requiring adjustment of the replacement dose of thyroxine for hypothyroidism.
Increased dose requirement 1. decreased intestinal absorption of T4      - dietary fiber or soy protein supplements 1 ,2,3      - reduced gastric acid secretion: H.pylori infection, atrophic gastritis, protonpump inhibitors 4,-8 ,      - malabsorption: coeliac disease (9,10),short bowel syndrome (11,12,13,14),lactose intolerance (15), intestinal Giardiasis (16), bariatric surgery (78-81)- bile-acid sequestrants: colestipol 17 , cholestyramine 18,19 , colesevelam (20 ) l - agents that bind L-T4: sucralfate (22-24), aluminium hydroxide 25-27 , ferrous s sulfate (28-30) calcium carbonate (31-33), sevelamar (34,36), lanthanumcarbonate (35) 2. increased need for T4     - weight gain- estrogens (40-47)     - pregnancy 49-54 3. Increased metabolic clearance of T4     - antiepileptic drugs (phenobarbital 59 , phenytoin 60 , carbamazepine 61 ) tuberculostatic drugs (rifampicin 62 )

4. precise mechanism unknown     - amiodarone 63 , sertraline 64 , chloroquine (65)

- tyrosine kinase inhibitors:imatinib (67,68),motesanib (69,70),sorafenib(71,72)

Decreased dose requirement 1. decreased need for T4      weight loss      androgens 73 2. decreased metabolic clearance of T4       old age 74,75,76

9.8.4 INTERFERENCE WITH CO-EXISTENT CONDITIONS

Hypocortisolemia. The co-existence of thyroid hormone deficiency and glucocorticoid deficiency is not rare. Primary hypothyroidism due to chronic autoimmune thyroiditis is associated with primary adrenocortical insufficiency due to autoimmune adrenalitis. The very cause of central hypothyroidism in many instances will also result in ACTH deficiency and secondary adrenocortical insufficiency. If the two entities co-exist, it is important to replace glucocorticoid before starting thyroxine. For, treatment of hypothyroidism in patients with glucocorticoid deficiency may precipitate an adrenal crises because the adrenal is incapable to meet the increasing demand for cortisol induced by the rise of the metabolic rate 1 . Some patients with adrenal insufficiency have slightly elevated TSH levels without serological evidence of chronic autoimmune thyroiditis; TSH normalizes with glucocorticoid replacement therapy (2,3,4). It illustrates the small inhibitory effect of cortisol on TSH secretion (5).

Ischemic heart disease. Although treatment of hypothyroidism with levothyroxine will improve myocardial function and reduce peripheral vascular resistance, it will increase the need for oxygen in the myocardium 6,7,8 . In patients with an already compromised myocardial blood supply due to coronary atherosclerosis, thyroxine treatment may provoke anginal symptoms. In a large series of hypothyroid patients, new-onset angina occurred in 2% upon thyroxine treatment; pre-existent angina worsened in 16%, did not change in 46%, and improved in 38% 9 . Patients with preexisting angina should be evaluated for obstructive coronary lesions before thyroxine therapy begins. Retrospective studies suggest that the possibility of myocardial infarction is greater than is the possibility of an adverse event during angiography or angioplasty (10,11,12,13,14).However, it is quite surprising that major surgery, such as coronary artery bypass grafting, can be very easily withstood by the patient with even moderate hypothyroidism as long as attention is paid to reducing the level of analgesics, maintaining adequate ventilation, and controlling the administration of free water 12 . In a few patients, remediable lesions will not be present or, even with bypass grafting, complete correction of the hypothyroid state will not be possible. In such patients, submaximal amounts of levothyroxine supplemented by other agents to enhance myocardial function may be helpful in allowing the reestablishment of normal thyroid function 14 .

Drugs. The metabolism of many drugs is slowed in hypothyroidism, resulting in higher sensitivity to a loading dose and a lower maintenance dose. Marked respiratory depression can occur after a single small dose of morphine. An increase in the dose of digoxin or insulin is sometimes noticed once euthyroidism has been restored.

Growth hormone deficiency. A decrease in serum FT4 and an increase in T3 has been reported following growth hormone (GH) administration with or without a reduction in serum TSH, but literature data are not consistent about these changes (15). The significance of these changes is uncertain, although one study reports a good correlation between changes in serum T3 and resting energy expenditure and cardiac isovolumetric contraction time upon GH treatment (16). The changes in thyroid function can be transient and may revert to normal after a few months. However, in adult hypopituitary patients, GH replacement has been reported to unmask central hypothyroidism in 36%-47% of apparently euthyroid patients, necessitating thyroxine replacement (15,17). At highest risk are patients with organic pituitary disease or multiple pituitary hormone deficiencies. It is therefore prudent to monitor thyroid function in hypopituitary patients starting GH therapy.

Chronic renal failure. Subclinical hypothyroidism is a relatively common condition (approximately 18%) among patients with chronic kidney disease not requiring chronic dialysis, and it is independently associated with progressively lower glomerular filtration rates in unselected oupatient adults (18).

9.9 MYXEDEMA COMA

Definition and pathogenesis.

Myxedema coma is a rare, life-threatening clinical condition in patients with long-standing severe untreated hypothyroidism in whom adaptive mechanisms fail to maintain homeostasis. Most patients, however, are not comatose, and the entity rather represents a form of decompensated hypothyroidism 1 ,-5 . Usually a precipitating event disrupts homeostasis which is maintained in hypothyroid patients by a number of neurovascular adaptations. These adaptations include chronic peripheral vasoconstriction, diastolic hypertension and diminished blood volume; in this way a normal body core temperature is preserved. The hypothyroid heart also compensates by performing more work at a given amount of oxygen by better coupling of ATP to contractile events. In severely hypothyroid patients homeostasis might no longer be maintained if blood volume is reduced any further (e.g. by gastrointestinal bleeding or the use of diuretics), if respiration already compromised by a reduced ventilatory drive is further hampered by intercurrent pulmonary infection, or if CNS regulatory mechanisms are impaired by stroke, the use of sedatives or hyponatremia 2 .

Diagnosis .

The three key features of myxedema coma are 1 : 1. Altered mental status. The patient may be entirely obtruded or may be roused by stimuli. Usually lethargy and sleepiness have been present for many months. Sleep may have occupied 20 hours or more of the day and may have interfered even with eating. There may actually have been transient episodes of coma at home before a more complete variety developed.

2. Defective thermoregulation: hypothermia, or the absence of fever despite infectious disease. Usually coma comes on during the winter months. The severely myxedematous patient becomes essentially poikilothermic. With cold weather the body temperature may drop sharply. The temperature is subnormal, often much depressed: a temperature of 74 F (23º3 C) has been recorded. A thermometer reading lower than the usual 97 F must be used, or hypothermia may be missed.

3. Precipitating event: cold exposure, infection, drugs (diuretics, tranquillizers, sedatives, analgetics), trauma, stroke, heart failure, gastrointestinal bleeding.

Diagnosis on clinical grounds is relatively easy once the possibility is considered. Previous hypothyroidism had been diagnosed in 39% to 61% of all cases (6,7).The pulse is slow, and the absence of mild diastolic hypertension is a warning sign of impending myxedema coma 1 . Any patient with hypothermia and obtundation should be considered as having potential myxedema coma, especially if chronic renal insufficiency and hypoglycemia can be ruled out. The diagnosis can be confirmed by finding a reduced FT4 and marked elevation of serum TSH. However, TSH will not be elevated in myxedema coma due to central hypothyroidism (4% to 18% of cases (6,7,8). Sometimes serum TSH is just slightly elevated, possibly related to co-existent nonthyroidal illness 19 . Creatine phosphokinase is often elevated. Both hypoxia (80%)and hypercapnia (54%) may be present (7) . Hypothermia with a temperature less than 94ºF (34º4C) is seen in 88%.

Treatment. Myxedema coma is a medical emergency. Early diagnosis, rapid administration of thyroid hormones and adequate supportive measures ( Table 9-14 ) are essential for the prognosis.

Table 9-14. Recommendations for the treatment of myxedema coma.

• hypothyroidism large initial intravenous dose of 300-500 µg T4; if no response within 24 hours, add T3.alternative: initial intravenous dose of 200-300 μg T4 plus 10-25 μg T3
• hypocortisolemia intravenous hydrocortisone 200-400 mg daily
• hypoventilation don’t delay intubation and mechanical ventilation too long
• hypothermia blankets, no active rewarming
• hyponatremia mild fluid restriction; conivaptan ?
• hypotension cautious volume expansion with crystalloid or whole blood
• hypoglycemia glucose administration
• precipitating event identification and elimination by specific treatment (liberal use of antibiotics)

 

 

In view of the rarity of myxedema coma, it has been difficult to perform randomized studies to resolve the issue of whether T4 or T3 is the most appropriate treatment. There are advocates of T4 therapy alone 9 . T3 therapy alone 10,11 , and combinations thereof 12 . Differences in opinion about the optimal treatment are caused by a/ the lack of RCTs, b/ the precarious balance between the need to attain effective thyroid hormone levels in target tissues as fast as possible and the risk of precipitating fatal tachycardia or myocardial infarction, and c/ the impairment of T4 into T3 conversion associated with severe illness and inadequate caloric intake, which favours T3 therapy over T4. If T4 alone is used, it should be given parenterally in doses of 300 to 500 µg to replace the calculated T4 deficit 13 . Since the average volume of distribution of T4 in a 70-kg human is approximately 7 L, 420 µg should cause an increase of 77 nM/L in the serum T4 concentration.After this initial ‘loading’ dose, a maintenance L-T4 dose of 75-100 µg/day is given intravenously or orally if the patient is alert. Serum T4 with this schedule usually increase into the normal range within 24 hours, and an increase of serum T3 can also be observed (14). Initial L-T4 doses larger than 500 μg have no advantage and are associated with higher mortality (15). Mortality was 17% in patients selected at random to receive 500 μg T4 intavenously as bolus followed by 100 μg T4 daily, but 60% in patients treated with 100 μg T4 daily (8). However, the difference in mortality between both groups was not significant likely due to small sample size (n=11). If T3 alone is used, it may be given as a 10-20 μg intavenous bolus followed by 10 μg every 4 h for the first 24 h, and 10 μg every 6 h for days 2 and 3 (4).L-T3 doses larger than 75 μg per day are associated with higher mortality (15).The patient should be switched to oral therapy when possible. There has been one case report of a patient with myxedema-associated cardiogenic shock who did respond to T3 but not to T4 treatment (16) but exposing tissues to very high doses of T3 is not without risk. If T4 in combination with T3 is used, 200 to 300 µg of T4 and 10 to 25 µg T3 are given intravenously as an initial dose. After 24 h 100 μg T4 is given intravenously, followed by 50 μg T4 daily from the third day until the patient regains consciousness. Intravenous T3 is continued at a dose of 10 μg every 8-12 h until the patients is conscious and can take T4 maintenance dose 4,12 .

Intravenous glucocorticoid should also be administered during the first days of therapy, since in severe hypothyroidism pituitary-adrenal function is impaired, and the cortisol production rate is lower. While this low production is adequate when cortisol metabolism is reduced, as it is in hypothyroidism, the rapid restoration of a normal metabolic rate from the above treatment may precipitate transient adrenal insufficiency. In addition, the patient should be intubated and measures taken to retain body heat. Central warming may be attempted but peripheral warming should not, since it may lead to vasodilatation and shock. The cutaneous blood flow is markedly reduced in the hypothyroid patient in order to conserve body heat. Warming blankets will defeat this mechanism. Mechanical ventilation may be needed, particularly when obesity and myxedema are combined.Hyponatremia is characteristic and free water restriction and the use of isotonic sodium chloride will usually restore normal serum sodium, as will improved cardiovascular function, which is one cause of the impaired free water clearance.The new vasopressin antagonist conivaptan might be useful in treating hyponatremia as high vasopressin levels have been observed in myxedema coma, but so far no case report has been published in which this drug had been administered Serum glucose should be monitored. Supplemental glucose may be necessary, especially if adrenal insufficiency is present. Hypotension may develop, particularly if myxedema is severe. Volume expansion is usually required to remedy this, since patients are usually maximally vasoconstricted. Dopamine should be added if fluid therapy does not restore efficient circulation. Concomitantly, a vigorous search for precipitating factors should be instituted. Determining whether an infection is present should be a priority, since as many as 35 percent of patients with myxedema coma have infection. Since hypothyroid patients cannot mount an adequate temperature response, the usual signs of infection, including tachycardia, fever, and elevated white blood count, may be absent. Prophylactic antibiotics are indicated until infection can be ruled out; upper respiratory infection should be eliminated. While the hypothyroid patient withstands the stress of surgery in general very well 17 , inadvertently excessive narcotics, sedatives, and hypnotics can tip a severely hypothyroid patient into coma. History taking from family members can be very rewarding in detecting predisposing events. This is illustrated by a case report describing myxedema coma in an elderly woman who had been eating excessive amounts of raw bok choy (Chinese white cabbage) daily for several months in the belief that it would help control her diabetes; the goitrogenic action of compounds like thiocyanates and oxazolidines (generated by eating raw cabbage) was identified as the cause of her coma 20 .

Prognosis .

Most patients begin to show increases in body temperature within the first 24 hours of treatment. The absence of an increase in body temperature within 48 hours should lead to consideration of more aggressive therapy, specifically T3 therapy if it has not already been initiated. Most patients regain consciousness within a few days. Mortality is between 25% and 52%, and sepsis is the predominant cause of death (6,7). Predictors of mortality include hypotension and bradycardia at presentation, need for mechanical ventilation, hypothermia unresponsive to treatment, sepsis, intake of sedative drugs, lower Glasgow Coma Scale, high APACHE II score and high SOFA score (6). Clinical suspicion, early recognition, prompt thyroid hormone replacement, and appropriate support cares remain the key to successful tretament of this rare but often fatal emergency (18).

9.10 SUBCLINICAL HYPOTHYROIDISM

9.10.1 DIAGNOSIS AND ETIOLOGY

Subclinical hypothyroidism is defined as an increased serum TSH in the presence of a normal serum FT4 concentration. Increased refers to values above and normal to values within population-based reference ranges of these hormones. It is however not so simple to diagnose accurately subclinical hypothyroidism in day-to-day practice applying this biochemical definition. Diagnosis of subclinical hypothyroidism is hampered by uncertainty about what constitutes appropriate reference intervals, and by biologic variation in especially TSH. The upper limit of the TSH reference interval was 4.12 mU/L in the National Health and Nutrition Eamination Survey III (NHANES III) for a large reference population that was free of thyroid disease and representative of the U.S. population, in which subjects were excluded who had thyroid antibodies or were taking thyroid medications or other medications affecting thyroid measurements (1). Distribution curves of TSH are skewed to higher TSH concentrations, and under the assumption that this represents undetected thyroid disease the National Academy of Clinical Biochemistry suggested the upper normal limit of TSH should be 2.5 mU/L (2). This proposal has been very controversial, also because it would label as abnormal 10% to 20% of individuals of all ages and 35% of people older than 70 years (3,4,5). More recent population-based studies in which subjects were also excluded in case of abnormal thyroid ultrasonography, observed 97.5 th percentiles of TSH of 3.77 in Germany (6) and 4.1 mU/L in the USA (7), close to the original NHANES III value of 4.12 mU/L. The explanation for the skew in TSH distribution curves toward higher serum TSH is most likely that the upper normal limit of TSH increases with advanced age (8). When NHANES III data were reanalyzed by TSH distribution curves for specific age deciles, a progressive shift in the curves to higher TSH with age was observed, rather than a skew to higher values. E.g. the upper normal limit of TSH is 7.5 mU/L in subjects older than 80 years; 70% of subjects older than 80 years would have been labeled as having raised TSH when an upper normal limit of 4.5 mU/L is used. Age-specific reference ranges are thus recommended. Besides conceptual problems with reference ranges, the diagnosis of subclinical hypothyroidism is jeopardized by considerable biologic variation in TSH values. A particular study enrolled 21 patients with subclinical hypothyroidism (identified with serum TSH between 5 and 12 mU/L and normal T4, confirmed on two occasions 3 months apart) without former thyroid disease, who underwent monthly repeated measurements without intervention (9). In the one-year follow-upp period, one patient appeared to be euthyroid at all visits, and one patient developed profound overt hypothyroidism and was treated.The remaining patients had subclinical hypothyroidism at 74% of the visits, overt hypothyroidism at 22% and normal thyroid function tests at 4% of the visits. Diagnosis of overt hypothyroidism was highly dependent on T4 reference limits. In individual patients serum TSH was correlated to both TPO antibodies and to urinary iodine excretion, but not to hypothyroid symptoms and signs (10). The study shows how TSH and FT4 vary around the outer limits of their reference ranges, how limited the information is obtained from a single set of thyroid function tests, and how biologic variation may change the diagnosis from visit to visit. It is therefore recommended, especially If a slightly increased serum TSH is found, to take a second blood sample after 3-6 months in order to ascertain the diagnosis of subclinical hypothyroidism. If a normal TSH is found in the second blood sample, this can rarely be attributed to ultradian and circadian TSH rhythms ( the relative risk of misjudging mean TSH serum levels by a single TSH determination between 07.00 and 17.00 hours is only 0.09% for values above 4.0 mU/l ) 11 .

It is highly relevant after the biochemical diagnosis to establish a nosologic diagnosis to evaluate which condition is responsible for the elevated TSH. The llist of possible causes is very long (21). The most common causes of subclinical hypothyroidism are chronic autoimmune thyroiditis (Hashimoto’s disease), previous 131I therapy or thyroidectomy, and inappropriate dosage of thyroxine or antithyroid drugs. Loss-of-function mutations in the gene encoding for the TSH receptor are relatively common in isolated hyperthyrotropinemia, especially in children and adolescents (12,13). Other causes of an elevated TSH are interference of heterophilic TSH antibodies in TSH immunoassys, nonthyroidal illness syndrome (recovery phase), impaired renal function, untreated adrenal insufficiency (Addison’s disease) and obesity. Strictly speaking, the latter group should not be labeled as subclinical hypothyroidism because thyroid disease is absent and management is directed to the nonthytoidal cause. The case of obesity is illustrative in this respect (14). In patients with morbid obesity (BMI range of 30-67 kg/m²) TSH levels correlate positively with BMI (r=0.91), and the mean BMI change from 49 to 32 kg/m² after bariatric surgery is associated with a reduction in mean TSH levels from 4.5 to 1.9 mU/L; FT4 levels are not associated with BMI, and subclinical hypothyroidism observed in 10.5% disappears after weight reduction (15). Thyroid autoimmunity is not a major cause sustaining the high rate of an elevated TSH in morbid obesity (16). Thus it is important to establish if subclinical hypothyroidism is caused by an underlying thyroid disease. Apart from history and physical examination, this can be done easily by measuring TPO antibodies in serum.

It is highly relevant after the biochemical diagnosis to establish a nosologic diagnosis to evaluate which condition is responsible for the elevated TSH. The llist of possible causes is very long (21). The most common causes of subclinical hypothyroidism are chronic autoimmune thyroiditis (Hashimoto’s disease), previous 131I therapy or thyroidectomy, and inappropriate dosage of thyroxine or antithyroid drugs. Loss-of-function mutations in the gene encoding for the TSH receptor are relatively common in isolated hyperthyrotropinemia, especially in children and adolescents (12,13). Other causes of an elevated TSH are interference of heterophilic TSH antibodies in TSH immunoassys, nonthyroidal illness syndrome (recovery phase), impaired renal function, untreated adrenal insufficiency (Addison’s disease) and obesity. Strictly speaking, the latter group should not be labeled as subclinical hypothyroidism because thyroid disease is absent and management is directed to the nonthytoidal cause. The case of obesity is illustrative in this respect (14). In patients with morbid obesity (BMI range of 30-67 kg/m²) TSH levels correlate positively with BMI (r=0.91), and the mean BMI change from 49 to 32 kg/m² after bariatric surgery is associated with a reduction in mean TSH levels from 4.5 to 1.9 mU/L; FT4 levels are not associated with BMI, and subclinical hypothyroidism observed in 10.5% disappears after weight reduction (15). Thyroid autoimmunity is not a major cause sustaining the high rate of an elevated TSH in morbid obesity (16). Thus it is important to establish if subclinical hypothyroidism is caused by an underlying thyroid disease. Apart from history and physical examination, this can be done easily by measuring TPO antibodies in serum.

The prevalence of subclinical hypothyroidism in the general population is rather high in the order of 4% to 8%; it is higher in iodine-replete areas than in iodine-deficient areas (17,18) (see also section 9.2). In the classical population-based study among adults in the English county of Whickham the prevalence was 75 per 1000 women and 28 per 1000 men 19 . About 75% have TSH values between 5 and 10 mU/L, and 25% have TSH values greater than 10 mU/L (20). The higher prevalence of subclinical hypothyroidism in females than in males and in older than in younger subjects is in agreement with the higher prevalence of thyroglobulin and thyroid peroxidase (microsomal) antibodies in women and in elderly people.

9.10.2 NATURAL HISTORY

The natural history of subclinical hypothyroidism is reported in many studies, although it remains difficult to predict whether the increased TSH levels will return spontaneously to within the normal range, will remain stable, or will increase to higher values with development of overt hypothyroidism. In general it can be said that the higher the initial TSH, the higher the risk of progression; the presence of TPO antibodies potentiates the risk. Spontaneous normalization of increased TSH values in subclinical hypothyroidism is a well-known phenomenon, but the reported frequency of normalization differs markedly between studies from 4% up to 52% 1-10 . Reasons for the wide variation in normalization of TSH are differences in duration of follow-up, heterogeneity of study populations, and possibly age. In a prospective observational study normalization of TSH occurred at a median time of 18 months (range 6-60 months) (11). Some studies involve homogeneous populations (e.g.only subjects with proven autoimmune thyroiditis), whereas others do not specify the cause of subclinical hypothyroidism and consequently may harbour many subjects without underlying thyroid disease. E.g. studies in children and adolescents with subclinical hypothyroidism report stable, normalized or increasing TSH values in 31%, 31% and 38% respectively over a mean follow-up period of 41 months when all participants had Hashimoto’s thyroiditis (12), and 47%, 41% and 12% over a mean follow-up period of 24 months when all participants had idiopathic subclinical hypothyroidism (13).Normalization of TSH is more frequent in subjects with moderate TSH elevation up to 10 mU/L, with or without thyroid autoimmunity. Normalization of TSH might also be more common in old age. In a study of 107 subjects (mostly with autoimmune thyroiditis) with a mean age of 62 yr, 37% normalized TSH at a mean follow-up of 2.7 yr (9,11). In contrast, in a study of 21 subjects (with unspecified cause of elevated TSH) with a mean age of 85 yr, 52% normalized TSH at a mean follow-up of 3 yr (10).

Progression to overt hypothyroidism ranges from 7.8% to 17.8% in various studies 2,4,5 . According to the initial serum TSH concentrations (TSH 4-6, >6-12, >12 mU/L, Kaplan-Meier estimates of the incidence of overt hypothyroidism in subclinically hypothyroid women were 0%, 42.8%, and 76.9% respectively after 10 years (or 0%, 3%, and 11% respectively per year) 8 . The incidence of overt hypothyroidism was higher in patients with TPO antibodies ((58.5% vs 23.2%). The importance of thyroid antibodies is also evident from a Dutch study: 9.6% of 55-year old women with TPO antibodies had raised TSH levels 10 years later, in contrast to 3.2% of women without antibodies 14 . The most extensive data are from a 20-year follow-up in the participants of the Whickham survey 15 . The incidence of overt hypothyroidism was 4.1 per 1000 women per year and 0.6 per 1000 men per year. Odds ratio’s (with 95% CI) for development of spontaneous hypothyroidism in surviving women are 14 (9-24) for raised TSH regardless of thyroid antibody status, 13 (8-19) for positive thyroid antibodies regardless of TSH, and 38 (22-65) for raised TSH and positive thyroid antibodies combined. Odds ratio’s for men are higher: 44 (19-104) for raised TSH regardless of thyroid antibody status, 25 (10-63) for positive thyroid antibodies regardless of TSH, and 173 (81-370) for raised TSH and positive thyroid antibody status combined. Most interestingly, the risk for developing hypothyroidism in the Whickham survey already starts at TSH levels of 2.0 mU/L (see also section 9.2). Similar cutoff values of 2.5 mU/L for predicting hypothyroidism have been reported in a 5-yr follow-up study among Dutch healthy female relatives of patients with autoimmune thyroid disease (16), and in an Australian population-based study with a 13-yr follow-up (17). A hypoechographic pattern on thyroid ultrasound also increases the risk of progression, even in the absence of TPO antibodies in serum (18). In subjects ≥65 yr, persistence of subclinical hypothyroidism, after 2 and 4 yr was 56%; resolution of elevated TSH was more common with a TSH 4.5-6.9 mU/L (46% vs 10% with TSH 7-9.9 mU/L and 7% with TSH ≥10 mU/L) and with TPO-Ab negativity (48% vs 15% for positive TPO-Ab) (19). TSH ≥10 mU/L was independently associated with progression to overt hypothyroidism. Transitions between euthyroidism and subclinical hypothyroidism were more common between 2 and 4 yr; age and sex did not affect transitions. Taking a high-dose phytoestrogen dietary supplementation (30 g soy protein with 16 mg phytoestrogens, representative of a vegetarian diet) for 8 weeks increases 3-fold the risk in subjects with subclinical hypothyroidism to develop overt hypothyroidism (20).

9.10.3 SYSTEMIC MANIFESTATIONS

A multitude of papers have been published on alterations in subclinical hypothyroidism as compared to euthyroid subjects. The vast literature on this topic from 1990 through April 2007 has been nicely reviewed by Biondi and Cooper (1) and updated by Cooper and Biondi in 2012 62 . The abnormalities listed in Table 9-15 , have been reported in some but not all studies on subclinical hypothyroidism. The described abnormalities are in general minor, and more frequent in subjects with the highest TSH values 2 .

Effect on symptoms, quality-of-life (QoL), and cognitive function. Hypothyroid symptoms occured more often than in controls in some small studies and in the Colorado study 3,4,,5 , but not in a large population-based study 6 . A study among healthy females with a family history of thyroid disease, recruited by advertisement, indicated a higher lifetime frequency of depression in subjects with subclinical hypothyroidism (56%) than in euthyroid subjects (20%) 7 . But again large population-based studies did not reveal lower well-being, impaired QoL, or more depression and anxiety 6,8-12 . Impaired memory function has been reported in subclinical hypothyroidsm in some early small series of patients (13,14), but more recent large population-based studies have not corroborated this observation also not in the elderly: cognitive functions were not different from controls (6,8,9,10,12). In contrast, a study using functional MRI suggested that working memory (but not other memory functions) is impaired by subclinical hypothyroidism (15). Its findings are apparently confirmed by another small study showing that cognitive impairment in subclinical hypothyroidism appears predominantly mnemonic in nature, suggesting that the etiology is not indicative of general cognitive slowing (16).Subclinical hypothyroidism seems to be less symptomatic in the elderly (8,63,64). Subjects with subclinical hypothyroidism in their 8 th decade of life even had increased walking speed and retention of physical function as compared with their peers who were euthyroid (55).

Table 9-15. Abnormalities reported in some but not all studies on subclinical hypothyroidism.

Symptoms • hypothyroid symptoms • impaired well-being and quality of life • impaired cognitive functions (working memory) • mood disturbances
Signs • impaired left ventricle diastolic and systolic function•hypertension • increased systemic vasular resistance • increased central arterial stiffness • impaired endothelium function • increased carotid intima-media thickness • impaired muscle energy metabolism • impaired peripheral nerve conduction latency and amplitude • impaired stapedial reflex
Biochemistry • high serum total and LDL cholesterol • high HOMA index (insulin resistance) • high serum C-reactive protein • low factor VIIa • high serum lactate during exercise • low serum IGF-1, high serum leptin

Cardiac effects . The many cardiac effects are reviewed by Biondi and Cooper (1,62). Impaired left ventricular diastolic function at rest has been clearly demonstrated in subclinically hypothyroid subjects by Doppler echocardiography and radionuclide ventriculography: isovolumetric relaxation time is prolonged and time-to-peak filling rate is impaired as compared to controls. Left ventricular systolic function at rest was reported as normal, but using the more sensitive Doppler echocardiography seems to be impaired as documented by an increased preejection period (PEP) to left ventricular ejection time (LVET) ratio. Cardiac MRI, a most accurate procedure to evaluate cardiac volumes and function, demonstrates a significant decrease in preload (end-diastolic volume) and a significant increase in the afterload (systemic vascular resistance), thereby leading to impaired cardiac performance (17). Pulsed wave tissue Doppler imaging allows to measure velocities at any point of the ventricular wall during the cardiac cycle; myocardial time intervals were prolonged at both the posterior septum and the mitral annulus compared to controls (18). Ultrasonic myocardial texture analysis reveals altered myocardial composition, suggesting early myocardial structural changes(19). Recent studies suggest furthermore impaired coronary flow reserve (20,21). In summary, the most consistent cardiac abnormality in suclinical hypothyroidism is impaired left ventricular diastolic function, characterized by slowed myocardial relaxation and impaired ventricular filling. Impaired left ventricular systolic function is not consistently reported but has been identified with more sensitive techniques (1). Cardiac performance is also impaired during exercise (22,23).The cardiac changes in subclinical hypothyroidism are less severe but otherwise similar to those observed in overt hypothyroidism, suggesting a continuum in the cardiac effect of thyroid hormone.

Vascular effects. A higher prevalence of diastolic hypertension in subclinical hypothyroidism as compared to controls has been reported in some but not all studies (1,24,25). Three factors contribute to systemic hypertension in overt hypothyroidism: increased peripheral vascular resistance, increased arterial stiffness, and endothelial dysfunction. The same factors seem to operate in subclinical hypothyroidism. Systemic vascular resistance is increased in some studies but not in others (1,19). As T3 directly affects vascular smooth muscle cells promoting relaxation, subclinical hypothyroidism might affect vascular tone. Increased arterial stiffness has been identified as an independent risk factor for cardiovascular morbidity and mortality. Increased arterial stiffness has been demonstrated in several studies using various techniques like pulse wave velocity (24,26). The vascular endothelium regulates vascular smooth muscle function by diffusion of nitric oxide from the endothelium to the smooth muscle cells, inducing relaxation. Flow-mediated endothelium-dependent vasodilatation is significantly impaired in subclinical hypothyroidism compared to controls (27,28,29). The endothelium dysfunction is attributed to reduced nitric oxide availability (28). Low-grade chronic inflammation could be responsible for impaired nitric oxide availability by a cyclo-oxygenase 2-dependent pathway, increasing oxidative stress in subclinical hypothyroidism due to Hashimoto’s thyroiditis (30). Carotid intima-media thickness is useful in the early diagnosis of atherosclerosis and coronary heart disease. Patients with subclinical hypothyroidism have higher carotid intima-media thickness than age-and sex-matched controls in some but not all studies (29,31,32). The observed vascular changes potentially increase the risk of atherosclerosis and coronary artery disease.

Effects on biochemical tests. Serum lipids (total cholesterol, LDL cholesterol) may be increased in subclinical hypothyroidism, but the existing data in the literature are inconsistent (1). The lipid pattern is more abnormal with serum TSH > 10 mU/L. The Whickham Survey did not observe increased total cholesterol levels in subclinical hypothyroidism versus controls (33), in contrast to other population-based studies like the NHANES III (34) and the Busselton study (34) which did observe higher cholesterol levels in subclinical hypothyroidism; however, in the latter two studies the difference with controls disappeared almost completely after adjustment for age and sex. Other population-based studies report that serum TSH levels > 5.5 mU/L are associated with a rise in serum total cholesterol of 0.23 mmol/l (36), and that an increase of 1 mU/L of serum TSH is associated with an increase of serum cholesterol of 0.09 mmol/l in women and of 0.16 mmol/l in men (37). In older women LDL cholesterol was 13% higher and HDL cholesterol was 12% higher with TSH values > 5.5 mU/L compared to normal TSH values (38). The lipid peroxidation marker malondialdehyde is elevated in subclinical hypothyroidism compared to controls (39). Remnant lipoproteins are more often present in the fasting serum in subclinical hypothyroidism than in controls (40). Serum triglycerides are usually normal. No associations with Lp(a) have been observed. The relationship between TSH and LDL cholesterol may depend on other factors like the presence of insulin resistance (41) and smoking (42). The risk of hypercholesterolemia in subclinical hypothyroidism was restricted to smokers in one study (42). Fasting insulin levels in subclinical hypothyroidism are higher than in controls, and the homeostasis model of assessment (HOMA-IR) index suggests insulin resistance in one study but not in another (43,44). Homocysteine levels are apparently not related to subclinical hypothyroidism (31,45). Fasting insulin correlated positively with hsCRP (highly sensitive C-reactive protein), a strong predictor of cardiovascular risk (43,46).C-reactive protein levels are higher in subclinical hypothyroidism than in controls in some but not all studies (43,46,47).Another study observed higher hsCRP, total and LDL cholesterol, asymmetric dimethylarginine and arginine levels but lower nitric oxide levels in subclinical hypothyroidism than in controls (48). Alterations in coagulation parameters have also been reported (1). Global fibrinolytic capacity was lower in subclinical hypothyroidism than in controls. Subjects with subclinical hypothyroidism identified in a population based study, as compared to age- and sex-matched controls, had no changes in hemostatic factors but their factor VIIa levels were 10% lower (49). In summary, data on a potential association of subclinical hypothyroidism and traditional cardiovascular risk factors (like cholesterol) and nontraditional cardiovascular risk factors (C-reactive protein, coagulation parameters) are not consistent.

Other effects . In the central nervous system an abnormal stapedial reflex but no abnormalities in brainstem auditory evoked potentials has been observed (1,50). Discrete changes in peripheral nerve function are reported: conduction velocities are normal, but motor distal latencies are prolonged and amplitudes are decreased relative to controls 1,51 . Muscle metabolism is impaired: during exercise (but not at rest) blood lactate is higher in subclinical hypothyroidism than in controls, consistent with impaired mitochondrial oxidative function 52,53 . There exists lower exercise tolerance and less muscle strength (54). In contrast, a community-based study in 70-79 year old subjects does not demonstrate increased risk of mobility problems (tested by mean usual and rapid gait speed, cardiorespiratory fitness and walking ease) in subclinical hypothyroidism; in fact, those with TSH levels between 4.5 and 7.0 mU/L showed a slight functional advantage over euthyroid subjects (55). In another age- and sex-adjusted analysis, subclinical hypothyroidism was associated with lower vital capacity at rest and a lower work rate at the ventilator anaerobic threshold (23). Subclinical hypothyroidism is more common in patients with common bile duct stones as compared to nongallstone controls (56), but it was not analyzed if this association is independent of obesity. It has also been suggested that the prevalence of subclinical hypothyroidism is higher in patients with deep venous thrombosis (57). A study in postmenopausal women reports that subclinical hypothyroidsm affects not bone turnover but bone structure in the calcaneus (lower heel QUS) (58). Serum leptin concentrations are higher in subclinically hypothyroid than euthyroid postmenopausal women, even when controlling for body mass index (59). Plasma total and acylated ghrelin concentrations are not significantly changed by subclinical hypothyroidism (60). Fasting serum IGF-1 levels are lower in subclinical hypothyroidism than in controls (61). A prospective cohort study among community-dwelling US subjects ≥65 yr with a follow-up of 13 yr reported an association between endogenous subclinical hypothyroidism and incident hip fracture in men (hazard ratio 2.45, 95% CI 1.27-4.73) but not in women (65).

9.10.4. ASSOCIATIONS WITH CARDIOVASCULAR MORBIDITY AND MORTALITY.

Studies in the early 1970s suggested preclinical hypothyroidism as a risk factor for coronary heart disease, presumably via increased cholesterol levels 1,2 . Since then, many but not all studies have demonstrated a higher prevalence of classical (like hypercholesterolemia) and nonclassical risk factors for cardiovascular disease in subclinical hypothyroidism, as outlined in section 9.10.3. So the question has arisen whether or not subclinical hypothyroidism is associated with a higher prevalence or incidence of cardiovascular disease. To answer this question, a number of epidemiological studies have been performed. Because of inconsistencies in the obtained results of population-based studies, the data has been subjected to meta-analyses.

Population-based studies . We will review both cross-sectional (transversal) and follow-up (longitudinal) population-based studies in chronological order. In 1996, a 20-yr follow-up of the Whickham Survey in the UK among men and women of 18 yr and older, did not find an association between autoimmune thyroid disease at study entry (defined as treated hypothyroidism, positive thyroid antibodies and/or or elevated serum TSH) and subsequent development of ischemic heart disease or increased circulatory or all-cause mortality 3 . In 2000, a 4.5-yr follow-up of the Rotterdam Study among women of 55 yr and older (mean age 69±7.5 years), subclinical hypothyroidism was not associated with an increased incidence of myocardial infarction. However, at study entrance subclinical hypothyroidism was associated with a higher age-adjusted prevalence of aortic atherosclerosis (odds ratio 1.7, 95% CI 1.1 to 2.6) and myocardial infarction (odds ratio 2.3, 95% CI 1.3 to 4.0) 4 . Additional adjustment for body mass index, serum cholesterol, blood pressure, smoking and the use of β-blockers did not affect these estimates. The population attributable risk for subclinical hypothyroidism associated with myocardial infarction was 14%, within the range of that for known major risk factors for cardiovascular disease (hypercholesterolemia 18%, smoking 15%, hypertension 14%, diabetes 14%). In 2001, a 10-yr follow-up study in the UK among men and women of 60 yr and older, found no association of subclinical hypothyroidism and death from circulatory disease, but 40% of subclinically hypothyroid subjects developed overt hypothyroidism and started L-T4 therapy (5). In 2004, a 10-yr follow-up study among atomic bomb survivors from Nagasaki among men and women of 40 yr and older, found that subclinical hypothyroidism was associated with an increased mortality from all causes only in men after 3-6 years, but not after 10 years (6); the cross-sectional analysis at baseline showed an increased risk of ischemic heart disease. In 2004, a 4-yr follow-up study in Leiden among men and women of 85 yr, subclinical hypothyroidism was associated with greater longevity and a decreased risk of death from cardiovascular disease, attributed to a lower metabolic rate (7). In 2005, the 20-yr follow-up in the Busselton study in Western Australia among men and women of 17-89 yr, identified subclinical hypothyroidism as an independent predictor of coronary heart disease but not of death from cardiovascular disease (8). In the longitudinal analysis the increased risk was present at TSH levels of both 4-10 mU/L and >10 mU/L, but the risk at baseline in the cross-sectional analysis was only present at TSH levels >10 mU/L. in 2005, a 4-yr follow-up study of the Health, Aging and Body Composition Study in the USA among men and women of 70-79 yr, concluded that subclinical hypothyroidism is associated with an increased risk of congestive heart failure at TSH levels of 7.0 mU/L or greater, but not with other cardiovascular events and mortality (9). TSH levels between 4.5 and 6.9 mU/L carried no risk. By using TSH as a continuous variable, each standard deviation increase of 4.0 mU/L was associated with a 30% increase in congestive heart failure. In 2006, a 13-yr follow-up of the Cardiovascular Health Study in the USA among men and women of 65 yr and older, reported no association between subclinical hypothyroidism and cardiovascular disorders or mortality (10). In 2007, a 2.7-yr follow-up study in hospitalized patients admitted to the department of cardiology in Pisa among men and women of mean age 61 yr, observed lower survival rates for cardiac death and overall death in subclinical hypothyroidism than in euthyroidism with hazard ratio’s (after adjustment for several risk factors) of 2.40 (95% CI 1.36-4.21) and 2.01 (95% CI 1.33-3.04) respectively (11). In this study subclinical hypothyroidism was defined as TSH levels between 4.5 and 10 mU/l with FT4 and FT3 within the reference range; its cause was atrophic thyroiditis in 39%, Hashimoto’s thyroiditis in 36% and thyroidectomy or 131I therapy in 25%. In 2007, a cross-sectional population-based study in Tromso, Norway among men and women of 55-74 yr, found that subjects with subclinical hypothyroidism (TSH 3.5-10 mU/L) had no signs of cardiac dysfunction (12). In 2008, a 12-yr follow-up in the Cardiovascular Health Study in the USA among men and women 65 yr and older, indicated a greater incidence of heart failure in subclinically hypothyroid participants with TSH of 10 mU/L or greater compared with euthyroid participants with TSH values of 0.45 to 4.50 mU/l (adjusted hazard ratio 1.88, 95 CI 1.05-3.34); no increased risk of heart failure was observed in subclinical hypothyroidism with TSH values 4.5 to 9.9 mU/L (13). In 2010, a 10.6-yr follow-up in the EPIC-Norfolk study in the UK among men and women 45-79 yr, no association was found between subclinical hypothyroidism and the risk of coronary heart disease, despite the association between thyroid hormone levels and cardiovascular risk factors (14). In 2010, a 7.5-yr follow-up in the Japanese-Brazilian Thyroid Study among men and women 30 yr and older, observed subclinical hypothyroidism was associated with all-cause mortality (adjusted hazard ratio 2.3, 95% CI 1.2-4.4) but not with cardiovascular mortality (hazard ratio 1.6, 95% CI 0.6-4.2) (15). In 2010, the 20-yr follow-up of the Whickham Survey in the UK among men and women of 18 yr and older, was reanalyzed (16). Incident ischemic heart disease was significantly higher in the group with subclinical hypothyroidism vs. the euthyroid group (hazard ratio 1.76 , 95% CI 1.15-2.71), and incident ischemic heart disease mortality was also increased in subclinical hypothyroidism (hazardratio 1.79, 95% CI 1.02-3.56). Subsequent treatment of subclinical hypothyroidism with L-T4 appears to attenuate ischemic heart disease-related morbidity and mortality. In 2011 analysis of the PreCis database (from the Cleveland Clinic Preventive Cardiology Clinic) revealed an association between moderate (TSH 6.1-10 mU/L) but not with mild (TSH 3.1-6.0 mU/L) subclinical hypothyroidism and coronary heart disease prevalence and all-cause mortality in both genders in subjects under the age of 65 (but not in the age gourp older than 65 yr) (24). The PROSPER study among men and women aged 70-82 yr at high cardiovascular risk showed in 2012 an association between subclinical hypothyroidism and heart failure only at TSH >10 mU/L (25).Data from the MrOs cohort of men ≥65 yr followed for 8.3 yr did not show any relation between subclinical hypothyroidism and mortality, but only 8 men had TSH ≥10 mU/L (26). A study from Taiwan among the adult populationa with a 10-yr follow-up found an association between subclinical hypothyroidism and all-cause mortality and cardiovascular disease (27).The ‘oldest old’ from the Cardiovascular Health Study were retested after 13 yr (mean age 85 yr) ((28): there was a 13% increase in TSH, 1,7% increase in FT4 and 13% decrease in T3 over this period.There was no association between subclinical hypothyroidism or persistent TPO-Ab and death, but higher FT4 levels were associated with death. The findings raise concern for treatment of mildly elevated TSH levels in old age.In another report from the Cardiovascular Healthstudy among individuals ≥65 yr published in 2013, the 10-yr risk of incident coronary heart disease, heart failure and cardiovascular death was not related to persistent subclinical hypothyroidism (29).A nested case-control study among postmenopausal women within the Women’s Health Initiative cohort did not find an association between subclinical hypothyroidism and incident myocardial infarction (30). Among participants in the 3 rd NHANES, subclinical hypothyroidism was associated with greater mortality in those with congestive heart failure but not in those without (31). In summary , there are major discrepancies in epidemiological data about cardiovascular risk in subclinical hypothyroidism (1). This may be due to differences in the populations studied in terms of age, sex, race, life style and duration of follow-up, to differences in the extent of the TSH elevation, and to differences in the assessment of cardiovascular endpoints. Nevertheless,a general trend can be detected: the risk of adverse health outcomes is higher with higher TSH values (especially at TSH ≥ 10 mU/L), and is lower with advancing age (especially at age >65-70 yr).

Meta-analysis studies. In 2006, a meta-analysis of 14 observational studies published from 1996 to April 2005, indicated that subclinical hypothyroidism increased the risk of coronary heart disease (odds ratio 1.65, 95% CI 1.28-2.12) (17). The odds ratio’s varied little in analyses adjusted for various factors and analyses limited to various subgroups. In 2007, a meta-analysis of 4 studies published from 1966 to April 2007, revealed fthat subclinical hypothyroidism had a hazard ratio of 1.21 (95% CI 0.86-1.69) for circulatory mortality and of 1.25 (95% CI 1.03-1.53) for all-cause mortality (18). In 2008, a meta-analysis of 10 studies published up to January 2008, reports a relative risk for subclinical hypothyroidism for coronary heart disease of 1.20 (95% CI 0.97-1.49) (19). Risk estimates were higher among participants younger than 65 years (RR 1.51, 95% CI 1.09-2.09) for studies with mean participant age <65 yr, and the RR was1.05 (95% CI 0.90-1.22) for studies with mean participant age 65 yr and older. The RR was 1.18 (95% 0.98-1.42) for cardiovascular mortality and 1.12 (95% CI 0.99-1.26) for all-cause mortality. In 2008, a meta-analysis of 9 studies published up to July 2007, calculated a hazard ratio for subclinical hypothyroidism for all-cause mortality of 1.02 (95% CI 0.78-1.35) in cohorts from the community and of 1.76 (95% CI 1.36-2.30) in cohorts of participants with comorbidities (20). In 2008, a meta-analysis of 4 studies published in the period 2001-2005, showed a significant risk of subclinical hypothyroidism for coronary heart disease: the relative risk was 1.53 (95% CI 1.31-1.79) at baseline, and 1.19 (95% CI 1.02-1.34) at follow-up (21). The relative risk for all-cause mortality at follow-up was not significant, but the relative risk of cardiovascular mortality at follow-up was 1.28 (95% CI 1.02-1.60). In 2008, a meta-analysis of 15 studies published up to May 2007, concluded that incidence and prevalence of ischemic heart disease were higher in subclinically hypothyroid subjects compared with euthyroid subjects from studies including those younger than 65 yr, but not in studies of subjects aged older than 65 yr: odds ratio 1.57 (95% CI 1.19-2.06) vs 1.01 (95% CI 0.87-1.18), and 1.68 (95% CI 1.27-2.23) vs 1.02 (95% CI 0.85-1.22) (22). Cardiovascular/all-cause mortality was also elevated in participants from the younger than 65-yr studies, but not from the studies of older people: odds ratio 1.37 (95% CI 1.04-1.79) vs 0.85 (95% CI 0.56-1.29). Prevalent ischemic heart disease was higher in subclinically hypothyroid subjects of both genders, although this was significant only in women. These data suggest that increased vascular risk may only be present in younger individuals with subclinical hypothyroidism. In 2010, a meta-analysis based on individual participant data of 9 prospective cohort studies was presented, encompassing 41,685 participants with 381,647 person-years of follow-up (23). Subclinical hypothyroidism defined as TSH values of 4.5 to 19.9 mU/L with normal T4 concentrations was present in 2,621 subjects. Compared with euthyroidism, the hazard ratio for coronary heart disease events increased with higher TSH concentrations, from 1.07 (95% CI 0.84-1.35) for TSH 4.5-6.9 mU/L, 1,12 (95% CI 0.88-1.44) for TSH 7.0-9.9 mU/L, to 2.00 (95% CI 1.25-3,20) for TSH 10 mU/L and higher. Total mortality was not increased, but coronary heart disease mortality was increased only at TSH levels of 10 mU/L and higher (hazard ratio 1.64, 95% CI 1.11-2.42). Results were similar after further adjustment for traditional risk factors. Risks did not significantly differ by age, gender or preexisting cardiovascular disease.In 2012 the results were published of a pooled analysis of individual participant data from 6 prospective cohorts in which subclinical hypothyroidism was defined as TSh of 4.5-19.9 mU/L with normal FT4 (32). Risks of heart failure evnts were increased with higher TSH levels: hazard ratio was 1.01 (95% CI 0.81-1.26) for TSH 4.5-6.9 mU/L, 1.65 (0.84-3.23) for TSH 7.0-9.9 mU/L, and 1.86 (1.27-2.72) for TSH 10.0-19.9 mU/L (P for trend <0.01); risks remained similar after adjustment for cardiovascular risk factors. In summary , meta-analysis studies provide fair evidence that subclinical hypothyroidism is associated with cardiovascular morbidity and mortality. The risk on cardiovascular events is apparently dose-dependent: the higher the TSH, the greater the risk, with highest risk in subjects with TSH levels of ≥ 10 mU/L The risk is likely lower in subjects older than 65-70 yr.

9.10.5. TREATMENT

The many studies on the efffect of levothyroxine treatment in subclinical hypothyroidism have yielded inconsistent results. In order to get a more clear picture, a meta-analysis has been performed by the Cochrane Colloboration on studies published until May 2006 1 . The meta-analysis included 12 randomized controlled trials (RCTs) of 6-14 months duration involving 350 people. The daily average L-T4 dose required to normalize TSH in the active group varied between 67.5 to 85.5 μg (range 50 to 125 μg); subclinical hyperthyroidism at the end of the intervention was reported in 2 studies. In the placebo groups spontaneous normalization of TSH occurred in 42%, 25% and 24% of patients in 3 studies; these studies included patients without thyroid disease. The reader may consult the Cochrane review for references to studies evaluated in the meta-analysis 1 . Studies published after the meta-analysis will be referred to separately.

Treatment effects on symptoms and signs . The Cochrane meta-analysis did not observe statistically significant improvement in symptoms, mood or quality of life; one study showed a statistically significant improvement in cognitive function. More recent evaluations include a RCT among subjects of 65 yr and older in the UK (providing no evidence for improvement of cognitive function with L-T4) (2), and two non-RCTs in which attention and memory improved in L-T4 treated subjects relative to controls (3,4). A placebo-controlled randomized clinical trial in the UK demonstrated significant improvement in tiredness upon levothyroxine treatment (35) The Cochrane meta-analysis evaluated many parameters of systolic and diastolic heart function. Significant improvement after L-T4 treatment was observed for some parameters, like isovolumic relaxation time, index of myocardial performance, cycle variation index and left ventricular ejection time; systemic vascular resistance was not improved. Some studies report a lower systolic and diastolic blood pressure upon L-T4 treatment of subclinical hypothyroidism (5,6,7); one study reports a 6% reduction in supine mean arterial pressure (6). More recently, L-T4 treatment apparently improves arterial stiffness as evident from one RCT (8) and three non-RCTs (7,9,10). With regard to endothelium dysfunction, endothelial progenitor cells, expressing both endothelial and stem cell markers, are offered as a novel risk marker of cardiovascular disease. Subclinical hypothyroidism is associated with a lower number of endothelial progenitor cells in peripheral blood; the count increased after L-T4 treatment to values of healthy controls (11). Recent studies also indicate regression of the increased carotid intima-media thickness upon L-T4 treatment as evident from one RCT (12) and three non-RCTs (7,13,14). With respect to serum lipids, an early non-systematic review concluded that normalization of serum TSH in subclinical hypothyroidism decreases serum cholesterol on average by 0.4 mmol/l (95% CI 0.2-0.6 mmol/l) 16 . A subsequent meta-analysis also concluded that normalization of serum TSH decreases serum LDL-cholesterol by 0.26 mmol/l (95% CI 0.12-0.41 mmol/l) 17 , but the analysis has been criticized because inclusion of both observational and randomized studies with some of poor quality. The Cochrane meta-analysis found no significant improvement of total cholesterol upon L-T4 treatment. HDL- as well as LDL-cholesterol also did not reveal significant effects of L-T4 treatment, although a subgroup with LDL values >155 mg/dl showed significant effects. No differences between intervention groups were seen in the outcomes of triglycerides, apolipoprotein A and B and lipoprotein (a). A more recent RCT observed significant decreases in total and LDL cholesterol in the T4-treated subjects as compared to the placebo group (15) The reduction in serum total and LDL cholesterol was larger in individuals with TSH levels > 8.0 mU/L. The influence of subclinical hypothyroidism seems directly proportional to the degree of TSH elevation (18). Higher pretreatment cholesterol levels may also be associated with a greater reduction in total and LDL cholesterol 17,19 . It has been established from large trials outside the thyroid field that cardiovascular disease is reduced by 15% for each 10% reduction in plasma LDL cholesterol or 25% by a 38 mg/dl reduction in plasma total cholesterol. A 10% reduction in cholesterol may reduce the risk of cardiovascular mortality by 20% (1). Recent studies evaluated still other effects of L-T4 treatment in subclinical hypothyroidism. A RCT demonstrated increased cardiopulmonary exercise performance after L-T4 therapy in comparison to no treatment 20 . L-T4 treatment of subjects older than 70 yr with subclinical hypothyroidism did not document any benefit in terms of functional mobility 21 . A non-RCT shows normalization of reduced glomerular filtration rate and increased serum Cystatin-C levels upon L-T4 treatment of subclinical hypothyroidism (7). A RCT in iron-deficient subjects with subclinical hypothyroidism demonstrated greater increase in hemoglobin levels upon treatment with L-T4 plus iron than in treatment with iron alone (22). L-T4 treatment of subclinical hypothyroidism in pregnant women improves maternal and fetal outcomes of pregnancy, and is recommended 23 . A RCT in infertile women reports improved outcomes of in vitro fertilization upon L-T4 treatment as compared to placebo treatment 24 .

Treatment effects on cardiovascular morbidity and mortality, There are no placebo-controlled randomized clinical trials that assess the effect of long-term L-T4 treatment in subclinical hypothyroidism on cardiovascular morbidity or mortality. However, interesting data have emerged from the UK General Practitioner Research Database. Individuals with new subclinical hypothyroidism (TSH 5-10 mU/L) were identified in 2001 with outcomes analysed until March 2009 (34). After a median follow-up of 7.6 yr, 52.8% of the younger subjects (40-70 yr) and 49.9% of the older subjects (>70 yr) with subclinical hypothyroidism were treated with levothyroxine. Incident ischemic heart disease in the younger age group was observed in 4.2% in treated subjects and in 6.6% of the untreated subjects (multivariate-adjusted hazard ratio 0.61, 95% CI 0.39-0.95). In the older age group there were ischemic heart disease events in 12.7% in treated subjects and 10.7% in untreated subjects (hazard ratio 0.99, 95% CI 0.59-1.33). The data suggest that treatment of subclinical hypothyroidism with levothyroxine was associated with fewer ischemic heart disease events in younger individuals, but not in older people.

Recommendations when to treat . Whether or not subclinical hypothyroidism should be treated was and still is hotly debated; there are strong defenders as well as strong opponents to levothyroxine treatment 25-28 . A 2004 scientific review by a panel of experts concluded that data supporting associations of subclinical thyroid disease with symptoms or adverse clinical outcomes or benefits of treatment are few, and that the consequences of subclinical thyroid disease are minimal 29 ; consequently, the panel recommended against routine treatment of subclinical hypothyroidism, albeit recognizing the possible need for treatment in selected individual cases. The 2007 Cochrane meta-analysis also could find no evidence supporting treatment (1). Subsequent meta-analyses of long-term follow-up population-based studies seem to indicate that subclinical hypothyroidism is indeed associated with a modest risk on cardiovascular morbidity and mortality (although this may be age-dependent), but proof that levothyroxine treatment decreases the risk is lacking. This would require appropriately powered, randomized,placebo- controlled, double-blinded interventional trials with long follow-up. The debate whether or not to treat, thus continues (30,31).Given the current state of affairs with a lack of controlled trials reporting on long-term outcome, the decision whether or not to treat has to be taken in the face of uncertainty. This is not rare in medical practice, and the physician copes with such problems by an individualized approach taking into account best available circumvential evidence and clinical judgment.

An algoritm for the individual management of subclinical hypothyroidism is given in figure 9-7.It is recommended to confirm the existence of subclinical hypothyroidism in a second blood sample taken about 3-6 months later. This recommendation in current guidelines is given in view of the high chance of spontaneous normalization of the elevated TSH value 36,37 .It might be useful to already order assay of TPO-Ab and serum lipids in the second sample, because it may be relevant for further management in case subclinical hypothyroidism turns out to be persisitent. If subclinical hypothyroidism is confirmed, a strong case can be made for levothyroxine treatment when TSH values are > 10 mU/L.. This is indeed recommended by one guideline 36 . The other guideline would recommend treatment at TSH >10 mU/L in subjects below the age of 70 yr, and in subjects older than 70 yr only in the presence of symptoms or high cardiovascular risk 37 . The discrepancy between the two guidelines is caused by uncertainty about the role of age: one meta-analysis concludes the association of subclinical hypothyroidism with cardiovascular morbidity and mortality is independent of age, whereas the other concludes the risk is only present in subjects younger than 65-70 yr. The observation that subclinical hypothyroidism might have some survival value in the elderly age group has led both guidelines to the recommendation not to treat elderly individuals with TSH values of 4-10 mU/L. In younger subjects with mild to moderate subclinical hypothyroidism (TSH values between 4 and 10 mU/L) one may opt to institute levothyroxine treatment in the presence of symptoms (in view of the chance that symptoms will improve), TPO-Ab (especially in case TPO-Ab concentration is high with the risk of imminent progression to overt hypothyroidism, or cardiovascular risk factors (in the hope based on circumstantial evidence obtained from population-based association studies and some observational intervention studies to diminish the risk of developing cardiovascular events). If these three conditions are absent, most will agree it is better not to treat. In case no treatment is given, follow-up with regular repeat TSH measurements is indicated. However there is certainly a role for clinical judgement in thes patients. Many practitioners will elect to try replacement therapy in patients with SCH who are symptomatic, especizally in patients under age 70, with careful attention to maintaining TSH in the normal range. For guidelines how to manage subclinical hypothyroidism in infertile women, in pregnant women or in women planning pregnancy, please consult the chapter on Thyroid and Pregnancy.

figure7

Figure 9-7. Algoritm for the individual management of subclinical hypothyroidism.(TPO-Ab = thyroid peroxidase antibodies; risk factors = cardiovascular risk factors).

Table 9-16

Table 9-16. Summary of data on subclinical hypothyroidism.

Prevalence • approximately 4-8% in general population • more common in women than in men • more common in elderly subjects than in young subjects• more common in iodine-replete than in iodine-deficient areas
Natural history • spontaneous normalization of TSH varies from 5% to 50% • progression to overt hypothyroidism in appr. 2%-3% per year, but 4% to 5% in the presence of thyroid antibodies
Treatment • dependent on TSH, TPO antibodies, (desire of) pregnancy, symptoms, age, cardiovascular risk factors

 

9.11 SCREENING FOR HYPOTHYROIDISM

In view of the rather high prevalence of thyroid function disorders and the availability of a suitable screening test in the form of the sensitive TSH assay, the question arises if screening programs are warranted in the general adult population 1 . Case-finding strategies have been employed successfully: previously unknown hypothyroidism was found in 0.64% of middle-age women in connection with screening for cervical carcinoma 2 , and in 0.3% of women attending a primary care unit 3 ; the prevalence of subclinical hypothyroidism in the latter study was 1.2%.Case-finding in women over 40 years of age can be useful. Patients admitted to geriatric units also benefit from routine testing as 2% to 5% have treatable thyroid disease, but patients hospitalized with acute illness do not benefit from routine thyroid function tests due to frequent interference of test results by the sick euthyroid syndrome 4 . The cost-effectiveness of periodic screening for mild thyroid failure has been investigated using a state-transition computer decision model that account for case-finding, medical consequences of mild thyroid failure, and costs of care during 40 years of simulated follow-up 5 . The cost-effectiveness of screening 35-year old patients with a serum TSH assay every 5 years was $ 9223 per QALY (quality-adjusted life year) for women and $22595 for men. The cost-effectiveness compares favorably with other generally accepted prevention programs. The authors recommend screening in the general community for mild hypothyroidism with serum TSH (combined with serum cholesterol) every five years at the age of 35 years 5,6 . An update on screening for thyroid disease in the general adult population, however, argues that the evidence of the efficacy of treatment for subclinical thyroid dysfunction is inconclusive and that large randomized trials are needed to determine the likelihood that treatment will improve the quality of life in otherwise healthy subjects who have mildly elevated TSH levels 7 . On the other hand, the update favors office-based screening to detect overt thyroid dysfunction in women older than 50 years of age: in this group, 1 in 71 women screened would benefit from relief of symptoms. Taken together, the presently available data do not justify yet screening of the healthy adult population for hypothyroidism. It is not recommended by th US Preventive Services Task Forcealtough the American Thyroid Association recommend screening every 5 years in women and men older than 35 years.Case-finding, i.e. testing on patients visiting their physician for unrelated reasons, seems currently the best approach to detect previously unsuspected hypothyroidism; it is especially worthwhile in women over 40 years of age. Screening of all pregnant women on thyroid disorders has not yet been accepted, but is most likely beneficial in view of a recent report indicating better outcomes with universal screening as compared to case finding (12). Table 9-17 provides a useful list of indications for screening for hypothyroidism 8 .

Table 9-17. Indications for screening for hypothyroidism. (Reproduced with permission) 8

ESTABLISHED Congenital hypothyroidism Treatment of hyperthyroidism Neck irradiation Pituitary surgery or irradiation Patients taking amiodarone or lithiumPROBABLY WORTHWHILE Pregnancy Type I diabetes antepartum Previous episode of postpartum thyroiditis Unexplained infertility Women over 40 with non-specific complaints Refractory depression; bipolar affective disorder with rapid cycling 10 Turner’s syndrome; Down’s syndrome Autoimmune Addison’s disease UNCERTAIN Breast cancer 11 DementiaFamily history of autoimmune thyroid disease Obesity Idiopathic oedemaNOT INDICATED Acutely ill patients with no clinical reason to suspect thyroid disease

References

References for Section 9.1

9.1.1. Gull WW: On a cretinoid state supervening in adult life in women. Trans Clin Soc London 1874; 7: 180.

9.1.2. Ord WM: On myxedema, a term proposed to be applied to an essential condition in the "cretinoid" affection occasionally observed in middle-aged women. Medico-Chir Trans 1878; 61: 57.

9.1.3. Reverdin JL: In discussion. Société médicale de Genève. Rev Med Suisse Romande 1882; 2: 539.

9.1.4. Kocher T: Ueber Kropfexstirpation und ihre Folgen. Arch Klin Chir 1883; 29:254.

9.1.5. Report of a Committee of the Clinical Society of London to Investigate the Subject of Myxedema. London, Longmans. Green & Co. Ltd. 1888.

References for Section 9.2

9.2.1. Spencer CA, LoPresti JS, Guttler RB, et al.: Application of a new chemiluminescent thyrotropin assay to subnormal measurements. J Clin Endocrinol Metab 1990; 70: 453-460.

9.2.2. Wiersinga WM. Adult hypothyroidism and myxedema coma. In: DeGroot LJ, Jameson JL (eds): Endocrinology (5th ed.). Philadelphia, WB Saunders Company, 2004, ch. 107

9.2.3. Evered DC, Ormston BJ, Smith PA, et al.: Grades of hypothyroidism. Brit Med J 1973; 1: 657-662.

9.2.4. Ishii H, Inada M, Tanaka K, et al.: Induction of outer and inner ring monodeiodinases in human thyroid gland by thyrotropin. J Clin Endocrinol Metab 1983; 57: 500-505.

9.2.5. Laurberg P: Mechanisms governing the relative proportions of thyroxine and 3,5,3'-triiodothyronine in thyroid secretion.Metabolism 1984; 33: 379-392.

9.2.6. Lum SM, Nicoloff JT, Spencer CA, Kaptein EM: Peripheral tissue mechanism for maintenance of serum triiodothyronine values in a thyroxine-deficient state in man. J Clin Invest 1984; 73: 570-575.

9.2.7. Martino E, Bartalena L, Pinchera A. Central hypothyroidism. In: Braverman LE, Utiger RD (eds): The Thyroid: A Fundamental and Clinical Text (8th ed.). Philadelphia, JB Lippincott Williams & Wilkins, 2000, pp 762-773.

9.2.8. Tunbridge WMG, Evered DC, Hall R, et al.: The spectrum of thyroid disease in the community: the Whickham Survey. Clin Endocrinol 1977; 7: 481-493.

9.2.9. Vanderpump MPJ, Tunbridge WMG, French JM, et al.: The incidence of thyroid disorders in the community: a twenty-year follow-up of the Whickham Survey. Clin Endocrinol 1995; 43: 55-68.

9.2.10. Dos Remedios LV, Weber PM, Feldman R, et al.: Detecting unsuspected thyroid dysfunction by the free thyroxine index. Arch Intern Med 1980; 140: 1045-1049.

9.2.11. Sawin CT, Castelli WP, Hershman JM, et al.: The aging thyroid: Thyroid deficiency in the Framingham study. Arch Intern Med 1985; 145: 1386-1388.

9.2.12. Okamura K. Ueda K, Sone H, et al.: A sensitive TSH assay for screening of thyroid functional disorder in elderly Japanese. J Am Geriatr Soc 1989; 37: 317-322.

9.2.13. Sundbeck G, Lundberg PA, Lindstedt G, et al.: Incidence and prevalence of thyroid disease in elderly women: Results from the longitudinal population study of elderly people in Gothenburg, Sweden. Age and Ageing 1991; 20: 291-298.

9.2.14. Wang C, Crapo LM: The epidemiology of thyroid disease and implication for screening. Endocrinol Metab Clin North Am 1997; 26: 189-218.

9.2.15. Geul KW, van Sluisveld ILL, Grobbee DE, et al.: The importance of thyroid microsomal antibodies in the development of elevated serum TSH in middle-aged women: Associations with serum lipids. Clin Endocrinol 1993; 39: 275-280.

9.2.16. Walsh JP, Bremner AP, Feddema P, et al. Thyrotropin and thyroid antibodies as predictors of hypothyroidism: a 13-year, longitudinal study of a community-based cohort using current immunoassay techniques. J Clin Endocrinol Metab 2010; 95: 1095-1104.

9.2.17. Asvold BO, Vatten LJ, Midthjell, Bjoro T. Serum TSH wthin the reference range as a predictor of future hypothyroidism: 11-year follow-up of the HUNT study in Norway. J Clin Endocrinol Metab 2012; 97: 93-99.

References for Section 9.3.1

9.3.1.1. Faglia G, Bitensky L, Pinchera A, et al.: Thyrotropin secretion in patients with central hypothyroidism: evidence for reduced biological activity of immunoreactive thyrotropin. J Clin Endocrinol Metab 1979; 48: 989-998.

9.3.1.2. Miura Y, Perkel VS, Papenberg KA, et al.: Concanavalin-A, lentil, and ricin lectin affinity binding characteristic of human thyrotropin: differences in the sialylation of thyrotropin in sera of euthyroid, primary, and central hypothyroid patients. J Clin Endocrinol Metab 1989; 69: 985-995.

9.3.1.3. Horimoto M, Nishikawa M, Ishihara T, et al.: Bioactivity of thyrotropin (TSH) in patients with central hypothyroidism: comparison between in vivo 3,5,3'-triiodothyronine response to TSH and in vitro bioactivity of TSH. J Clin Endocrinol Metab 1995; 80: 1124-1128.

9.3.1.4. Samuels MH, Lillehei K, Kleinschmidt-Demasters BK, et al.: Patterns of pulsatile glycoprotein secretion in central hypothyroidism and hypogonadism. J Clin Endocrinol Metab 1990; 70: 391-395.

9.3.1.5. Adriaanse R, Brabant G, Endert E, Wiersinga WM: Pulsatile TSH release in patients with untreated pituitary disease. J Clin Endocrinol Metab 1993; 77: 205-209.

9.3.1.6. Collu R, Tang J, Castagné J, et al.: A novel mechanism for isolated central hypothyroidism: inactivating mutations in the thyrotropin-releasing hormone receptor gene. J Clin Endocrinol Metab 1997; 82: 1361-1365.

9.3.1.7. Dacou-Voutetakis C, Feltquate DM, Drakopoulo M, et al.: Familial hypothyroidism caused by a nonsense mutation in the thyroid-stimulating hormone ß-subunit gene. Am J Hum Genet 1990; 46: 988-993.

9.3.1.8. Hayashizaki Y, Hiraoka Y, Tatsumi K: Deoxyribonucleic acid analyses of five families with familial inherited thyroid stimulating hormone deficiency. J Clin Endocrinol Metab 1990; 71: 792-796.

9.3.1.9. Persani L. Clinical review: Central hypothyroidism: pathogenic, diagnostic, and therapeutic challenges. J Clin Endocrinol Metab 2012; 97: 3068-3078.

9.3.1.10. Arafah BM. Reversible hypopituitarism in patients with large non-functioning pituitary adenomas. J Clin Endocrinol Metab 1986; 62: 1173-1179.

9.3.1.11. Constine LS, Woolf PD, Cann D, et al.: Hypothalamic-pituitary dysfunction after radiation for brain tumors. N Engl J Med 1993; 328: 87-94.

9.3.1.12. Snijder PJ, Fowble BF, Schatz NJ, et al.: Hypopituitarism following radiation therapy of pituitary adenomas. Am J Med 1986; 81: 457-462.

9.3.1.13. Edwards BM, Clark JDA: Post-traumatic hypopituitarism: six cases and a review of the literature. Medicine 1986; 65: 281-290.

9.3.1.14. Cosman F, Post KD, Holub D, Wardlaw SL, et al.: Lymphocytic hypophysitis. Report of 3 new cases and review of the literature. Medicine 1989; 68: 240-256.

9.3.1.15. Kaptein EM, Spencer CA, Kamile MB, Nicoloff JT: Prolonged dopamine administration and thyroid hormone economy in normal and critically ill subjects. J Clin Endocrinol Metab 1980; 51: 387-393.

9.3.1.16. Vagenakis AG, Braverman LE, Azizi F, et al.: Recovery of pituitary thyrotropic function after withdrawal of prolonged thyroid suppression therapy. N Engl J Med 1975; 293: 681-684.

9.3.1.17 . Sherman SI, Gopal J, Haugen BR, Chiu AC, Whaley K, Nowlakha P, Duvic M. Central hypothyroidism associated with retinoid X receptor-selective ligands. New Engl J Med 340:1075-1079, 1999.

9.3.1.18. Oliveira JHA, Persani L, Beck-Peccoz P, et al. Investigating the paradox of hypothyroidism and increased serum thyrotropin (TSH) levels in Sheehan’s syndrome: characterization of TSH carbohydrate content and bioactivity. J Clin Endocrinol Metab 2001; 86: 1694-1699.

9.3.1.19. Bonomi M, Proverbio MC, Weber G, et al. Hyperplastic thyroid gland, high serum glycoprotein α-subunit, and variable circulating thyrotropin (TSH) levels as hallmark of central hypothyroidism due to mutations of the TSHβ gene. J Clin Endocrinol Metab 2001; 86: 1600-1604.

9.3.1.20. Vuissoz JM, Deladoëy J, Buyukgebiz A et al. New autosomal recessive mutation of the TSHβ subunit gene causing central isolated hypothyroidism. J Clin Endocrinol Metab 2001; 86: 4468-4471.

9.3.1.21. Lando A, Holm K, Nysom K, et al. Thyroid function in survivors of childhood acute lymphoblastic leukemia: the significance of prophylactic cranial irradiation. Clin Endocrinol 2001; 55: 21-25.

9.3.1.22. Rose SR. Cranial irradiation and central hypothyroidism. Trends Endocrinol Metab 2001; 12:97-104.

9.3.1.23. Persani L, Ferretti E, Borgato S, et al. Circulating thyrotropin bioactivity in sporadic central hypothyroidism. J Clin Endocrinol Metab 2000; 85:3631.

9.3.1.24. Schmiegelow M, Feldt-Rasmussen U, Rasmussen AK, et al. A Population-based study of thyroid function after radiotherapy and chemotherapy for a childhood brain tumor. J Clin Endocrinol Metab 2003; 88:136-140.

9.3.1.25. Benvenga S, Campenmi A, Ruggeri RM, Trimarchi F. Hypopituitarism secondary to head trauma. J Clin Endocrinol Metab 2000; 85:1353-136.

9.3.1.26. Bellastella A, Bizzarro A, Coronella C, et al. Lymphocytic hypophysitis: a rare or underestimated disease? Eur J Endocrinol 2003; 149:363-376.

9.3.1.27. Schneider HJ, Kreitschmann-Andermahr I, Ghigo E, Stalla GK, Agha A. Hypothalamopituitary dysfunction following traumatic brain injury and aneurysmal subarachnoid hemorrhage. A systematic review. JAMA 2007; 298: 1429-1438.

9.3.1.28. Liu S, Ogilvie KM, Klausing K, et al. Mechanism of selective retinoid X receptor agonist-induced hypothyroidism in the rat. Endocrinology 2002; 143: 2880-2885.

9.3.1.29.Golden WM, Weber KB, Hernandez TL et al. Single-dose rexinoid rapidly and specifically suppresses serum thyrotropin in healthy subjects. J Clin Endocrinol Metab 2007; 92: 124-130.

9.3.1.30. Smit JW, Stokkel MP, Pereira AM et al. Bexarotene induced hypothyroidism: bexarotene stimulates the peripheral metabolism of thyroid hormones. J Clin Endocrinol Metab 2007; 92: 2496-2499.

9.3.1.31. Haugen BR. Drugs that suppress TSH or cause central hypothyroidism. Best Pract Res Clin Endocrinol Metab 2009; 23: 793-800.

9.3.1.32. Sun Y, Bak B, Schoenmakers N et al. Loss-of-function mutations in IGSF1 cause an X-linked syndrome of central hypothyroidism and testicular enlargement. Nat Genet 2012; 44: 1375-1381.

9.3.1.33. Joustra SD, Schoenmakers N, Persani L et al. The IGSF1 deficiency syndrome: characteristics of male and female patients. J Clin Endocrinol Metab 3013; Oct.9 [Epub ahead of print].

References for Section 9.3.2

9.3.2.1. Hayashi Y, Tamai H, Fukata S, et al.: A long-term clinical, immunological, and histological follow-up study of patients with goitrous chronic lymphocytic thyroiditis. J Clin Endocrinol Metab 1985; 61: 1172-1178.

9.3.2.2.. Michaelson ED, Young RL: Hypothyroidism with Graves' disease. JAMA 1970; 23: 1351.

9.3.2.3. Arikawa K, Ichikawa Y, Yoshida T, et al.: Blocking type antithyrotropin receptor antibody in patients with nongoitrous hypothyroidism: its incidence and characteristics of action. J Clin Endocrinol Metab 1985; 60: 953-959

9.3.2.4. Kraiem Z, Lahat N, Glaser B, et al.: Thyrotropin receptor blocking antibodies: incidence, characterization and in vivo synthesis. Clin Endocrinol 1987; 27: 409-421

9.3.2.5. Rieu M, Portos C, Lissak B, et al.: Relationship of antibodies to thyrotropin receptors and to thyroid ultrasonographic volume in euthyroid and hypothyroid patients with autoimmune thyroiditis. J Clin Endocrinol Metab 1996; 80: 641-645.

9.3.2.6. Tomer Y, Huber A. The etiology of autoimmune thyroid disease: a story of genes and environment. J Autoimmun 2009; 32: 231-239.

9.3.2.7. Jacobson EM, Tomer Y. The genetic basis of thyroid autoimmunity. Thyroid 2007; 17: 949-961.

9.3.2.8. Laurberg P, Pedersen KM, Hreidarsson A, et al.: Iodine intake and the pattern of thyroid disorders: a comparative epidemiological study of thyroid abnormalities in the elderly in Iceland and in Jutland, Denmark. J Clin Endocrinol Metab 1998; 83: 765-769.

9.3.2.9. Laurberg P, Cerqueira C, Ovesen L et al. Iodine intake as a determinant of thyroid disorders in populations. Best Pract Res Clin Endocrinol Metab 2010; 24: 13-27.

9.3.2.10. Sundrick RS, Bagchi N, Brown TR: The role of iodine in thyroid autoimmunity: from chickens to humans: a review. Autoimmunity 1992; 13: 61-68.

9.3.2.11. Effraimidis G, Tijssen JG, Wiersinga WM. Discontinuation of smoking increases the risk for developing thyroid peroxidase antibodies and/or thyroglobulin antibodies: a prospective study. J Clin Endocrinol Metab 2009; 94: 1324-1328.

9.3.2.12. Asvold BO, Bjoro T, Nilsen TI, Vatten LJ. Tobacco smoking and thyroid function: a population-based study. Arch Intern Med 2007; 167: 1428-1432.

9.3.2.13. Carle A, Bulow Pedersen I, Knudsen N et al. Smoking cessation is followed by a sharp but transient rise in the incidence of overt autoimmune hypothyroidism – a population-based, case-control study. Clin Endocrinol 2012; 77: 764-772.

References for Section 9.3.3.

9.3.3.1. Nikolai TF: Recovery of thyroid function in primary hypothyroidism. Am J Med Sci 1989; 297: 18-21.

9.3.3.2. Kasagi K, Iwata M, Misaki T, Konishi J. Effect of iodine restriction on thyroid function in patients with primary hypothyroidism. Thyroid 2003; 13: 561-567.

9.3.3.3. Okamura K, Sato K, Ikenoue H, et al.: Reevaluation of thyroidal radioactive iodine uptake test, with special reference to reversible primary hypothyroidism with elevated thyroid radioiodine uptake. J Clin Endocrinol Metab 1988; 67: 720-726.

9.3.3.4. Comtois R, Faucher L, Laflèche L: Outcome of hypothyroidism caused by Hashimoto's thyroiditis. Arch Int Med 1995; 155: 1404-1408.

9.3.3.5. Takasu N, Komiya I, Asawa T, et al.: Test for recovery from hypothyroidism during thyroxine therapy in Hashimoto's thyroiditis. Lancet 1990; 336: 1084-1086.

9.3.3.6. Takasu N, Yamada T, Takasu M, et al.: Disappearance of thyrotropin-blocking antibodies and spontaneous recovery from hypothyroidism in autoimmune thyroiditis. N Eng J Med 1992; 326: 513-518.

9.3.3.7. Kraiem Z, Baron E, Kahana L, et al.: Changes in stimulating and blocking TSH receptor antibodies in a patient undergoing three cycles of transition from hypo to hyperthyroi-dism and back to hypothyroidism. Clin Endocrinol 1992; 36: 211-216.

9.3.3.8. Gerstein HC: How common is postpartum thyroiditis? A methodologic overview of the literature. Arch Int Med 1990; 150: 1397-1400.

9.3.3.9. Kuypens JL, Pop VJ, Vader HL, et al.: Prediction of postpartum thyroid dysfunction: can it be improved? Eur J Endocrinol 1998; 139: 36-43.

9.3.310. Alvarez-Marfany M, Roman SH, Drexler AJ, et al.: Long-term prospective study of postpartum thyroid function in women with insulin dependent diabetes mellitus. J Clin Endocrinol Metab 1994; 79: 10-16.

9.3.3.11. Othman S, Phillips DI, Parkes AB, et al.: A long-term follow-up of postpartum thyroiditis. Clin Endocrinol 1990; 32: 559-564.

9.3.3.12. Kuypens JL, de Haan-Meulman M, Vader HL, et al.: Cell-mediated immunity and postpartum thyroid dysfunction: a possibility for the prediction of disease? J Clin Endocrinol Metab 1998; 83: 1959-1966.

9.3.3.13. Harris B, Othman S, Davies JA, et al.: Association between postpartum thyroid dysfunction and thyroid antibodies and depression. Brit Med J 1992; 305: 152-156.

9.3.3.14. Pop VJ, de Vries E, van Baar A, et al.: Maternal thyroid peroxidase antibodies during pregnancy: a marker of impaired child development? J Clin Endocrinol Metab 1995; 80: 3561-3566.

9.3.3.15. Pop VJ, Kuypens JL, Van Baar AL, et al.: Low maternal free thyroxine concentrations during early pregnancy are associated with impaired psychomotor development in infancy. Clin Endocrinol 1999; 50: 149-156.

9.3.3.16. Henrichs J, Bongers-Schokking JJ, Schenk JJ. Matrenal thyroid function during early pregnancy and cognitive functioning in early childhood: the Generation R study. J Clin Endocrinol Metab 2010; 95: (doi:10.2010/jc.2010-0415)

9.3.3.17. Vialettes B, Guillerand MA, Viens P, et al.: Incidence rate and risk factors for thyroid dysfunction during recombinant interleukin-2 therapy in advanced malignancies. Acta Endocrinol 1993; 129: 31-38.

9.3.3.18. Preziati D, La Rosa L, Covini G, et al.: Autoimmunity and thyroid function in patients with chronic active hepatitis treated with recombinant interferon alpha-2a. Europ J Endocrinol 1995; 132: 587-593.

9.3.3.19. Marazuela M, Garcia-Buey L, Gonzalez-Fernandez B, et al.: Thyroid autoimmune disorders in patients with chronic hepatitis C before and during interferon-a therapy. Clin Endocrinol 1996; 44: 635-642.

9.3.3.20. Prummel MF, Laurberg P. Interferon- α and autoimmune thyroid disease. Thyroid 2003; 13:547-551.

9.3.3.21. Mandac JC, Chaudrhy S, Sherman KE, Tomer Y. The clinical and physiological spectrum of interferon-alpha inducedthyroiditis: toward a new classification. Hepatology 2006; 43: 661-672.

9.3.3.22. Roti E, Minelli R, Giuberti T et al. Multiple changes in thyroid function in patients with chronic active HCV hepatitis treated with recombinant interferon-alpha. Am J Med 1996; 101: 482-487.

9.3.3.23. Tomer Y, Menconi F. Interferon induced thyroiditis. Best Pract Res Clin Endocrinol Metab 2009; 23: 703-712.

9.3.3.24. Akeno N, Blackard JT, Tomer Y. HCV E2 protein binds directly to thyroid cells and induces IL-8 production: a new mechanism for HCV induced thyroid autoimmunity. J Autoimmun 2008; 31: 339-344.

9.3.3.25. Baufin E, Marcellin P, Pouteau M et al. Reversibility of thyroid dysfunction induced by recombinant alpha interferon in chronic hepatitis C. Clin Endocrinol 1993; 39: 657-661.

9.3.3.26. Akeno N, Smith EP, Stefan M et al. IFN-α mediates the development of autoimmunity both by direct tissue toxicity and through immune cell recruitment mechanisms.J Immunol 2011; 186: 4693-4706.

References for Section 9.3.4.

9.3.4.1. Nofal MN, Beierwaltes WH, Patno ME: Treatment of hyperthyroidism with sodium iodide I-131, a 16-year experience. JAMA 1966; 197: 605-610.

9.3.4.2. Hagen GA, Ouellette RP, Chapman EM: Comparison of high and low dosage of 131I in the treatment of thyrotoxicosis. N Engl J Med 1967; 277: 559.

9.3.4.3. Roudebush CP, Hoye KE, DeGroot LJ: Compensated low-dose 131I therapy of Graves' disease. Ann Intern Med 1977; 87: 441.

9.3.4.4. Nygaard B, Hegedüs L, Gervil M, et al.: Influence of compensated 131I therapy on thyroid volume and incidence of hypothyroidism in Graves' disease. J Int Med 1995; 238: 491-497.

9.3.4.5. Stoffer SS, Hamburger JI: Inadvertent 131I therapy for hyperthyroidism in the first trimester of pregnancy. J Nucl Med 1976; 17: 146.

9.3.4.6. Huysmans DA, Corstens FH, Kloppenborg PW: Long-term follow-up in toxic solitary autonomous thyroid nodules treated with radioactive iodine. J Nucl Med 1991; 32: 27-30.

9.3.4.7. Nygaard B, Hegedüs L, Nielsen KG, et al.: Long-term effect of radioactive iodine on thyroid function and size in patients with solitary autonomously functioning toxic thyroid nodules. Clin Endocrinol 1999; 50: 197-202.

9.3.4.8. Smith RE, Adler RA, Clark P, et al.: Thyroid function after mantle irradiation in Hodgkin's disease. JAMA 1981; 245: 46-49.

9.3.4.9. Tell R, Sjödin H, Lundell G, et al.: Hypothyroidism after external radiotherapy for head and neck cancer. Int J Radiation Oncology Biol Phys 1997; 39: 303-308.

9.3.4.10. Smolarz K, Malke G, Voth E, et al. Hypothyroidism after therapy for larynx and pharynx carcinoma. Thyroid 2000; 10: 425-429.

9.3.4.11. Mercado G, Adelstein DJ, Saxton JP, et al. Hypothyroidism. A frequent event after radiotherapy and after radiotherapy with chemotherapy for patients with head and neck carcinoma. Cancer 2001; 92:2892-2897.

9.3.4.12. Ishiguro H, Yasuda Y, Tomita Y et al. Long-term follow-up of thyroid function in patients who received bone marrow transplantation during childhood and adolescence. J Clin Endocrinol Metab 2004; 89: 5981-5986.

References for Section 9.3.5.

9.3.5.1. Chau AM, Lynch MJG, Bailey JD et al.: Hypothyroidism in cystinosis: a clinical, endocrinologic and histologic study involving sixteen patients with cytinosis. Am J Med 1970; 48: 678.

9.3.5.2. Barsano CP: Other forms of hypothyroidism. In: Braverman LE, Utiger RD (eds): The Thyroid: A Fundamental and Clinical Text (7th ed.). Philadelphia, JB Lippincott, 1996, pp 768-778.

9.3.5.3. Paes JE, Burman KD, Cohen J et al. Acute bacterial suppurative thyroiditis: a clinical review and expert opinion. Thyroid 2010; 20: 247-256.

References for Section 9.3.7.

9.3.7.1. Wiersinga WM, Touber JL, Trip MD, van Royen EA: Uninhibited thyroidal uptake of radioiodine despite iodine excess in amiodarone-induced hypothyroidism. J Clin Endocrinol Metab 1986; 63: 485-491.

9.3.7.2. Braverman LE: Iodine and the thyroid: 33 years of study. Thyroid 1994; 4: 351-356.

9.3.7.3. Markou K, Georgopoulos N, Kyriazopoulou V, Vagenakis AG. Iodine-induced hypothyroidism. Thyroid 2001; 11: 501-510.

9.3.7.4. Trip MD, Wiersinga WM, Plomp TA: Incidence, predictability, and pathogenesis of amiodarone-induced thyrotoxicosis and hypothyroidism. Am J Med 1991; 91: 507-511.

9.3.7.5. Martino E, Safran M, Aghini-Lombardi F, et al.: Environmental iodine intake and thyroid dysfunction during chronic amiodarone therapy. Ann Int Med 1984; 101: 28-34.

9.3.7.6. Pedersen IB, Laurberg P, Knudsen N et al. An increased incidence of overt hypothyroidism after iodine fortification of salt in Denmark: a propsective population study. J Clin Endocrinol Metab 2007; 92: 3122-3127.

References for Section 9.3.8.

9.3.8.1. Braverman LE, Woeber KA, Ingbar SH: Induction of myxedema by iodide in patients euthyroid after radioiodine or surgical treatment of diffuse toxic goiter. N Engl J Med 1969; 281: 816.

9.3.8.2. Braverman LE, Ingbar SH, Vagenakis AG, et al.: Enhanced susceptibility to iodide myxedema in patients with Hashimoto's thyroiditis. J Clin Endocrinol 1971; 32: 515

9.3.8.3. Berens SC, Bernstein RS, Robbins J, Wolff J: Antithyroid effects of lithium. J Clin Invest 1970; 49: 1357.

9.3.8.4. Perrild H, Hegedüs L, Baastrup PC, et al.: Thyroid function and ultrasonically determined thyroid size in patients receiving long-term lithium treatment. Am J Psychiatry 1990; 147: 1508-1521.

9.3.8.5. Gaitan E, Cooksey RC, Legan J, et al.: Antithyroid and goitrogenic effects of coal-water extracts from iodine-sufficient goiter areas. Thyroid 1993; 3 49-53.

9.3.8.6. Langer P, Tajtakova M, Fodor G, et al.: Increased thyroid volume and prevalence of thyroid disorders in an area heavily polluted by polychlorinated biphenyls. Eur J Endocrinol 1998; 139: 402-409.

9.3.8.7. Braverman LE, He X, Pino S et al. The effect of perchlorate, thiocyanate, and nitrate on thyroid function in workers exposed to perchlorate long-term. J Clin Endocrinol Metab 2005; 90: 700-706.

9.3.8.8. Braverman LE, Pearce EN, He X et al. Effects of six months of daily low-dose perchlorate exposure on thyroid function in healthy volunteers. J Clin Endocrinol Metab 2006; 91: 2721-2724.

9.3.8.9. Pearce EN, Lazarus JH, Smyth PP et al. Perchlorate and thiocyanate exposure and thyroid function in first-trimester pregnant women. J Clin Endocrinol Metab 2010; 95: 3207-3215.

9.3.8.10. Illouz F, Laboureau-Soares S, Dubois S, Rohmer V, Rodien P. Tyrosine kinase inhibitors and modifications of thyroid function tests: a review. Eur J Endocrinol 2009; 160: 331-336.

9.3.8.11. Desai J, Yassa L, Marqusee E et al. Hypothyroidism after sunitinib treatment for patients with gastrointestinal stromal tumors. Ann Intern Med 2006; 145: 660-664.

9.3.8.12. Wolter P, Stefan C, Decallonne B et al. The clinical implications of sunitinib-induced hypothyroidism: a propsective evaluation. Brit J Cancer 2008; 99: 448-454.

9.3.8.13. Mannavola D, Coco P, Vannucchi G et al. A novel tyrosine kinase selective inhibitor, sunitinib, induces transient hypothyroidism by blocking iodine uptake. J Clin Endocrinol Mrtab 2007; 92: 3531-3534.

9.3.8.14. Salem AK, Fenton MS, Marion KM, Hershman JM. Effect of sunitinib on growth and function of FRTL-5 thyroid cells. Thyroid 2008; 18: 631-635.

9.3.8.15. Wong E, Rosen LS, Mulay M et al. Sunitinib induces hypothyroidism in advanced cancer patients and may inhibit thyroid peroxidase activity. Thyroid 2007; 17: 351-355.

9.3.8.16. Grossmann M, Premaratne E, Desai J, Davis ID. Thyrotoxicosis during sunitinib treatment for renal cell carcinoma. Clin Endocrinol 2008; 69: 669-672.

9.3.8.17. Rogiers A, Wolter P, Op de Beeck K, Thijs M, Decallone B, Schoffski P. Shrinkage of thyroid volume in sunitinib treated patients with renal cell carcinoma – a potential marker of irreversible thyroid dysfunction? Thyroid 2010; 20: 317-322.

9.3.8.18. Makita N, Miyakawa M, Fujita T, Iiri T. Sunitinib induces hypothyroidism with a markedly reduced vascularity. Thyroid 2010; 20: 323-326.

9.3.8.19. Hershman JM. How does sunitinib cause hypothyroidism? Thyroid 2010; 20: 243-244.

9.3.8.20. Tamaskar I, Bukowski R, Elson P et al. Thyroid function test abnormalities in patients with metastatic renal cell carcinoma treated with sorafenib. Ann Oncol 2008; 19: 265-268.

9.3.8.21. Miyake H, Kurahashi T, Yamanaka K et al. Abnormailties of thyroid function in Japanese patients with metastatic renal cell carcinoma treated with sorafenib: a prospective evaluation. Urol Oncol 2009; Nov.12 (Epub ahead of print)

9.3.8.22. Makita N, Iiri T. Tyrosine kinase inhibitor-induced thyroid disorders: a review and hypothesis. Thyroid 2013; 23: 151-159.

9.3.8.23. Torino F, Barnabei A, Paragliola R, Baldelli R, Appetecchia M, Corsello SM. Thyroid dysfunction as an unintended side effect of anticancer drugs. Thyroid 2013; 23: 1345-1366.

References for Section 9.3.9

9.3.9.1. Huang SA, Tu HM, Harney JW, et al. Severe hypothyroidism caused by type 3 iodothyronine deiodinase in infantile hemangiomas. New Engl J Med 2000; 343: 185-189.

9.3.9.2. Huang SA, Fish SA, Dorfman DM et al. A 21-year-old woman with consumptive hypothyroidism due to a vascular tumor expressing type 3 iodothyronine deiodinase. J Clin Endocrinol Metab 2002; 87: 4457-4461.

9.3.9.3. Ruppe MD, Huang SA, Jan de Beur SM. Consumptive hypothyroidism caused by paraneoplastic production of type 3 iodothyroinine deiodinase. Thyroid 2005; 15: 1369-1372.

9.3.9.4. Howard D, La Rosa FG, Huang S et al. Consumptive hypothyroidism resulting from hepatic vascular tumors in an athyreotic adult. J Clin Endocrinol Metab 2011; 96: 1966-1970.

References for Section 9.4.

9.4.1. Smith TJ, Bahn RS, Gorman CA: Connective tissue, glycosaminoglycans, and diseases of the thyroid. Endocr Rev 1989; 10: 366-391.

9.4.2. Smith TJ, Horwitz AL, Refetoff S. The effect of thyroid hormone on glycosaminoglycan accumulation in human skin fibroblasts. Endocrinology 1981; 108: 2397.

9.4.3. Smith TJ, Murata Y, Korwitz AL, et al. Regulation of glycosaminoglycan synthesis by thyroid hormone in vitro. J Clin Invest 1982; 70: 1066.

9.4.4. Klion FM, Segal R, Schaffner F: The effect of altered thyroid function on the ultrastructure of the human liver. Am J Med 1971; 150: 137.

9.4.5. Wilkens L: Epiphysial dysgenesis associated with hypothyroidism. Am J Dis Child 1941; 61: 13.

9.4.6. Price TR, Netsky MG: Myxedema and ataxia: Cerebellar alterations and "neural myxedema bodies". Neurology 1966; 16: 957.

9.4.7. Rossman NP: Neurological and muscular aspects of thyroid dysfunction in childhood. Pediatr Clin North Am 1976; 23: 575.

9.4.8. Naeye RL: Capillary and venous lesions in myxedema. Lab Invest 1963; 12: 465.

9.4.9. Salomon MI, DiScala V, Grisham E et al: Renal lesions in hypothyroidism: a study based on kidney biopsies. Metabolism 1967; 16: 846.

9.4.10. Ezrin C, Swanson HE, Humphrey JG et al: The cells of the human adenohypophysis in thyroid disorders. J Clin Endocrinol Metab 1959; 19: 958.

9.4.11. Yamada T, Tsukui T, Ikejiri K et al: Volume of sella turcica in normal subjects and in patients with primary hypothyroidism and hyperthyroidism. J Clin Endocrinol Metab 1976; 42: 817.

9.4.12. Bloodworth JMB, Kirkendall WM, Carr TL: Addison's disease associated with thyroid insufficiency and atrophy (Schmidt syndrome). J Clin Endocrinol Metab 1954; 14: 540.

References for Section 9.5.1.

9.5.1.1. Diekman MJM, Romijn JA, Endert E, Sauerwein H, Wiersinga WM. Thyroid hormones modulate serum leptin levels: observations in thyrotoxic and hypothyroid women. Thyroid 1998; 8: 1081-1086.

9.5.1.2. Hsieh CJ, Wang PW, Wong ST et al. Serum leptin concentrations of patients with sequential thyroid function changes. Clin Endocrinol 2002; 57: 29-34.

9.5.1.3. Iglesias P, Alvarez Fidalgo P, Codoceo R, Diez JJ. Serum concentrations of adipocytokines in patients with hyperthyroidism and hypothyroidism before and after control of thyroid function. Clin Endocrinol 2003; 59: 621-529.

9.5.1.4. Brackhik M, Marcisz C, Giebel S, Orzel A. Serum leptin and ghrelin levels in premenopausal women with stable body mass index during treatment of thyroid dysfunction. Thyroid 2008; 18: 545-550.

9.5.1.5. Gimenez-Palop O, Gimenez-Perez G, Mauricio D et al. Circulating ghrelin in thyroid dysfunction is related to insulin resistance and not to hunger, food intake or anthropometric changes. Eur J Endocrinol 2005; 153: 73-79.

9.5.1.6. Gjedde S, Vestergaard ET, Gormsen LC et al. Serum ghrelin levels are increased in hypothyroid patients and become normalized by L-thyroxine treatment. J Clin Endocrinol Metab 2008; 93: 2277-2280.

9.5.1.7. Altinova AE, Toruner FB, Akturk M et al. Adiponectin levels and cardiovascular risk factors in hypothyroidism and hyperthyroidism. Clin Endocrinol 2006; 65: 530-535.

9.5.1.8. Krassas GE, Pontikides N, Loustis K, Koliakos G, Constantinidis T, Kaltsas T. Resistin levels are normal in hypothyroidism and remain unchanged after attainment of euthyroidism: relationship with insulin levels and anthropometric parameters. J Endocrinol Invest 2006; 29: 606-612.

9.5.1.9. Kokkinos A, Mourouzis I, Kyriaki D, Pantos C, Katsilambros N, Cokkinos DV. Possible implications of leptin, adiponectin and ghrelin in the regulation of energy homeostasis by thyroid hormone. Endocrine 2007; 32: 30-32.

9.5.1.10. Lewallen CG, Rall JE, Berman M: Studies of iodoalbumin metabolism. II. The effects of thyroid hormone. J Clin Invest 1959; 38: 88.

9.5.1.11. Smith TJ, Murata Y, Horwitz AL, et al.: Regulation of glycosaminoglycan accumulation by thyroid hormone in vitro. J Clin Invest 1982; 70: 1066.

9.5.1.12. Smith TJ, Horwitz AL, Refetoff S: The effect of thyroid hormone on glycosaminoglycan accumulation in human skin fibroblasts. Endocrinology 1981; 108: 2397.

9.5.1.13. Crispell KR, Parson W, Hollifield G: A study of the rate of protein synthesis before and during the administration of L-triiodothyronine to patients with myxedema and healthy volunteers using N-15 glycine. J Clin Invest 1956; 35: 164.

9.5.1.14. Lamberg BA, Gräsbeck R: The serum protein pattern in disorders of thyroid function. Acta Endocrinol 1955; 19: 91.

9.5.1.15. Dimitriadis G, Mitrou P, Lambadiari V et al. Insulin action in adipose tissue and muscle in hypothyroidism. J Clin Endocrinol Metab 2006; 91: 4930-4937.

9.5.1.16. Brauman H, Corvilain J: Growth hormone response to hypoglycemia in myxedema. J Clin Endocrinol Metab 1968; 28: 301.

9.5.1.17. Shah JH, Motto GS, Papagiannes E, William GA: Insulin metabolism in hypothyroidism. Diabetes 1975; 24: 922.

9.5.1.18. Maratou E, Hadjidakis DJ, Kollias A et al. Studies of insulin resistance in patients with clinical and subclinical hypothyroidism. Eur J Endocrinol 2009; 160: 785-790.

9.5.1.19. Shah JH, Cerchio GM: Hypoinsulinemia of hypothyroidism. Arch Intern Med 1973; 132: 657.

9.5.1.20. Erdogan M, Canataroglu A, Ganidagli S, Kulaksizoglu M. Metabolic syndrome prevalence in subclinic and overt hypothyroid patients and the relation among metabolic syndrome parameters. J Endocrinol Invest 2010, Jul 22 (Epub ahead of priint).

9.5.1.21. Handisurya A, Pacini G, Tura A, Gessi A, Kautzky-Willer A. Effects of T4-replacement therapy on glucose metabolism in subjects with subclinical and overt hypothyroidism. Clin Endocrinol 2008; 69: 963-969.

9.5.1.22. Hecht A, Gershberg H: Diabetes mellitus and primary hypothyroidism. Metabolism 1968; 17: 108.

9.5.1.23. Abrams JJ, Grundy SM: Cholesterol metabolism in hypothyroidism and hyperthyroidism in man. J Lipid Res 1981; 22: 323.

9.5.1.24. Kritchevsky D: Influence of thyroid hormones and related compounds on cholesterol biosynthesis and degradation: a review. Metabolism 1960; 9: 984.

9.5.1.25. Kurland GS, Lucas JL, Freedberg AS: The metabolism of intravenously infused 14C-labeled cholesterol in euthyroidism and myxedema. J Lab Clin Med 1961; 57: 574.

9.5.1.26. Walton KW, Scott PJ, Dykes PW, Davies JWL: The significance of alterations in serum lipids in thyroid dysfunction. II. Alterations of the metabolism and turnover of 131I low-density lipoproteins in hypothyroidism and thyrotoxicosis. Clin Sci 1965; 29: 217.

9.5.1.27 Chait A, Bierman EL, Albers JJ: Regulatory role of triiodothyronine in the degradation of low density lipoprotein by cultered human skin fibroblasts. J Clin Endocrinol Metab 1979; 48: 887.

9.5.1.28. Scarabottolo L, Trezzi E, Roma P, Gatapano AL: Experimental hypothyroidism modulates the expression of the low density lipoprotein receptor by the liver. Atherosclerosis 1986; 59: 329.

9.5.1.29. Thompson GR, Soutar AK, Spengel FA, et al: Defects of receptor mediated low density lipoprotein catabolism in homozygous familial hypercholesterolemia and hypothyroidism in vivo. Proc Natl Acad Sci USA 1981; 78: 2591.

9.5.1.30. Diekman T, Demacker PNM, Kastelein JJP, et al.: Increased oxidizability of low-density lipoproteins in hypothyroidism. J Clin Endocrinol Metab 1998; 83: 1752-1755.

9.5.1.31. Pazos F, Alvarez JJ, Rubies-Prat J, et al.: Long term thyroid replacement therapy and levels of lipoprotein (a) and other lipoproteins. J Clin Endocrinol Metab 1995; 80: 562-566.

9.5.1.32. O'Brien T, Katz K, Hodge D, et al.: The effect of treatment of hypothyroidism and hyperthyroidism on plasma lipids and apolipoproteins AI, AII and E. Clin Endocrinol 1997; 46: 17-20.

9.5.1.33. Martinez-Triguero ML, Hernandez-Myares A, Nguyen TT, et al.: Effect of thyroid hormone replacement on lipoprotein (a), lipids, and apolipoproteins in subjects with hypothyroidism. Mayo Clin Proc 1998; 73: 837-841.

9.5.1.34. Dullaart RPF, Hoogenberg K, Groener JEM, et al.: The activity of cholesterol ester transfer protein is decreased in hypothyroidism: a possible contribution to alterations in high-density lipoproteins. Europ J Clin Invest 1990; 20: 581-587.

9.5.1.35. Tan KCB, Shiu SWM, Kung AWC: Plasma cholesteryl ester transfer protein activity in hyper- and hypothyroidism. J Clin Endocrinol Metab 1998; 83: 140-143.

9.5.1.36. Diekman MJM, Anghelescu N, Endert E, Bakker O, Wiersinga WM. Changes in plasma low-density lipoprotein (LDL)- and high-density lipoprotein cholesterol in hypo- and hyperthyroid patients are related to changes in free thyroxine, not to polymorphisms in LDL receptor or cholesterol ester transfer protein genes. J Clin Endocrinol Metab 2000; 85: 1857-1862.

9.5.1.37. Gjedde S, Gormsen LC, Rungby J et al. Decreased lipid intermediate levels and lipid oxidation rates despite normal lipolysis in patients with hypothyroidism. Thyroid 2010; 20: 843-849.

9.5.1.38. Tanaci N, Ertugrul DT, Sahin M et al. Postprandial lipemia as a risk factor for cardiovascular disease in patients with hypothyroidism. Endocrine 2006; 29: 451-456.

9.5.1.39. Makino M, Oda N, Miura N, et al. Effect of eicosapentaenoic acid ethyl ester on hypothyroid function. J Endocrinol 2001; 171: 259-265.

9.5.1.40. Ito M, Arishima T, Kudo T et al. Effect of levothyroxine replacement on non-high-density lipoprotein cholesterol in hypothyroid patients. J Clin Endocrinol Metab 2007; 92: 608-611.

9.5.1.41. Ito M, Takamutsu J, Matsuo T, et al. Serum concentrations of remnant-like particles in hypothyroid patients before and after thyroxine replacement. Clin Endocrinol 2003; 58:621-626.

9.5.1.42. Weintraub M, Grosskopf I, Trostanesky Y, et al. Thyroxine replacement enhances clearance of chylomicron remnants in patients with hypothyroidism. J Clin Endocrinol Metab 1999; 84: 2532-2536.

9.5.1.43. Pearce EN, Wilson PW, Yang Q, Vasan RS, Braverman LE. Thyroid function and lipid subparticle sizes in patients with short-term hypothyroidism and a population-based cohort. J Clin Endocrinol Metab 2008; 93: 888-894.

9.5.1.44. Baskol G, Atmaca H, Tanriverdi F, Baskol M, Kocer D, Bayram F. Oxidative stress and enzymatic antioxidant status in patients with hypothyroidism before and after treatment. Exp Clin Endocrinol Diab 2007; 115: 522-526.

9.5.1.45.Torun AN, Kulaksizoglu S, Kulaksizoglu M, Pamuk BO, Isbilen E, Tutuncu NB. Serum total antioxidant status and lipid peroxidation marker malondialdehyde levels in overt and subclinical hypothyroidism. Clin Endocrinol 2009; 70: 469-474.

9.5.1.46. Nanda N, Bobby Z, Hamide A, Koner BC, Sridhar MG. Association between oxidative stress and coronary lipid risk factors in hypothyroid women is independent of body mass index. Metabolism 2007; 56: 1350-1355.

9.5.1.47. Gjedde S, Gormsen LC, Riis AL et al. Reduced expression of uncoupling protein 2 in adipose tissue in patients with hypothyroidism. J Clin Endocrinol Metab 2010; 95: 3537-3541.

9.5.1.48. Teixeira PF, Cabral MD, Silva NA et al. Serum leptin in overt and subclinical hypothyroidism: effect of levothyroxine treatment and relationship to menopausal status and body composition. Thyroid 2009; 19: 443-450.

9.5.1.49. Kaplan O, Uzum AK, Aral H et al. Unchanged serum adipokine concentrations in the setting of short –term thyroidectomy-induced hypothyroidism. Endocr Pract 2012; 18: 887-893.

9.5.1.50. Kosowicz J, Baumann-Antczak A, Ruchala M, Grycznska M, Gurgul E, Sowinski J. Thyroid hormones affect plasma ghrelin and obestatin levels. Horm Metab Res 2011; 43: 121-125.

9.5.1.51. Guzel S, Seven A, Guzel EC, Buyuk B, Celebi A, Aydemir B. Visfatin, leptin, and TNF-α: interrelated adipokines in insulin-resistant clinical and subclinical hypothyroidism. Endocr Res 2013: [Epub ahead of print].

9.5.1.52. Goldberg IJ, Huang LS, Huggins LA et al. Thyroid hormone reduces cholesterol via a non-LDL receptor-mediated pathway. Endocrinology 2012; 153: 5143-5149.

9.5.1.53. Lin JZ, Martagon AJ, Hsueh WA et al. Thyroid hormone receptor agonists reduce serum cholesterol independent of the LDL receptor. Endocrinology 2012; 153: 6136-6144.

References for Section 9.5.2.

9.5.2.1. Smith TJ, Murata Y, Horwitz AL et al: Regulation of glycosaminoglycan accumulation by thyroid hormone in vitro. J Clin Invest 1982; 70 1066.

9.5.2.2. Aikawa JK: The nature of myxedema: Alterations in the serum electrolyte concentrations and radiosodium space and in the exchangeable sodium and potassium contents. Ann Intern Med 1956; 44: 30.

9.5.2.3. Crispell KR, Williams GA, Parson W, Hollifeld G: Metabolic studies in myxedema following administration of l-triiodothyronine: (1) Duration of negative nitrogen balance: (2) effect of testosterone proprionate: (3) comparison with nitrogen balance in a healthy volunteer. J Clin Endocrinol Metab 1957; 17: 221.

9.5.2.4. Smith TJ, Horwitz AL, Refetoff S: The effect of thyroid hormone on glycosaminoglycan accumulation in human skin fibroblasts. Endocrinology 1981; 108: 2397.

9.5.2.5. Smith TJ, Bahn RS, Gorman CA. Connective tissue, glycosaminoglycans, and diseases of the thyroid. Endocr Rev 1989; 10: 366-391.

9.5.2.6. Boelaert K, Newby PR, Simmonds MJ et al. Prevalence and relative risk of other autoimmune diseases in subjects with autoimmune thyroid disease. Am J Med 2010; 123: 183e1-183e9.

9.5.2.7. Artantas S, Gul U, Kilic A, Guler S. Skin findings in thyroid diseases. Eur J Intern Med 2009; 20: 158-161.

9.5.2.8. Levy Y, Segal N, Weintrob N, Danon YL. Chronic urticaria: association with thyroid autoimmunity. Arch Dis Child 2003; 88: 517-519.

References for Section 9.5.3.

9.5.3.1. Smith CD, Ain KB: Brain metabolism in hypothyroidism studied with 31P magnetic-resonance spectroscopy. Lancet 1995; 345: 619-620.

9.5.3.2. Dugbartey AT: Neurocognitive aspects of hypothyroidism. Arch Int Med 1998; 158: 1413-1418.

9.5.3.3. Jellinek EH, Kelly RE: Cerebellar syndrome in myxoedema. Lancet 1960; 2: 225.

9.5.3.4. Sanders V: Neurological manifestations of myxedema. N Engl J Med 1962; 266: 577, 599

9.5.3.5. Cremer GM, Goldstein NP, Paris J: Myxedema and ataxia. Neurology 1969; 19: 37.

9.5.3.6. Crevasse LE, Logue RB: Peripheral neuropathy in myxedema. Ann Intern Med 1959;50: 1433.

9.5.3.7. Nickel SN, Frame B: Nervous and muscular systems in myxedema. J Chronic Dis 1961; 14: 570.

9.5.3.8. Murray IPC, Simpson JA: Acroparaesthesia in myxoedema. Lancet 1958; 1: 1360.

9.5.3.9. Bland JH, Frymoyer JW: Rheumatoid syndrome of myxedema. N Engl J Med 1970; 282: 1171.

9.5.3.10. Frymoyer JW, Bland JH: Carpal-tunnel syndrome in patients with myxedematous arthropathy. J Bone Joint Surg 1973; 55A: 78.

9.5.3.11. Mra Z, Wax MK. Effects of acute thyroxin depletion on hearing in humans. The Laryngoscope 1999; 109: 343-350.

9.5.3.12. Cleare AJ, McGregor A, O'Keane V: Neuroendocrine evidence for an association between hypothyroidism, reduced central 5-HT activity and depression. Clin Endocrinol 1995; 43: 713-719.

9.5.3.13. Aronson R, Offman HJ, Joffe RT, Naylor CD: Triiodothyronine augmentation in the treatment of refractory depression. Arch Gen Psychiatry 1996; 53: 842-848.

9.5.3.14. Murray IPC: The reaction time in myxoedema.Lancet 1956; 2: 384.

9.5.3.15. Kapur VK, Koepsell TD, de Maine J, et al.: Association of hypothyroidism and obstructive sleep apnea. Am J Resp Crit Care Med 1998; 158: 1379-1383.

9.5.3.16. Tachman ML, Guthrie GP Jr: Hypothyroidism: diversity of presentation. Endocr. Rev 1984; 5: 456.

9.5.3.17. Loosen PT: Thyroid function in affective disorders and alcoholism. Endocrinol Metab Clin North Am 1988; 17: 55.

9.5.3.18. Clarnette RM, Peterson CJ: Hypothyroidism: does treatment cure dementia? J Geriatr Psychiat Neurol 1994; 7: 23-27.

9.5.3.19. Scheinberg P, Stead EA, Braman ES, Warren JV: Correlative observations on cerebral metabolism. J Clin Invest 1950; 29: 1139.

9.5.3.20. Sensenbach W, Madison L, Eisenberg S, Ochs L: The cerebral circulation and metabolism in hyperthyroidism and myxedema. J Clin Invest 1954; 33: 1434

9.5.3.21. O'Brien MD, Harris PWR: Cerebral-cortex perfusion rates in myxedema. Lancet 1965; 1: 1170.

9.5.3.22. Duyff RF, Bosch J vd, Laman DM, et al. Neuromuscular findings in thyroid dysfunction: a prospective clinical and electrodiagnostic study. J Neurol Neurosurg Psychiatry 2000; 68: 750-755.

9.5.3.23. Gunarsson T, Sjöberg S, Eriksson M, Nordin C. Depressive symptoms in hypothyroid disorder with some observations on biochemical correlates. Neuropsychobiology 2001; 43: 70-74.

9.5.3.24. Kinuya S, Michigiski T, Tonami N, et al. Reversible cerebral hypoperfusion observed with Tc-99m HMPAO SPECT in reversible dementia caused by hypothyroidism. Clin Nucl Med 1999; 24: 666-668.

9.5.3.25. Constant EL, Volder AG de, Ivanoiu A, et al. Cerebral blood flow and glucose metabolism in hypothyroidism: a positron emission tomography study. J Clin Endocrinol Metab 2001; 86: 3864-3870.

9.5.3.26. Chong JY, Rowland LP, Utiger RD. Hashimoto encephalopathy: syndrome or myth? Arch Neurol 2003; 60: 164-171.

9.5.3.27. Fatourechi V. Hashimoto’s encephalopathy: myth or reality? An endocrinologist’s perspective. Best Pract Res Clin Endocrinol Metab 2005; 19: 53-66.

9.5.3.28. Burmeister LA, Ganguli M, Dodge HH, et al. Hypothyroidism and cognition: preliminary evidence for a specific defect in memory. Thyroid 2001; 11:1177-1185.

9.5.3.29. Miller KJ, Parsons TD, Whybrow PC et al. Verbal memory retrieval deficits associated with untreated hypothyroidism. J Neuropsychiatry Clin Neurosci 2007; 19: 132-136.

9.5.3.30. Anjana Y, Tandon OP, Vaney N, Madhu SV. Cognitive status in hypothyroid female patients: event-related evoked potential study. Neuroendocrinology 2008; 88: 59-66.

9.5.3.31. Correia N, Mullally S, Cooke G et al. Evidence for a specific defect in hippocampal memory in overt and subclinical hypothyroidism. J Clin Endocrinol Metab 2009; 94: 3789-3797.

9.5.3.32. Heinrich TW, Grahm G. Hypothyroidism presenting as psychosis: myxedema madness revisited. Prim Care Companion J Clin Psychiatry 2003; 5: 260-266.

9.5.3.33. Appelhof BC, Brouwer JP, van Dijck R et al. Triiodothyronine addition to paroxetine in the treatmentof major depressive disorder. J Clin Endocrinol Metab 2004; 89: 6271-6276.

9.5.3.34. Cooper-Kazaz R, Apter JT, Cohen R et al. Combined treatment with sertraline and liothyronine in major depression. A randomized, double-blind, placebo-controlled trial. Arch Gen Psychiatry 2007; 64: 679-688.

9.5.3.35. Thomsen AF, Kvist TK, Andersen PK, Kessing LV. Increased risk of developing affective disorder in patients with hypothyroidism: a register-based study. Thyroid 2005; 15: 700-707.

9.5.3.36. Bauer M, Silverman DH, Schlagenhauf F et al. Brain glucose metabolism in hypothyroidism: a positron emission tomography study before and after thyroid hormone replacement therapy. J Clin Endocrinol Metab 2009; 94: 2922-2929.

9.5.3.37. Comer DM, McConnell EM. Hypothyroid-associated sensorineuronal deafness. Ir J Med Sci 2010; 178: 621-622.

9.5.3.38. He XS, Ma N, Pan ZL et al. Functional magnetic resource imaging assessment of altered brain function in hypothyroidism during working memory processing. Eur J Endocrinol 2011; 164: 951-959.

9.5.3.39. Cooke G, Mullally S, Correia N, O’Mara S, Gibney J. Hippocampal volume is decreased in adult-onset hypothyroidism. Thyroid 2013; [Epub ahead of print].

9.5.3.40. Montero-Pedrazuela A, Fernandez-Lamo J, Alieva M, Pereda-Perez I, Venero C, Guadano-Ferraz A. Adult-onset hypothyroidism enhances fear memory and upregulates mineralocorticoid and glucocorticoid receptors in the amygdala. PLoS One 2011; 6: e26582.

9.5.3.41. Koromilas C, Liapi C, Schulpis KH, Kalafatakis K, Zarros A, Tsakiris S. Structural and functional alterations in the hippocampus due to hypothyroidism. Metab Brain Dis 2010; 3: 339-354.

9.5.3.42. Pilhatsch M, Marxen M, Winter C, Smolka MN, Bauer M. Hypothyroidism and mood disorders: integrating novel insights from brain imaging techniques. Thyroid Res 2011; 4 Suppl 1: S3.

References for Section 9.5.4.

9.5.4.1. Graettinger JS, Muenster JJ, Checchia CS et al.: A correlation of clinical and hemodynamic studies in patients with hypothyroidism. J Clin Invest 1958; 37: 502.

9.5.4.2. DeGroot WJ, Leonard JJ: The thyroid state and the cardiovascular system. Mod Concepts Cardiovasc Dis 1969; 38: 23.

9.5.4.3. Buccino RA, Spann JF Jr, Sonnenblock EH, Braunwald E: Effect of thyroid state on myocardial contractility. Endocrinology 1968; 82: 191.

9.5.4.4. Levey GS, Skelton L, Epstein SE: Decreased myocardial adenyl cyclase activity in hypothyroidism. J Clin Invest 1969; 48: 2244.

9.5.4.5. Santos AD, Miller RP, Puthenpurakal KM, et al: Echocardiographic characterization of the reversible cardiomyopathy of hypothyroidism. Am J Med 1980; 68: 675.

9.5.4.6. Fuller H Jr, Spittell JA Jr, McConahey WM, Schirger A: Myxedema and hypertension. Postgrad Med 1966; 40: 425.

9.5.4.7. Polikar R, Burger AG, Scherrer U, Nicod P: The thyroid and heart. Circulation 1993; 87: 1435.

9.5.4.8. Zondek H: Das Myxödemherz. Muenchen Med Wochenschr 1918; 65: 1180.

9.5.4.9. Ladenson PW, Sherman SI, Baughman KL et al.: Reversible alterations in myocardial gene expression in a young man with dilated cardiomyopathy and hypothyroidism. Proc Natl Acad Sci USA 1992; 89: 5251.

9.5.4.10. Ladenson PW: Recognition and management of cardiovascular disease related to thyroid dysfunction. Am J Med 1990; 88: 638.

9.5.4.11. Gordon AH: Pericardial effusion in myxedema. Trans Assoc Am Physians 1935; 50: 272.

9.5.4.12. Smolar EN, Rubin JE, Avramides A, Carter AC: Cardiac tamponade in primary myxedema and review of the literature. Am J Med Sci 1976; 272: 345.

9.5.4.13. Davis PJ, Jacobsen S: Myxedema with cardiac tamponade and pericardial effusion of "gold paint" appearance. Arch Intern Med 1967; 120: 615.

9.5.4.14. Hardisty CA, Naik DR, Munro DS: Pericardial effusion in hypothyroidism. Clin Endocrinol 1980; 13: 349.

9.5.4.15. Zondek H: The electrocardiogram in myxoedema. Br Heart J 1964; 26: 227.

9.5.4.16. Lee JK, Lewis JA: Myxoedema with complete A-V block and Adams-Stokes disease abolished with thyroid medication. Br Heart J 1962; 24: 253.

9.5.4.17. Cohen RD, Lloyd-Thomas HG: Exercise electrocardiogram in myxoedema. Br Med J 1966; 2: 327.

9.5.4.18. Fredlund B-O, Olsson SB: Long QT interval and ventricular tachycardia of "torsade de pointe" type in hypothyroidism. Acta Med Scand 1983; 213: 231.

9.5.4.19. Nesher G, Zion MM: Recurrent ventricular tachycardia in hypothyroidism - report of a case and review of the literature. Cardiology 1988; 75: 301.

9.5.4.20. Crowley WF, Ridgway EC, Bough EW et al.: Noninvasive evaluation of cardiac function in hypothyroidism: Response to gradual thyroxine replacement. N Engl J Med 1977; 296: 1.

9.5.4.21. Rodbard D, Fujita T, Rodbard S: Estimation of thyroid function by timing the arterial sounds. JAMA 1967; 201: 884.

9.5.4.22. Blumgart HL, Freedberg AS, Kurland GS: Hypercholesterolemia, myxedema, and atherosclerosis. Am J Med 1953; 14: 665.

9.5.4.23. Tielens E, Pillay M, Storm C, Berghout A. Cardiac function at rest in hypothyroidism evaluated by equilibrium radionuclide angiography. Clin Endocrinol 1999; 50: 497-502.

9.5.4.24. Tielens, Pillay M, Storm C, Berghout A. Changes in cardiac function at rest before and after treatment in primary hypothyroidism. Am J Cardiol 2000; 85: 376-380.

9.5.4.25. Steinberg AD: Myxedema and coronary artery disease - a comparative autopsy study. Ann Intern Med 1968; 68: 338.

9.5.4.26. Becker C: Hypothyroidism and atherosclerotic heart disease: pathogenesis, medical management, and the role of coronary artery bypass surgery. Endocr Rev 1985; 6: 432

9.5.4.27. Keating FR, Parkin TW, Selby JB, Dickinson LS: Treatment of heart disease associated with myxedema. Prog Cardiovasc Dis 1961; 3: 364.

9.5.4.28. Levine HD: Compromise therapy in the patient with angina pectoris and hypothyroidism: a clinical assessment. Am J Med 1980; 69: 411.

9.5.4.29. Diekman MJM, Harms MPM, Endert E, Wieling W, Wiersinga WM. Endocrine factors related to changes in total peripheral vascular resistance after treatment of thyrotoxic and hypothyroid patients. Eur J Endocrinol 2001; 144: 339-346.

9.5.4.30. Obuobie K, Smith J, Evans LM, et al. Increased central arterial stiffness in hypothyroidism. J Clin Endocrinol Metab 2002; 87: 4662-4666.

9.5.4.31. Dernellis J, Panaretou M. Effects of thyroid replacement therapy on arterial blood pressure in patients with hypertension and hypothyroidism. Am Heart J 2002; 143: 718-724.

9.5.4.32. Cappola AR, Ladenson PW. Hypothyroidism and atherosclerosis. J Clin Endocrinol Metab 2003; 88: 2438-2444.

9.5.4.33. Diekman MJM, van der Put NM, Blom HJ, et al. Determinants of changes in plasma homocysteine in hyperthyroidism and hypothyroidism. Clin Endocrinol 2001; 54: 197-204.

9.5.4.34. Christ-Crain M, Meier C, Guglielmetti M, et al. Elevated C-reactive protein and homocysteine values: cardiovascular risk factors in hypothyroidism? A cross-sectional and a double-blind placebo-controlled trial. Atherosclerosis 2003; 166: 379-386.

9.5.4.35. Razvi S, Weaver JU, Vanderpump MP, Pearce SH. The incidence of ischemic heart disease and mortality in people with subclinical hypothyroidism: reanalysis of the Whickham Survey cohort. J Clin Endocrinol Metab 2010; 95: 1734-1740.

9.5.4.36. Hak AE, Pols HA, Visser TJ, et al. Subclinical hypothyroidism is an independent risk factor for atherosclerosis and myocardial infarction in elderly women: the Rotterdam study. Ann Int Med 2000; 132: 270-278.

9.5.4.37. Kilein I, Danzi S. Thyroid disease and the heart. Circulation 2007; 116: 1725-1735.9.5.4.38. Danzi S, Klein I. Thyroid hormone and the cardiovascular system. Med Clin North Am 2012; 96: 257-268.

9.5.4.39. Pakula D, Marek B, Kajdaniuk D et al. Plasma levels of NT-pro-brain natriuretic peptide in patients with overt and subclinical hyperthyroidism and hypothyroidism. Endokrynol Pol 2011; 62: 523-528.

9.5.4.40. Biondi B. Mechanisms in endocrinology: heart failure and thyroid dysfunction. Eur J Endocrinol 2012; 167: 609-618.

9.5.4.41. Butala A, Chaudhari S, Sacerdote A. Cardiac tamponade as a presenting manifestation of severe hypothyroidism. BMJ Case Rep 2013;. Doi: 10.1136/bcr-12-2011-5281.

9.5.4.42. Thvilum M, Brandt F, Almind D, Christensen K, Brix TH, Hegedus L. Type and extent of somatic morbidity before and after the diagnosis of hypothyroidism. A nationwide register study. PLoS One 2013; 8: e75789. Doi: 10.371/journal.pone.0075789.

References for Section 9.5.5.

9.5.5.1. Wilson WR, Bedell GN: The pulmonary abnormalities in myxedema. J Clin Invest 1960; 39: 42.

9.5.5.2. Zwillich CW, Pierson DJ, Hofeldt FD et al.: Ventilatory control in myxedema and hypothyroidism. N Engl J Med 1975; 292: 662.

9.5.5.3. Nordqvist P, Dhunér KG, Stenberg K, Örndahl G: Myxedema coma and CO2 retention. Acta Med Scand 1960; 166; 189.

9.5.5.4. Weg JG, Calverly JR, Johnson C: Hypothyroidism and alveolar hypoventilation. Arch Intern Med 1965; 115: 302.

9.5.5.5. Orr WC, Males JL, Imes NK: Myxedema and obstructive sleep apnea. Am J Med 1981; 70: 1061

9.5.5.6. Ladenson PW, Goldenheim PD, Ridgway EC: Prediction and reversal of blunted ventilatory responsiveness in patients with hypothyroidism. Am J Med 1988: 84: 877-883.

9.5.5.7. Pelttari L, Rauhala E, Polo O, et al.: Upper airway obstruction in hypothyroidism. J Intern Med 1994; 236: 177-181.

9.5.5.8. George JT, Thow JC, Rodger KA, Mannion R, Jayagopal V. Reversibility of fibrotic appearance of lungs with thyroxine replacement therapy in patients with severe hypothyroidism. Endocr Pract 2009; 15: 720-724.

9.5.5.9. Mickelson SA, Lian T, Rosenthal L. Thyroid testing and thyroid hormone replacement in patients with sleep disordered breathing. Ear, Nose & Throut Journal 1999; 78: 768-775.

9.5.5.10. Attal P, Chanson P. Endocrine aspects of obstructive sleep apnea. J Clin Endocrinol Metab 2010; 95: 483-495.

References for Section 9.5.6.

9.5.6.1. Kaminsky P, Robin-Lherbier B, Brunotte F, et al.: Energetic metabolism in hypothyroid skeletal muscle, as studied by phosphorus magnetic resonance spectroscopy. J Clin Endocrinol Metab 1992; 74: 124-129.

9.5.6.2. Monzani F, Caraccio N, Siciliano G, et al.: Clinical and biochemical features of muscle dysfunction in subclinical hypothyroidism. J Clin Endocrinol Metab 1997; 82: 3315-3318.

9.5.6.3. Khaleeli AA, Griffith DG, Edwards RHT: The clinical presentation of hypothyroid myopathy. Clinical Endocrinology 1983; 19: 365.

9.5.6.4. Hurwitz LJ, McCormick D, Allen IV: Reduced muscle a-glucosidase (acid-maltase) activity in hypothyroid myopathy. Lancet 1970; 1: 67.

9.5.6.5. Waldstein SS, Bronsky D, Shrifter HB, Oester YT: The electromyogram in myxedema. Arch Intern Med 1958; 101: 97.

9.5.6.6. Debré R, Sémélaigne G: Syndrome of diffuse muscular hypertrophy in infants causing athletic appearance: Its connection with congenital myxedema. Am J Dis Child 1935; 50: 1351.

9.5.6.7. Thomasen E: Myotonia, Thomsen's Disease, Paramyotonia, Dystrophia Myotonica, Aarhus, Denmark, Universitetsforlaget, 1948.

9.5.6.8. Hsu I-H, Thadhani RI, Daniels GH: Acute compartment syndrome in a hypothyroid patient. Thyroid 1995; 5: 305-308.

9.5.6.9. Lambert EH, Underdahl LO, Beckett S, Mederos LO: A study of the ankle jerk in myxedema. J Clin Endocrinol Metab 1951; 11: 1186.

9.5.6.10. Famulski KS, Pilarska M, Wrzosek A, Sarzala MG: ATPase activity and protein phosphorylation in rabbit fast skeletal muscle sarcolemma. Eur J Biochem 1988; 171: 363.

9.5.6.11. Simonides WS, van Hardeveld C, Larsen PR: Identification of sequences in the promoter of the fast isoform of sarcoplasmic reticulum Ca-ATPase (SERCA1) required for transcriptional activation by thyroid hormone. Thyroid 1992; 2: S-102.

9.5.6.12. Bland JH, Frymoyer JW: Rheumatoid syndrome of myxedema. N Engl J Med 1970; 282: 1171.

9.5.6.13. Frymoyer JW, Bland JH: Carpal-tunnel syndrome in patients with myxedematous arthropathy. J Bone Joint Surg 1973; 55A: 78.

9.5.6.14. Bonakdarpour A, Kirkpatrick JA, Renzi A, Kendall N: Skeletal changes in neonatal thyrotoxicosis. Radiology 1972; 102: 149.

9.5.6.15. Krane S, Bronwell GL, Stanbury JB, Corrigan H: The effect of thyroid disease on calcium metabolism in man. J Clin Invest 1956; 35: 874.

9.5.6.16. Bouillon R, De Moor P: Parathyroid function in patients with hyper- or hypothyroidism. J Clin Endocrinol Metab 1974; 38: 999.

9.5.6.17. Lever EG: Primary hyperparathyroidism masked by hypothyroidism. Am J Med 1983; 74: 144.

9.5.6.18. Lave CE, Bird ED, Thomas WC: Hypercalcemia in myxedema. J Clin Endocrinol Metab 1962; 22: 261.

9.5.6.19. Bouillon R, Muls E, De Moor P: Influence of thyroid function on the serum concentration of 1,25-dihydroxy vitamin D3. J Clin Endocrinol Metab 1980; 51: 793.

9.5.6.20. Mundy GR, Shapiro JL, Bandelin JG, et al.: Direct stimulation of bone resorption by thyroid hormones. J Clin Invest 1976; 58: 529.

9.5.6.21. Eriksen EF: Normal and pathological remodeling of human trabecular bone: Three dimensional reconstruction of the remodeling sequence in normals and in metabolic bone disease. Endocr Rev 1986; 7: 379.

9.5.6.22. Duyff RF, Bosch J vd, Laman DM, et al. Neuromuscular findings in thyroid dysfunction: a prospective clinical and electrodiagnostic study. J Neurol Neurosurg Psychiatry 2000; 68: 750-755.

9.5.6.23. Cakir M, Samanci N, Balci N, Balci MK. Musculoskeletal manifestations in patients with thyroid disease. Clin Endocrinol 2003; 59: 162-167.

9.5.6.24. Madariaga M. Polymyositis-like syndrome in hypothyroidism: review of cases reported over the past twenty-five years. Thyroid 2002; 12: 331-336.

9.5.6.25. Heemstra KA, Soeters MR, Fliers E et al. Type 2 iodothyronine deiodinase in skeletal muscle: effects of hypothyroidism and fasting. J Clin Endocrinol Metab 2009; 94: 2144-2150.

9.5.6.26. Vestergaard P, Mosekilde L. Fractures in patients with hyperthyroidism and hypothyroidism: a nationwide follow-up study in 16,249 patients. Thyroid 2002; 12: 411-419.

9.5.6.27. Vestergaard P, Rejnmark L, Mosekilde L. Influence of hyper- and hypothyroidism, and the effect of treatment with antithyroid drugs and levothyroxine on fracture risk. Calcif Tiss Int 2005; 77: 139-144.

References for Section 9.5.7.

9.5.7.1. Lerman J, Means JH: The gastric secretion in exophthalmic goitre and myxoedema. J Clin Invest 1932; 11: 167.

9.5.7.2. Donaich D, Roitt IM: An evaluation of gastric and thyroid auto-immunity in relation to hematologic disorders. Semin Hematol 1964; 1: 313.

9.5.7.3. Ebert EC. The thyroid and the gut. J Clin Gastroenterol 2010; 44: 402-406.

9.5.7.4. Lauritano EC, Bilotta AL, Gabrielli M et al. Association between hypothyroidism and small intestinal bacterial overgrowth. J Clin Endocrinol Metab 2007; 92: 4180-4184.

9.5.7.5. Rahman Q, Haboubi NY, Hudson PR, et al.: The effect of thyroxine on small intestinal motility in the elderly. Clin Endocrinol 1991; 35: 443-446

9.5.7.6. Hohl RD, Nixon RK: Myxedema ileus. Arch Intern Med 1965; 115: 145.

9.5.7.7. Nickerson JF, Hill SR, McNeil JH, Barker SB: Fatal myxedema with and without coma. Ann Intern Med 1960; 53: 475.

9.5.7.8. Tachman ML, Guthrie GP Jr: Hypothyroidism: diversity of presentation. Endocr Rev 1984; 5: 456.

9.5.7.9. Broitman SA, Bondy DC, Yachnin I et al.: Absorption and disabsorption of D-xylose in thyrotoxicosis and myxedema. N Engl J Med 1964; 270: 333.

9.5.7.10. Depczynski B, Ward R, Eisman J: The association between myxedematous ascites and extreme elevation of serum tumor markers. J Clin Endocrinol Metab 1996; 81: 4175.

9.5.7.11. Lorenzo y Losade H Jr, Staricoo EC, Cervino JM et al. Hypotonia of the galbladder of myxedematous origin. J Clin Endocrinol Metab 1957; 17: 133.

9.5.7.12. Cakir M, Kayacetin E, Toy H, Bozkurt S. Gallbladder motor function in patients with different thyroid hormone status. Exp Clin Endocrinol Diab 2009; 117: 395-399.

9.5.7.13. Fleisher G, McConahey W, Pankow M: Serum creatine kinase, lactic dehydrogenase, and glutamic-oxalacetic transaminase in thyroid diseases and pregnancy. Mayo Clinic Proc 1965; 40: 300.

9.5.7.14. Amino N, Kuro R, Yabu Y, et al.: Elevated levels of circulating carcinoembryonic antigens in hypothyroidism. J Clin Endocrinol Metab 1981; 52: 457.

9.5.7.15. Saha B, Maity C. Alterations of serum enzymes in primary hypothyroidism. Clin Chem Lab Med 2002; 40: 609-611.

9.5.7.16. Targher G, Montagnana M, Salvagno G et al. Association between serum TSH, free T4 and serum liver enzyme activities in a large cohort of unselected outpatients. Clin Endocrinol 2008; 68: 481-484.

References for Section 9.5.8.

9.5.8.1. Nedreb BG, Ericsson U-B, Nygård O, et al.: Plasma total homocysteine levels in hyperthyroid and hypothyroid patients. Metabolism 1998; 47: 89-93.

9.5.8.2. Yount E, Little JM: Renal clearance in patients with myxedema. J Clin Endocrinol Metab 1955; 15: 343.

9.5.8.3. Discala VA, Kinney MJ: Effects of myxedema on the renal diluting and concentrating mechanism. Am J Med 1971; 50: 325

9.5.8.4. Ford RV, Owens JC, Curd GW Jr et al.: Kidney function in various thyroid states. J Clin Endocrinol Metab 1961; 21: 548.

9.5.8.5. Moses AM, Gabrilove JL, Soffer LJ: Simplified water loading test in hypoadrenocorticism and hypothyroidism. J Clin Endocrinol Metab 1958; 18: 1413.

9.5.8.6. Crispell KR, Parson W, Sprinkle P: A cortisone-resistant abnormality in the diuretic response to ingested water in primary myxedema. J Clin Endocrinol Metab 1954; 14: 640.

9.5.8.7. Bleifer KH, Belsky JL, Saxon L, Papper S: The diuretic response to administered water in patients with primary myxedema. J Clin Endocrinol Metab 1960; 20: 409.

9.5.8.8. Davies CE, MacKinnon J, Platts MM: Renal circulation and cardiac output in "low-output" heart failure and in myxoedema. Br Med J 1952; 2: 595.

9.5.8.9. Pettinger WA, Talner L, Ferris TF: Inappropriate secretion of antidiuretic hormone due to myxedema. N Engl J Med 1965; 272: 361.

9.5.8.10. Showsky WR, Kikuchi TA: The role of vasopressin in the impaired water excretion of myxedema. Am J Med 1978; 64: 613.

9.5.8.11. Iwasaki Y, Oiso Y, Yamauchi K, et al.: Osmoregulation of plasma vasopressin in myxedema. J Clin Endocrinol Metab 1990; 70: 434.

9.5.8.12. Hanna FWF, Scanlon MF: Hyponatraemia, hypothyroidism and role of arginine-vasopressin. Lancet 1997; 350: 755-756.

9.5.8.13. Montenegro J, Gonzalez O, Saracho R, et al.: Changes in renal function in primary hypothyroidism. Am J Kidney Dis 1996; 27: 195-198.

9.5.8.14. Leeper RD, Benua RS, Brener JL, Rawson RW: Hyperuricemia in myxedema. J Clin Endocrinol Metab 1960; 20: 1457.

9.5.8.15. Jones JE, Desper PC, Shane SR, Flink EB: Magnesium metabolism in hyperthyroidism and hypothyroidism. J Clin Invest 1966; 45: 891.

9.5.8.16. Bernstein R, Midtb K, Urdal P, et al.: Serum N-terminal pro-atrial natriuretic factor 1-98 before and during thyroxine replacement therapy in severe hypothyroidism. Thyroid 1997; 7: 415-419.

9.5.8.17. Diekman MJM, Put NM vd, Blom HJ, Tijssen JGP, Wiersinga WM. Determinants of changes in plasma homocysteine in hyperthyroidism and hypothyroidism. Clin Endocrinol 2001; 54: 197-204.

9.5.8.18. Sahun M, Villabona C, Rosel P, et al. Hypothyroidism is associated with plasma hypo-osmolality and impaired water excretion that is vasopressin-independent. J Endocrinol 2001; 168:435-445.

9.5.8.19. Goede DL, Wiesli P, Brandle M et al. Effects of thyroxine replacement on serum creatinine and cystatin C in patients with primary and central hypothyroidism. Swiss Med Wkly 2009; 139: 339-344.

9.5.8.20. Schrier RW. Vasopressin and aquaporin 2 (AQP2) in clinical disorders of water homeostasis. Semin Nephrol 2008; 28: 289-296.

9.5.8.21. Chen Y-C, Cadnapaphomchai MA, Yang J et al. Nonosmotic release of vasopressin and renal aquaporins in impaired urinary dilution in hypothyroidism. Am J Physiol Renal Physiol 2005; 289: F672-F678.

9.5.8.22. Iglesias P, Diez JJ. Thyroid dysfunction and kidney disease. Eur J Endocrinol 2009; 160: 503-515.

9.5.8.23. Kimmel M, Braun N, Alscher MD. Influence of thyroid function on different kidney function tests. Kidney Blood Press Res 2012; 35: 9-17.

9.5.8.24. Ye Y, Gai X, Xie H, Jiao L, Zhang S. Impact of thyroid function on serum cystatin C and estimated glomerular filtration rate: a cross-sectional study. Endocr Pract 2013; 19: 397-403.

9.5.8.25. Tsuda A, Inaba M, Ichii M et al. Relationship between serum TSH levels and intrarenal hemodynamic parameters in euthyroid subjects. Eur J Endocrinol 2013; 169: 45-50.

9.5.8.26. Schwarz C, Leichtle AB, Arampatzis S et al. Thyroid function and serum electrolytes: does an association really exists? Swiss Med Wkly 2012; 142: w13669. Doi: 10.4414/smw.2012.13669.

9.5.8.27. Hammami MM, Almogbel F, Hammami S, Faifi J, Alqahtani A, Hashem W. Acute severe hypothyroidism is not associated with hyponatremia even with increased water intake: a prospective study in thyroid cancer patients. BMC Endocr Disord 2013; 13:27 [Epub ahead of print].

9.5.8.28. Peri A. Clinical review: the use of vaptans in clinical endocrinology. J Clin Endocrinol Metab 2013; 98: 1321-1332.

9.5.8.29. Karmisholt J, Andersen S, Laurberg P. Weight loss after therapy of hypothyroidism is mainly caused by excretion of excess body water associated with myxoedem. J Clin Endocrinol Metab 2011; 96: E99-E103.

References for Section 9.5.9.

9.5.9.1. Krassas GE, Poppe K, Glinoer D. Thyroid function and human reproductive health. Endocrine Rev 2010; 31: doi:10.1210/er.2009-0041.

9.5.9.2. Tagawa N, Takano T, Fukata S et al. Serum concentrations of androstenediol and androstenediol sulfate in patients with hyperthyroidism and hypothyroidism. Endocr J 2001; 48: 345-354.

9.5.9.3. Gordon GG, Southren AL: Thyroid-hormone effects on steroid-hormone metabolism. Bull NY Acad Med 1977; 53: 241.

9.5.9.4. Copinschi G, Leclercq R, Bruno OD, Cornil A: Effects of altered thyroid function upon cortisol secretion in man. Horm Metab Res 1971; 3: 437.

9.5.9.5. Gordon GG, Southern AL, Tochimoto S, et al.: Effect of hyperthyroidism and hypothyroidism on testosterone and androstenedione in man. J Clin Endocrinol Metab 1969; 29: 164.

9.5.9.6. De La Balze FA, Arrillaga F, Mancini RE, et al.: Male hypogonadism in hypothyroidism: a study of six cases. J Clin Endocrinol Metab 1962; 22: 212.

9.5.9.7. Hopwood NJ, Lockhart LH, Bryan GT: Acquired hypothyroidism with muscular hypertrophy and precocious testicular enlargement. J Pediatr 1974; 85: 233

9.5.9.8. Bruder JM, Samuels MH, Bremner WJ, et al.: Hypothyroidism-induced macroorchidism: use of a gonadotropin-releasing hormone agonist to understand its mechanism and augment adult stature. J Clin Endocrinol Metab 1995; 80: 11-16.

9.5.9.9. Anasti JN, Flack MR, Froehlich J, et al.: A potential novel mechanism for precocious puberty in juvenile hypothyroidism. J Clin Endocrinol Metab 1995; 80: 276-279.

9.5.9.10. Carani C, Isidori AM, Granata A et al. Multicenter study on the prevalence of sexual symptoms in male hypo- and hyperthyroid patients. J Clin Endocrinol Metab 2005; 90: 6472-6479.

9.5.9.11. Krassas GE, Tziomalos K, Papadopoulou F, Pontikides N, Perros P. Erectile dysfunction in patients with hyper- and hypothyroidism: how common and should we treat? J Clin Endocrinol Metab 2008; 93: 1815-1819.

9.5.9.12. Corrales Hernandez JJ, Miralles Garcia JM, Garcia Diez LC. Primary hypothyroidism and human spermatogenesis. Arch Androl 1990; 25: 21-27.

9.5.9.13. Jaya Kumar B, Khurana ML, Ammini AC, Kamarkar MG, Ahuja MM. Reproductive endocrine functions in men with primary hypothyroidism: effect of thyroxine replacement. Horm Res 1990; 34: 215-218.

9.5.9.14. Krassas GE, Papadopoulou F, Tziomalos K, Zeginiadou T, Pontikides N. Hypothyroidism has an adverse effect on human spermatogenesis: a prospective, controlled study. Thyroid 2008; 18: 1255-1259.

9.5.9.15. Longcope C, Abend S, Braverman LE, Emerson CH. Androstenedione and estrone dynamics in hypothyroid women. J Clin Endocrinol Metab 1990; 70: 903-907.

9.5.9.16. Costin G, Kershnar AK, Kogut MD, Turkington RW. Prolactin activity in juvenile hypothyroidism and precocious puberty. Pediatrics 1972; 50: 881.

9.5.9.17. Goldsmith RE, Sturgis SH, Lerman J, Stanbury JB: The menstrual pattern in thyroid disease. J Clin Endocrinol Metab 1952; 12: 846.

9.5.9.18. Samuels MH, Veldhuis JD, Henry P, Ridgway EC: Pathophysiology of pulsatile and copulsatile release of thyroid-stimulating hormone, luteinizing hormone, follicle-stimulating hormone, and alpha-subunit. J Clin Endocrinol Metab 1990; 71: 425-432

9.5.9.19. Krassas GE, Pontikides N, Kaltsas Th et al. Disturbances of menstruation in hypothyroidism. Clin Endocrinol 1999; 50: 655-659.

9.5.9.20. Benson RC, Dailey ME. The menstrual pattern in hyperthyroidism and subsequent posttherapy hypothyroidism. Surg Gynecol Obstet 1955; 100: 19-26.

9.5.9.21. Goldsmith RE, Sturgis SH, Lerman J, Stanbury JB. The menstrual pattern in thyroid disease. J Clin Endocrinol Metab 1952; 12: 846-855.

9.5.9.22. Scott Jr JC, Mussey E. Menstrual patterns in myxedema. Am J Obstet Gynecol 1964; 90: 161-165.

9.5.9.23. Joshi JV, Bhandarkar SD, Chadha M, Balaiah D, Shah R. Menstrual irregularities and lactation failure may precede thyroid dysfunction or goitre. J Postgrad Med 1993; 39: 137-141.

9.5.9.24. Ross F, Nusynowitz ML. A syndrome of primary hypothyroidism, amenorrhea and galactorrhea. J Clin Endocrinol Metab 1968; 28: 591.

9.5.9.25. Davis LE, Leveno KJ, Cunningham FG: Hypothyroidism complicating pregnancy. Obstet Gynecol 1988; 72: 108.

9.5.9.26. Potter JD: Hypothyroidism and reproductive failure. Surg Gynecol Obstet 1980; 150: 251.

9.5.9.27. Montoro M, Collea JV, Frasier SD, Mestman JH: Successful outcome of pregnancy in women with hypothyroidism. Ann Intern Med 1981; 94: 31.

9.5.9.28. Lincoln SR, Ke RW, Kutteh WH. Screening for hypothyroidism in infertile women. J Reprod Med 1999; 44: 455-457.

9.5.2.29. Cramer DW, Sluss PM, Powers RD et al. Serum prolactin and TSH in an in vitro fertilization population: is there a link between fertilization and thyroid function? J Assist Reprod Genet 2003; 20: 210-215.

9.5.9.30. Abalovich M, Gutierrez S, Alcaraz G, Maccallini G, Garcia A, Levalle O. Overt and subclinical hypothyroidism complicating pregnancy. Thyroid 2002; 12: 63-68.

9.5.9.31. Liu H, Momotani N, Noh JY et al.: Maternal hypothyroidism during early pregnancy and intellectual development of progeny. Arch Intern Med 1994; 154: 785.

9.5.9.32. Man EB, Brown JF, Serunian SA: Maternal hypothyroxinemia: Psychoneurological deficits of progeny. Ann Clin Lab Ser 1991; 21: 227.

9.5.9.33. Smallridge RC, Ladenson PW. Hypothyroidism in pregnancy: consequences to neonatal health. J Clin Endocrinol Metab 2001; 86: 2349-2353.

9.5.9.34. Haddow JE, Palomaki GE, Allan WC, et al. Maternal thyroid deficiency during pregnancy and subsequent neuropsychological development of the child. N Engl. J Med 1999; 341: 549-555.

9.5.9.35. Abalovich M, Amino N, Barbour LA et al. Management of thyroid dysfunction during pregnancy and postpartum. An Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab 2007; 92 (Supplement): S1-S47.

9.5.9.36. Krassas GE, Poppe K, Glinoer D. Thyroid function and human reproductive health. Endocr Rev 2010; 31: 702-755.

9.5.9.37. Stagnaro-Green A, Abalovich M, Alexander E et al. Guidelines of the American Thyroid Association for the diagnosis and management of thyroid diseases during pregnancy and postpartum. Thyroid 2011; 21: 1081-1125.

9.5.9.38. De Groot L, Abalovich M, Alexander EK et al. Management of thyroid dysfunction during pregnancy and postpartum: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab 2012; 97: 2543-2565.

References for Section-9.5.10.

9.5.10.1. Yamada T, Tsukui T, Ikejiri K, et al.: Volume of sella turcica in normal subjects and in patients with primary hypothyroidism and hyperthyroidism. J Clin Endocrinol Metab 1976; 42: 817.

9.5.10.2. Yamamoto K, Saito K, Takai T, et al.: Visual field defects and pituitary enlargement in primary hypothyroidism. J Clin Endocrinol Metab 1983; 57: 283.

9.5.10.3. Vagenakis AG, Dole K, Braverman LE: Pituitary enlargement, pituitary failure, and primary hypothyroidism. Ann Intern Med 1976; 85: 195.

9.5.10.4. Sarlis NJ, Brucker-Davis F, Doppman JL, Skarulis MC: MRI-demonstrable regression of a pituitary mass in a case of primary hypothyroidism after a week of acute thyroid hormone therapy. J Clin Endocrinol Metab 1997; 82: 808-811.

9.5.10.5. Honbo KS, Van Herle AJ, Kellett KA: Serum prolactin levels in untreated primary hypothyroidism. Am J Med 1978; 64: 782.

9.5.10.6. Onishi T, Miyai K, Aono T, et al.: Primary hypothyroidism and galactorrhea. Am J Med 1977; 63: 373.

9.5.10.7. Brauman H, Corvilain J: Growth hormone response to hypoglycemia in myxedema. J Clin Endocrinol Metab 1968; 28: 301.

9.5.10.8. Valcavi R, Valente F, Dieguez C, et al.: Evidence against depletion of the growth hormone (GH)-releasable pool in human primary hypothyroidism: studies with GH-releasing hormone, pyridostigmine, and arginine. J Clin Endocrinol Metab 1993; 77: 616-620.

9.5.10.9. Miell JP, Taylor AM, Zini M, et al.: Effects of hypothyroidism and hyperthyroidism on insulin-like growth factors (IGFs) and growth hormone- and IGF-binding proteins. J Clin Endocrinol Metab 1993; 76: 950.

9.5.10.10. Miell JP, Zini M, Quin JD, et al.: Reversible effects of cessation and recommencement of thyroxine treatment on insulin-like growth factors (IGFs) and IGF-binding proteins in patients with toal thyroidectomy. J Clin Endocrinol Metab 1994; 79: 1507-1512.

9.5.10.11. Gordon GG, Southren AL: Thyroid-hormone effects on steroid-hormone metabolism. Bull NY Acad Med 1977; 53: 241.

9.5.10.12. Brown H, Englert E, Wallach S: Metabolism of free and conjugated 17-hydroxycorticosteroids in subjects with thyroid disease. J Clin Endocrinol Metab 1958; 18: 167.

9.5.10.13. Copinschi G, Leclercq R, Bruno OD, Cornil A: Effects of altered thyroid function upon cortisol secretion in man. Horm Metab Res 1971; 3: 437.

9.5.10.14. Hellman L, Bradlow HL, Zumoff B, et al.: Thyroid-androgen interrelations and the hypocholesteremic effect of androsterone. J Clin Endocrinol Metab 1959; 19: 939.

9.5.10.15. Ogihara T, Yamamoto T, Miyai K, Kumahara Y: Plasma renin activity and aldosterone concentration of patients with hyperthyroidism and hypothyroidism. Endocrinol Jpn.

9.5.10.16. Saruta T, Kitajima W, Hayashi M, et al.: Renin and aldosterone in hypothyroidism: relation to excretion of sodium and potassium. Clin Endocrinol 1980; 12: 483.

9.5.10.17. Felber JP, Reddy WJ, Selenkow HA, Thorn GW: Adrenocortical response to the 48-hour ACTH test in myxedema and hyperthyroidism. J Clin Endocrinol Metab 1959; 19: 895.

9.5.10.18. Liddle GW, Estep HL, Kendall JW Jr, et al.: Clinical application of a new test of pituitary reserve. J Clin Endocrinol Metab 1959; 19: 875.

9.5.10.19. Gold EM, Kent JR, Forsham PH: Clinical use of a new diagnostic agent, methopyrapone (SU-4885), in pituitary and adrenocortical disorders. Ann Intern Med 1961; 54: 175.

9.5.10.20. Lessof MH, Maisey MN, Lyne C, Sturge RA: Effect of thyroid failure on the pituitary-adrenal axis. Lancet 1969; I: 642.

9.5.10.21. Ridgway EC, McCammon JA, Benotti J, Maloof F: Acute metabolic responses in myxedema to large doses of intravenous L-thyroxine. Ann Intern Med 1972; 77: 549.

9.5.10.22. Bigos ST, Ridgway EC, Kourides JA, Maloof F: Spectrum of pituitary alterations with mild and severe thyroid impairment. J Clin Endocrinol Metab 1978; 46: 317.

9.5.10.23. Kamilaris TC, DeBold CR, Pavlou SN, et al.: Effect of altered thyroid hormone levels on hypothalamic-pituitary-adrenal function. J Clin Endocrinol Metab 1987; 65: 994.

9.5.10.24. Clausen N, Lins PE, Adamson U, et al.: Couterregulation of insulin-induced hypoglycemia in primary hypothyroidism. Acta Endocrinol 1986; 111: 516.

9.5.10.25. Carpenter CCJ, Solomon N, Silverberg SG, et al.: Schmidt's syndrome (thyroid and adrenal insufficiency): a review of the literature and a report of fifteen new cases including ten instances of coexistent diabtes mellitus. Medicine 1964; 43: 153.

9.5.10.26. Salvi M, Fukazawa H, Bernard N, et al.: Role of autoantibodies in the pathogenesis and association of endocrine autoimmune disorders. Endocr Rev 1988; 9: 450.

9.5.10.27. Lack EE: Lymphoids "hypophysitis" with end organ insufficiency. Arch Pathol 1975; 99: 215.

9.5.10.28. Gharib H, Gastineau CF, Hodgson SF, et al.: Reversible hypothyroidism in Addison's disease. Lancet 1972; 2:734.

9.5.10.29. Christiansen CJ: Increased levels of plasma noradrenalin in hypothyroidism. J Clin Endocrinol Metab 1972; 35: 359.

9.5.10.30. Stoffer SS, Jiang N-S, Gorman CA, Pikler GM: Plasma catecholamines in hypothyroidism and hyperthyroidism. J Clin Endocrinol Metab 1973; 36: 587.

9.5.10.31. Heemstra KA, Burggraaf J, van der Klaauw AA, Romijn JA, Smit JW, Corssmit EP. Short-term overt hypothyroidism induces sympathovagal imbalance in thyroidectomized differentiated thyroid carcinoma patients. Clin Endocrinol 2010; 72: 417-421.

9.5.10.32. Raab W: Epinephrine tolerance of the heat altered by thyroxine and thiouracil. J Pharmacol Exp Ther 1944; 82: 330.

9.5.10.33. Raab W: Diminution of epinephrine sensitivity of the normal human heart through thiouracil. J Lab Clin Med 1945; 30: 774.

9.5.10.34. Schneckloth RE, Kurland GS, Freedberg AS: Effect of variation in thyroid function on the pressor response to norepinephrine in man. Metabolism 1953; 2: 546.

9.5.10.35. Polikar R, Kennedy B, Maisel A, et al.: Decreased adrenergic sensitivity in patients with hypothyroidism. J Am Coll Cardiol 1990; 15: 94-98.

9.5.10.36. Iglesias P, Bayon C, Mendez J, et al. Serum insulin-like growth factor type 1, insulin-like growth factor-binding protein-1, and insulin-like growth factor-binding protein-3 concentrations in patients with thyroid dysfunction. Thyroid 2001; 11: 1043-1048.

9.5.10.37. Schmid C, Zwimpfer C, Brandle M, Krayenbuhl PA, Zapf J, Wiesli P. Effect of thyroxine replacement on serum IGF-1, IGFBP-3, and the acid-labile subunit in patients with hypothyroidism and hypopituitarism. Clin Endocrinol 2006; 65: 706-711.

9.5.10.38. Eskes SA, Endert E, Fliers E, Wiersinga WM. Prevalence of growth hormone deficiency in Hashimoto’s thyroiditis. J Clin Endocrinol Metab 2010; 95: 2266-2270.

9.5.10.39. Raber W, Gesol A, Nowotny P, Vierhapper H. Hyperprolactinaemia in hypothyroidism: clinical significance and impact of TSH normalization. Clin Endocrinol 2003; 58: 185-191.

9.5.10.40. Silva JE, Bianco SD. Thyroid-adrenergic interactions: physiological and clinical implications. Thyroid 2008; 18: 157-165.

References for Section 9.5.11.

9.5.11.1. Watanakunakorn C, Hodges RE, Evans TC: Myxedema: a study of 400 cases. Arch Intern Med 1965; 116: 183.

9.5.11.2. Stern B, Altshule MD: Hematological studies in hypothyroidism following total thyroidectomy. J Clin Invest 1936; 15: 633.

9.5.11.3. Bomford R: Anaemia in myxoedema: and the role of the thyroid gland in erythropoiesis. Q J Med 1938; 7: 495.

9.5.11.4. Shalet M, Coe D, Reisman KR: Mechanism of erythropoietic action of thyroid hormone. Proc Soc Exp Biol Med 1966; 123: 443.

9.5.11.5. Tudhope GR, Wilson GM: Anemia in hypothyroidism. Q J Med 1960; 29: 513.

9.5.11.6. Muldowney FP, Crooke J, Wayne EJ: The total red cell mass in thyrotoxicosis and myxoedema. Clin Sci 1957; 16: 309.

9.5.11.7. Leithold SL, David D, Best WR: Hypothyroidism with anemia demonstrating abnormal vitamin B12 absorption. Am Jmed 1958; 24: 535.

9.5.11.8. Tudhope GR, Wilson GM: Deficiency of vitamin B12 in hypothyroidism. Lancet 1962; 1: 703.

9.5.11.9. Hines JD, Halstead CH, Criggs RC, Harris JW: Megaloblastic anemia secondary to folate deficiency associated with hypothyroidism. Ann Intern Med 1968; 68: 792.

9.5.11.10. Wardrop C, Hutchinson HE: Red cell shape in hypothyroidism. Lancet 1969; 1: 1243.

9.5.11.11. Lillington GA, Gastineau CF, Underdahl LO: The sedimentation rate in primary myxedema. Proc Staff Meetings Mayo Clinic 1959; 34: 605.

9.5.11.12. Squizzato A, Roumadi E, Buller HR, Gerdes VEA. Thyroid dysfunction and effects on coagulation and fibrinolysis: a systematic review. J Clin Endocrinol Metab 2007; 92: 2415-2420.

9.5.11.13. Ford HC, Carter JM: Haemostasis in hypothyroidism. Postgrad Med J 1990; 66: 280-284.

9.5.11.14. Hofbauer LC, Heufelder A: Coagulation disorders in thyroid disease. Europ J Endocrinol 1997; 136: 1-7.

9.5.11.15. Verkleij CJ, Stuijver DJ, van Zaane B et al. Thrombin-activatable fibrinolysis inhibitor in hypothyroidism and hyperthyroxinemia. Thromb Haemost 2013; 109: 214-220.

9.5.11.16. Manfredi E, van Zaane B, Gerdes VEA, Brandjes DPM, Squizzato A. Hypothyroidism and acquired von Willebrand’s syndrome: a systematic review. Haemophilia 2008; 14: 423-433.

9.5.11.17. Erfurth EM, Ericsson U-BC, Egervalh K, Lethagen SR: Effect of acute desmopressin and of long-term thyroxine replacement on haemostasis in hypothyroidism. Clin Endocrinol 1995; 42: 373-378.

9.5.11.18 Showsky WR, Kikuchi TA: The role of vasopressin in the impaired water excretion of myxedema. Am J Med 1978; 64: 613.

9.5.11.19. Chadarevian R, Bruckert E, Leenhardt L, Giral Ph, Ankri A, Turpin G. Components of the fibrinolytic system are differently altered in moderate and severe hypothyroidism. J Clin Endocrinol Metab 2001; 86: 732-737.

9.5.11.20. De Vito P, Incerpi S, Pedersen JZ, Luly P, Davis FB, Davis PJ. Thyroid hormones as modulators of immune activities at the cellular level. Thyroid 2011; 21: 879-890.

9.5.11.21. Kim JH, Park JH, Kim SY, Bae HY. The mean platelet volume is positively correlated with serum thyrotropin concentrations in a population of healthy subjects and subjects with unsuspected subclinical hypothyroidism. Thyroid 2013; 23: 31-37.

9.5.11.22. Erikci AA, Karagoz B, Ozturk A et al. The effect of subclinical hypothyroidism on platelet parameters. Hematology 2009; 14: 115-117.

9.5.11.23. Torun AN, Uzum AK, Aksoy N. Overt and mild subclinical hypothyroidism do not influence mean platelet volume in premenopausal women having low cardiac risk. Clin Appl Thromb Hemost 2012; 18: 312-315.

9.5.11.24. Stuijver DJ, Plantanida E, van Zaane B et al. Acquired von Willebrand syndrome in patients with overt hypothyroidism: a prospective cohort study. Haemophilia 2013 [Epub ahead of print]. Doi: 10.1111/hae.12275.

9.5.11.25. Yango J, Alexopoulou O, Eeckhoudt S, Hermans C, Daumerie C. Evaluation of the respective influence of thyroid hormones and TSH on blood coagulation parameters after total thyroidectomy. Eur J Endocrinol 2011; 164: 599-603.

9.5.11.26. Vescovi PP, Favorolo EJ, Lippi G et al. The spectrum of coagulation abnormalities in thyroid disorders. Semin Thromb Hemost 2011; 37: 7-10.

9.5.11.27. Marongiu F, Barcellona D, Mameli A, Mariotti S. Thyroid disorders and hypercoagulability. Semin Thromb Hemost 2011; 37: 11-16.

References for Section 9.6.

9.6.1. Ord WM: In Allbutt TC (ed): System of Medicine. New York. The Macmillan Co, 1897.

References for Sections 9.7.1-9.7.3.

9.7.1. Tachman ML, Guthrie GP: Hypothyroidism: diversity of presentation. Endocr Rev 1984; 5: 456-465.

9.7.2. Billewicz WL, Chapman RS, Crooks J, et al.: Statistical methods applied to the diagnosis of hypothyroidism. Q J Med (New Series) 1969; 38: 255-266.

9.7.3. Seshadri MS, Samuel BU, Kanagasabapathy AS, Cherian AM: Clinical scoring system for hypothyroidism: is it useful? J Gen Intern Med 1989; 4: 490-492.

9.7.4. Zulewski H, Müller B, Exer P, et al.: Estimation of tissue hypothyroidism by a new clinical score: evaluation of patients with various grades of hypothyroidism and controls. J Clin Endcrinol Metab 1997; 82: 771-776.

9.7.5. Doucet J, Trivalle Ch, Chassagne Ph, et al.: Does age play a role in clinical presentation of hypothyroidism? J Am Geriatr Soc 1994; 42: 984-986.

9.7.6. Müller B, Zulewski H, Huber P, et al.: Impaired action of thyroid hormone associated with smoking in women with hypothyroidism. New Engl J Med 1995; 333: 964-969.

9.7.7. Nikolai TF: Recovery of thyroid function in primary hypothyroidism. Am J Med Sci 1989; 297: 18-21.

9.7.8. Comtois R, Faucher L, Laflèche L: Outcome of hypothyroidism caused by Hashimoto's thyroiditis. Arch Int Med 1995; 155: 1404-1408.

9.7.9. Okamura K, Sato K, Ikenoue H, et al.: Reevaluation of thyroidal radioactive iodine uptake test, with special reference to reversible primary hypothyroidism with elevated thyroid radioiodine uptake. J Clin Endocrinol Metab 1988; 67: 720-726.

9.7.10. Takasu N, Komiya I, Asawa T, et al.: Test for recovery from hypothyroidism during thyroxine therapy in Hashimoto's thyroiditis. Lancet 1990; 336: 1084-1086.

9.7.11. Wiersinga WM, Touber JL, Trip MD, van Royen EA: Uninhibited thyroidal uptake of radioiodine despite iodine excess in amiodarone-induced hypothyroidism. J Clin Endocrinol Metab 1986; 63: 485-491.

9.7.12. Garber JR, Cobin RH, Gharib H, et al. Clinical practice guidelines for hypothyroidism in adults: cosponsored by the American Association of Clinical Endocrinologists and the American Thyroid Association. Thyroid 2012; 22: 1200-1235.

9.7.13. Bochukova E, Schoenmakers N, Agostini M, et al: A mutation in the thyroid hormone receptor alpha gene. . N Engl J Med 2012; 366: 243-249.

9.7.14. Van Mullem A, Van Heerebeek R, Chrysis D et al: Clinical phenotype and mutant TRα1. N Engl J Med 2012; 366: 1451-1453.

9.7.15. Pedersen OM, Aardal NP, Larssen TB, Varhaug JE, Myking O, Vik-Mo H: The value of ultrasonography in predicting autoimmune thyroid disease. Thyroid 2000; 10: 251-259.

9.7.16. Boucai L, Hollowell JG, Surks MI. An approach for development of age-, gender-, and ethnicity=specific thyrotropin reference limits. Thyroid 2011; 21: 5-11.

References for Section 9.8.1.

9.8.1.1. Hays MT: Localization of human thyroxine absorption. Thyroid 1991; 3: 241.

9.8.1.2. Hays MT: Thyroid hormone and the gut. Endocr Res 1988; 14: 203.

9.8.1.3. Fish LH, Schwartz HL, Cavanaugh J, et al.: Replacement dose, metabolism, and bioavailability of levothyroxine in the treatment of hypothyroidism. Role of triiodothyronine in pituitary feedback in humans. New Engl J Med 1987; 316: 764-770.

9.8.1.4. Bach-Huynh TG, Nayak B, Loh J et al.: Timing of levothyroxine administration affects serum thyrotropin concentration. J Clin Endocrinol Metab 2009; 94: 3905-3912.

9.8.1.5. Wennlund A: Variation in serum levels of T3, T4, FT4 and TSH during thyroxine replacement therapy. Acta Endocrinol 1986; 113: 47.

9.8.1.6. Browning MCK, Bennet WM, Kirkaldy AJ, Jung RT: Intra-individual variation of thyroxine, triiodothyronine, and thyrotropin in treated hypothyroid patients: Implications for monitoring replacement therapy. Clin Chem 1988; 34: 696.

9.8.1.7. Daniels GH. Response to “How do you approach the problem of TSH elevation in a patient on high-dose thyroid hormone replacement?” Clin Endocrinol 2009; 71: 603.

9.8.1.8. Dong BJ, Hauck WW, Gambertoglio JG et al. Bioequivalence of generic and brand-name levothyroxine products in the treatment of hypothyroidism. JAMA 1997; 227: 1205-1213.

9.8.1.9. Olveira G, Almaraz MC, Soriguer F et al. Altered bioavailability due to changes in the formulation of a commercial preparation of levothyroxine in patients with differentiated carcinoma. Clin Endocrinol 1997; 46: 707-711.

9.8.1.10. Carr D, McLeod DT, Parry G, Thornes HM. Fine adjustment of thyroxine replacement dosage: Comparison of the thyrotropin releasing hormone test using a sensitive thyrotropin assay with measurement of free thyroid hormones and cli nical assessment. Clin Endocrinol 1988; 28: 325-333.

9.8.1.11. Canaris GJ, Manowitz NR, Mayor GM, Ridgway EC. The Colorado thyroid disease prevalence study. Arch Int Med 2000; 160: 526-534.

9.8.1.12. Hollowell JG, Staehling NW, Flanders WD et al. Serum TSH, T4, and thyroid antibodies in the United States Population (1988 to 1994): National Hea;th and Nutritional Examination Survey (NHANES III). J Clin Endocrinol Metab 2002; 87: 489-499.

9.8.1.13. Food and Drug Administration, Center for Drug Evaluation and Research 2000 Guidance for Industry: Levothyroxine sodium tablets – In vivo pharmacokinetic and bioavailability studies and in vitro dissolution testing. www.fda.gov/cder/guidance/364fnl.htm

9.8.1.14. American Thyroid Association, Endocrine Society, American Association of Clinical Endocrinologists. Joint statement on the U.S. Food and Drug Administration’s decision regarding bioequivalence of levothyroxine sodium. Thyroid 2004; 14: 486.

9.8.1.15. Blakesley V, Awni W, Locke C, Ludden T, Granneman GR, Braverman LE. Are bioequivalence studies of levothyroxine sodium formulations in euthyroid volunteers reliable? Thyroid 2004; 14: 191- 200.

9.8.1.16. Stockigt J. Testing the bioavailability of oral L-thyroxine by studying its absorption: smoke or mirrors? Thyroid 2004; 14: 167-168.

9.8.1.17. Di Girolamo G, Keller GA, de Los Santos AR, Schere D, Gonzalez CD. Bioequivalence of two levothyroxine tablet formulations without and with mathematical adjustment for basal thyroxine levels in healthy Argentinian volunteers: a single-dose, randomized, open-label, crossover study. Clin Ther 2008; 30: 2015-2023.

9.8.1.18. Eisenberg M, DiStefano JJ. TSH-based protocol, tablet instability, and absorption effects on L-T4 bioequivalence. Thyroid 2009; 19: 103-110.

9.8.1.19 .Pabla D, Akhlaghi F, Zia H. A comparative pH-dissolution profile study of selected commercial levothyroxine products using inductively coupled mass spectrometry. Eur J Pharm Biopharm 2009; 72: 105-110.

9.8.1.20. Hennessey JV, Malabanan AO, Haugen BR, Levy EG. Adverse event reporting in patients treated with levothyroxine: results of the pharmacovigilance task force survey of the American Thyroid Association, American Association of Clinical Endocrinologists and The Endocrine Society. Endocr Pract 2010; 11: 1-41.

9.8.1.21. LeBoff MS, Kaplan MM, Silva JE, Larsen PR: Bioavailability of thyroid hormones from oral replacement preparations. Metabolism 1982; 31: 900.

9.8.1.22. Jackson S, William E, Cobb E: Why does anyone still use desiccated thyroid USP? Am J Med 1978; 64: 284-288.

9.8.1.23. Blumberg KR, Mayer WJ, Parikh DK, Schnell LA: Liothyronine and levothyroxine in Armour thyroid. J Pharm Sci 1993; 76: 346.

9.8.1.24. Rees-Jones RW, Rolla AR, Larsen PR: Hormone content of thyroid replacement preparations. JAMA 1980; 243: 549.

9.8.1.25. Rees-Jones RW, Larsen PR: Triiodothyronine and thyroxine content of desiccated thyroid tablets. Metabolism 1977; 26: 1213.

9.8.1.26. Escobar-Morreale HF, Escobar del Ray F, Obregon MJ, Morreale de Escobar G: Only the combined treatment with thyroxine and triiodothyronine ensures euthyroidism in all tissues of the thyroidectomized rat. Endocrinology 1996; 137: 2490-2502.

9.8.1.27. Garber JR, Cobin RH, Gharib H et al. Clinical practice guidelines for hypothyroidism in adults: cosponsored by th American Association of Clinical endocrinologists and the American Thyroid association. Thyroid 2012; 22: 1200-1235.

9.8.1.28. Celi FS, Zemskova M, Linderman JD et al. The pharmacodynamic equivalence of levothyroxine and liothyronine: a randomized, double blind, cross-over study in thyroidectomized patients. Clin Endocrinol 2010; 72: 709-713.

9.8.1.29. Hoang TD, Olsen CH, Mai VQ, Clyde PW, Shakir MKM. Desiccated thyroid extract compared with levothyroxine in the treatment of hypothyroidism: a randomized, double-blind, crossover study. J Clin Endocrinol Metab 2013; 98: 2013.

9.8.1.30. Grozinsky-Glasberg S, Fraser A, Nahshoni E, Weizman A, Leibovici L. Thyroxine-triiodothyronine combination therapy versus thyroxine monotherapy for clinical hypothyroidism: meta-analysis of randomized controlled trials. J Clin Endocrinol Metab 2006; 91: 2592-2599.

9.8.1.31. Nygaard B, Jensen EW, Kvetny J, Jarlov A, Faber J. Effect of combination therapy with thyroxine (T4) and 3,5,3’-triiodothyronine versus T4 monotherapy in patients with hypothyroidism, a double-blind, randomized cross-over study. Eur J Endocrinol 2009; 161: 895-902.

9.8.1.32. Wiersinga WM, Duntas L, Fadeyev V, Nygaard B, Vanderpump MP. 2012 ETA guidelines: the use of L-T4 + L-T3 in the treatment of hypothyroidism. Eur Thyroid J 2013; 1: 55-71.

9.8.1.33. Bunevicius R, Kazanavicius G, Zalinkevicius R, Prange AJ: Effects of thyroxine as compared with thyroxine plus triiodothyronine in patients with hypothyroidism. New Engl J Med 1999; 340: 424-429.

9.8.1.34. Toft AD: Thyroid hormone replacement - one hormone or two? New Engl J Med 1999; 340: 469-470.

9.8.1.35. Bunevicius R, Prange AJ. Mental improvement after replacement therapy with thyroxine plus triiodothyronine: relationship to cause of hypothyroidism. Int J Neuropsychopharmacol 2000; 3: 167-174.

9.8.1.36. Walsh JP, Shiels L, Lim EM, et al. Combined thyroxine/liothyronine treatment does not improve well-being, quality of life, or cognitive function compared to thyroxine alone: a randomized controlled trial in patients with primary hypothyroidism. J Clin Endocrinol Metab 2003; 88: 4543-4550.

9.8.1.37. Sawka AM, Gerstein HC, Marriott MJ, et al. Does a combination regimen of thyroxine (T4) and 3, 5,3’-triiodothyronine improve depressive symptoms better than T4 alone in patients with hypothyroidism? Results of a double-blind, randomized, controlled trial. J Clin Endocrinol Metab 88:4551-4555, 2003.

References for Section 9.8.2.

9.8.2.1. Roti E, Minelli R, Gardini E, Braverman LE: The use and misuse of thyroid hormone. Endocr Rev 1993; 14: 401-423.

9.8.2.2. Oppenheimer JH, Braverman LE, Toft A, et al.: Thyroid hormone treatment: when and what? J Clin Endocrinol Metab 1995; 80: 2875-2883.

9.8.2.3. Singer PA, Cooper DS, Levy EG, et al.: Treatment guidelines for patients with hyperthyroidism and hypothyroidism. JAMA 1995; 273: 808-812.

9.8.2.4. Vanderpump MPJ, Ahlquist JAO, Franklyn JA, Clayton RN: Consensus statement for good practice and audit measures in the management of hypothyroidism and hyperthyroidism. Brit Med J 1996; 313: 539-544.

9.8.2.5. Bearcroft CP, Toms GC, Williams SJ et al: Thyroxine replacement in post-radioiodine hypothyroidism. Clin Endocrinol 1991; 34: 115.

9.8.2.6. Kabadi UM, Jackson T: Serum thyrotropin in primary hypothyroidism. A possible predictor of optimal daily levothyroxine dose. Arch Intern Med 1995; 155: 1046-1048.

9.8.2.7. Santini F, Pinchera A, Marsili A et al. Lean body mass is a major determinant of levothyroxine dosage in the treatment of thyroid diseases. J Clin Endocrinol Metab 2005; 90: 124-127.

9.8.2.8. Hall RCW: Psychiatric effects of thyroid hormone disturbance. Psychosomatics 1983; 24: 7.

9.8.2.9. Josephson AM, Mackenzie TB: Thyroid-induced mania in hypothyroid patients. Br J Psychiatr 1980; 137: 222.

9.8.2.10. Josephson AM, Mackenzie TB: Appearance of manic psychosis following rapid normalization of thyroid status. Am J Psychiatr 1979; 136: 846.

9.8.2.11. Carr K, McLeod DT, Parry G, Thornes HM: Fine adjustment of thyroxine replacement dosage : comparison of the thyrotrophin releasing hormone tests using a sensitive thyrotrophin assay with measurement of free thyroid hormones and clinical assessment. Clin Endcrinol 1988; 28: 325.

9.8.2.12. Fish LH, Schwartz HL, Cavanaugh J, et al.: Replacement dose, metabolism, and bioavailability of levothyroxine in the treatment of hypothyroidism. Role of triiodothyronine in pituitary feedback in humans. New Engl J Med 1987; 316: 764-770.

9.8.2.13. Ferretti E, Persani L, Jaffrain-Rea M-L, Giambona S, Tamburrano G, Beck-Peccoz P. Evaluation of the adequacy of levothyroxine replacement therapy in patients with central hypothyroidism. J Clin Endocrinol Metab 1999; 84: 924-929.

9.8.2.14. Shimon I, Cohen O, Lubetsky A, Olchovsky D. Thyrotropin suppression by thyroid hormone replacement is correlated with thyroxine level normalization in central hypothyroidism. Thyroid 2002; 12: 823-827.

9.8.2.14. Grebe SKG, Cooke RR, Ford HC, et al.: Treatment of hypothyroidism with once weekly thyroxine. J Clin Endocrinol Metab 1997; 82: 870-875.

9.8.2.15. Wekking EM, Appelhof BC, Fliers E et al. Cognitive functioning and well-being in euthyroid patients on thyroxine therapy for hypothyroidism. Eur J Endocrinol 2005; 153: 747-753.

9.8.2.16. Saravanan P, Chau WF, Roberts N, Vedhara K, Greenwood R, Dayan CM. Psychological well-being in patients on ‘adequate ’ doses of L-thyroxine: results of a large, controlled community-based questionnaire study. Clin Endocrinol 2002; 57: 577-585.

9.8.2.17. Wiersinga WM. Do we need still more trials on T4 and T3 combination therapy in hypothyroidism? Eur J Endocrinol 2009; 161: 955-959.

9.8.2.18. Grozinsky-Glasberg S, Fraser A, Nahshoni E, Weizman A, Leibovici L. Thyroxine-triiodothyronine combination therapy versus thyroxine monotherapy for clinical hypothyroidism: meta-analysis of randomized controlled trials. J Clin Endocrinol Metab 2006; 91: 2592-2599.

9.8.2.19. Nygaard B, Jensen EW, Kvetny J, Jarlov A, Faber J. Effect of combination therapy with thyroxine (T4) and 3,5,3’-triiodothyronine (T3) versus T4 monotherapy in patients with hypothyroidism, a double-blind, randomized cross-over study. Eur J Endocrinol 2009;161: 895-902.

9.8.2.20. Bianco AC, Salvatore D, Gereben B, Berry MJ, Larsen PR. Biochemistry, cellular and molecular biology, and physiological roles of the iodothyronine selenodeiodinases. Endocr Rev 2002; 23: 38-89.

9.8.2.21. Escobar-Morreale HF, Botella-Carretero JI, Escobar del Rey F, Morreale de Escobar G. Review. Treatment of hypothyroidism with combinations of levothyroxine plus liothyronine. J Clin Endocrinol Metab 20005; 90: 4946-4954.

9.8.2.22. Saravanan P, Simmons DJ, Greenwood R, Peters TJ, Dayan CM. Partial substitution of thyroxine (T4) with triiodothyronine in patients on T4-replacement therapy: results of a large community-based randomized controlled trial. J Clin Endocrinol Metab 2005; 90: 805-812.

9.8.2.23. Appelhof BC, Fliers E, Wekking EM et al. Combined therapy with levothyroxine and liothyronine in two ratios, compared with levothyroxine monotherapy in primary hypothyroidism: a double-blind, randomized, controlled clinical ytrial. J Clin Endocrinol Metab 2005; 90: 2666-2674.

9.8.2.24. Kaplan MM, Sarne DH, Schneider AB. In search of the impossible dream? Thyroid hormone replacement that treats all symptoms in all hypothyroid patients. J Clin Endocrinol Metab 2003; 88: 4540-4542.

9.8.2.25. Jonklaas J, Davidson B, Bhagat S, Soldin SJ. Triiodothyronine levels in athyreotic individuals during levothyroxine therapy. JAMA 2008; 299: 769-777.

9.8.2.26. Walsh JP, Ward LC, Burke V et al. Small changes in thyroxine dosage do not produce measurable changes in hypothyroid symptoms, well-being, or quality-of-life: results of a double-blind, randomized clinical trial. J Clin Endocrinol Metab 2006; 91: 2624-2630.

9.8.2.27. Saranavan P, Siddique H, Simmons DJ, Greenwood R, Dayan CM. Twenty-four hour hormone profiles of TSH, free T3 and free T4 in hypothyroid patients on combined T3/T4 therapy. Exp Clin Endocrinol Diab 2007; 115: 261-267.

9.8.2.28. Henneman G, Docter R, Visser TJ, Postema PT, Krenning EP. Thyroxine plus low-dose, slow-release triiodothyronine replacement in hypothyroidism. Thyroid 2004; 14: 271-275.

9.8.2.29. Russell W, Harrison RF, Smith N, Darzy K, Shalet S, Weetman AP, Ross RJ. Free triiodothyronine has a distinct circadian rhythm that is delayed but parallels thyrotropin levels. J Clin Endocrinol Metab 2008; 93: 2300-2306.

9.8.2.30. Dayan CM, Panicker V. Novel insights into thyroid hormones from the study of common genetic variation. Nature Reviews Endocrinology 2009; 5: 211-218.

9.8.2.31. Saravanan P, Visser TJ, Dayan CM. Psychological well-being correlates with free thyroxine but not free 3,5,3’-triiodothyronine levels in patients on thyroid hormone replacement. J Clin Endocrinol Metab 2006; 91: 3389-3393.

9.8.2.32. Panicker V, Saravanan P, Vaidya B et al. Common variation in the DIO 2 gene predicts baseline psychological well-being and response to combination thyroxine plus triiodothyroinine therapy in hypothyroid patients. J Clin Endocrinol Metab 2009; 94: 1623-1629.

9.8.2.33. Appelhof BC, Peeters RP, Wiersinga WM et al. Polymorhisms in type 2 deiodinase are not associated with well-being, neurocognitive functioning, and preference for combined thyroxine/3,5,3’-triiodothyronine therapy. J Clin Endocrinol Metab 2005; 90: 6296-6299.

9.8.2.34. van der Deure WM, Appelhof BC, Peeters RP et al. Polymorphisms in the brain-specific thyroid hormone transporter OATPC1 are associated with fatigue and depression in hypothyroid patients. J Clin Endocrinol Metab 2008; 69: 804-811.

9.8.2.35. Kim BW, Bianco AC. For some, L-thyroxine replacement might not be enough: a genetic rationale. J Clin Endocrinol Metab 2009; 94: 1521-1523.

9.8.2.36. Sawin CT, Geller A, Wolf PA, et al.: Low serum thyrotropin concentrations as a risk factor for atrial fibrillation in older persons. New Engl J Med 1994; 331: 1249-1252.

9.8.2.37. Biondi B, Fazio S, Carella C, et al.: Cardiac effects of long term thyrotropin-suppressive therapy with levothyroxine. J Clin Endocrinol Metab 1993; 77: 334-338.

9.8.2.38. Leese GP, Jung RT, Guthrie C, et al.: Morbidity in patients on L-thyroxine: a comparison of those with a normal TSH to those with a suppressed TSH. Clin Endocrinol 1992; 37: 500-503.

9.8.2.39. Uzzan B, Campos J, Cucherat M, et al.: Effects on bone mass of long term treatment with thyroid hormones: a meta-analysis. J Clin Endocrinol Metab 1996; 81: 4278-4289

9.8.2.40. Leese GP, Jung RT, Guthrie C, et al.: Morbidity in patients on L-thyroxine: a comparison of those with a normal TSH to those with a suppressed TSH. Clin Endocrinol 1992; 37: 500-503.

9.8.2.41. Solomon BL, Wartofsky L, Burman KD: Prevalence of fractures in postmenopausal women with thyroid disease. Thyroid 1993; 3: 17-23.

9.8.2.42. Shimon I, Cohen O, Lubetsky A, Olchovsky D. Thyrotropin suppression by thyroid hormone replacement is correlated with thyroxine level normalization in central hypothyroidism. Thyroid 2002; 12:823-827.

9.8.2.43. Vestergaard P, Weeke J, Hoeck HC, et al. Fractures in patients with primary idiopathic hypothyroidism. Thyroid 2000; 10:335-340.

9.8.2.44. Sheppard MC, Holder R, Franklyn J. Levothyroxine treatment and occurrence of fracture of the hip. Arch Int Med 2002; 162: 338-343.

9.8.2.45. Flynn RW, Bonellie SR, Jung RT, MacDonald TM, Morris AD, Leese GP. Serum thyroid-stimulating hormone concentration and morbidity from cardiovascular disease and fractures in patients on long-term thyroxine therapy. J Clin Endocrinol Metab 2010; 95: 186-193.

9.8.2.46. Somwaru LL, Amold AM, Joshi N, Fried LP, Cappola AR. Thyroid function test abnormalities are common in elderly people taking thyroid hormone replacement medications. J Clin Endocrinol Metab 2009; 94: 1342-1345.

9.8.2.47. Roos A, Linn-Rasker SP, van Domburg RT, Tijssen JP, Berghout A. The starting dose of levothyroxine in primary hypothyroidism treatment. A prospective, randomized, double-blind trial. Arch Int Med 2005; 165: 1714-1720.

9.8.2.48. Wartofsky L. Levothyroxine therapy for hypothyroidism. Should we abandon conservative dosage titration? Arch Int Med 2005; 165: 1683-1684.

9.8.2.49, Bolk N, Visser TJ, Kalsbeek A, van Domburg RT, Berghout A. Effects of evening vs morning thyroxine ingestion on serum thyroid hormone profiles in hypothyroid patients. Clin Endocrinol 2007; 66: 43-48.

9.8.2.50. Benvenga S, Bartolone L, Pappalardo MA et al. Altered intestinal absorption of L-thyroxine caused by coffee. Thyroid 2008; 18: 293-301.

9.8.2.51. Garber JR, Cobin RH, Gharib H et al. Clinical practice guidelines for hypothyroidism in adults: cosponsored by the American Association of Clinical Endocrinologists and the American Thyroid Association. Thyroid 2012; 22: 1200-1235.

9.8.2.52. Gullo D, Latina A, Frasca F, Le Moli R, Pelegritti G, Vigneri R. Levothyroxine monotherapy cannot guarantee euthyroidism in all athyreotic patients. PLoS ONE 2011; 6: e22552.

9.8.2.53. Beck-Peccoz P. Treatment of central hypothyroidism. Clin Endocrinol 2011; 74: 671-672.

9.8.2.54. Karmisholt J, Andersen S, Laurberg P. Weight loss after therapy of hypothyroidism is mainly caused by excretion of excess body water associated with myxoedem. J Clin Endocrinol Metab 2011; 96: E99-E103.

9.8.2.55. Walker JN, Shillo P, Ibbotson V et al. A thyroxine absorption test followed by weekly thyroxine administration: a method to assess non-adherence to treatment. Eur J Endocrinol 2013; 169: 913-917.

References for Section 9.8.3.

9.8.3.1. Topliss DJ, Wright JA, Volpe R: Increased requirement for thyroid hormone after a jejunal bypass operation. Can Med Assoc J 1980; 123: 765-766.

9.8.3.1. Liel Y, Harman-Boehm I, Shany S: Evidence for a clinically important adverse effect of fiber-enriched diet on the availability of levothyroxine in adult hypothyroid patients. J Clin Endocrinol Metab 1996; 81: 857-859.

9.8.3.2. Bell DSH, Ovalle F. Use of soy protein supplement and resultant need for increased dose of levothyroxine. Endocr Pract 2001; 7: 193-194.

9.8.3.3. Chiu AC, Sherman SI. Effects of pharmacological fiber supplements on levothyroxine absorption. Thyroid 1998; 8: 667-671,

9.8.3.4. Centanni M, Gargano L, Canettieri G et al. Thyroxine in goiter, Helicobacter infection, and chronic gastritis. N Engl J Med 2006; 354: 1787-1795.

9.8.3.5. Checchi S, Montanaro A, Pasqui L et al. L-thyroxine requirement in patients with autoimmune hypothyroidism and parietal cell antibodies. J Ckin Endocrinol Metab 2008; 93: 465-469.

9.8.3.6. Sachmechi I, Reich DM, Aninyei M et al. Effect of proton pump inhibitors on serum thyroid-stimulating hormone level in euthyroid patients treated with levothyroxine for hypothyroidism. Endocr Pract 2007; 13: 345-349.

9.8.3.7.Dietrich JW, Gieselbrecht K, Holl RW et al. Absorption kinetics of levothyroxine is not altered by proton-pump inhibitor therapy. Horm Metab Res 2006; 38: 57-59.

9.8.3.8. Ananthakrishnan S, Braverman LE, Levin RM et al. The effect of famotidine, esomeprazole, and ezetimibe on levothyroxine absorption. Thyroid 2008; 18: 493-498.

9.8.3.9. D’Esteve-Bonetti L, Bennet AP, Malet D et al. Gluten-induced enteropathy (coeliac disease) revealed by resistance to treatment with levothyroxine and alfacalcidol in a sixty-eight-year old patient: a case report. Thyroid 2002; 12: 633-636.

9.8.3.10. McDermott JH, Coss A, Walsh CH. Celiac disease presenting as resistant hypothyroidism. Thyroid 2005; 15: 386-388.

9.8.3.11. Azizi F, Belur R, Albano J. Malabsorption of thyroid hormones after jejunoileal bypass for obesity. Ann Intern Med 1979; 90: 941-942.

9.8.3.12. Topliss DJ, Wright JA, Volpe R. Increased requirement for thyroid hormone after a jejunal bypass operation. Can Med Assoc J 1980; 123: 765-766.

9.8.3.13. Smyrniotis V, Vaos N, Arkadopoulos N, Kostopanagiotou G, Theodoraki K, Lambrou A. Severe hypothyroidism in patients dependent on prolonged thyroxine infusion through a jejunostomy. Clin Nutr 2000; 19: 65-67.

9.8.3.14. Stone E, Leiter LA, Lambert JR, Silverberg JDH, Jeejeebhoy KN, Burrow GN. L-thyroxine absorption in patients with short bowel. J Clin Endocrinol Metab 1984; 59: 139-141.

9.8.3.15. Munoz-Torres M, Varsavsky M, Alonso G. Lactose intolerance revealed by severe resistance to treatment with levothyroxine. Thyroid 2006; 16: 1171-1173.

9.8.3.16. Seppel T, Rose F, Schlaghecke R. Chronic intestinal Giardiasis with isolated thyroxine malabsorption as reason for severe hypothyroidism – implications for localization of thyroid hormoine absorption iin the gut. Exp Clin Endocrinol Diab 1996; 104: 180-182.

9.8.3.17. Witztum JL, Jacobs LS, Schonfeld G: Thyroid hormone and thyrotropin levels in patients placed on colestipol hydrochloride. J Clin Endocrinol Metab 1978; 46: 838-840.

9.8.3.18. Northcutt RC, Stiel JN, Hollifield JW et al. The influence of cholestyramine on thyroxine absorption. JAMA 1969; 208: 1857-1861.

9.8.3.19. Harmon SM, Seifert CF: Levothyroxine-cholestyramine interaction reemphasized. Ann Int Med 1991; 115: 658-659.

9.8.3.20. Weitzman SP, Ginsburg KC, Carlson HE. Colesevelam hydrochloride and lanthanum carbonate interfere with the absorption of levothyroxine. Thyroid 2009; 19: 77-79.

9.8.3.21. Liwanpo L, Hershman JM. Conditions and drugs interfering with thyroxine absorption. Best Pract Res Clin Endocrinol Metab 2009; 23: 781-792.

9.8.3.22. Havrankova J, Lahaie R: Levothyroxine binding by sucralfate. Ann Int Med 1992; 147: 445-446 [Erratum 1993; 118:398].

9.8.3.23. Sherman SI, Tielens ET, Ladenson PW. Sucralfate causes malabsorption of L-thyroxine. Am J Med 1994; 90: 531-535.

9.8.3.24. Campbell JA, Schmidt BA, Bantle JP. Sucralfate and the absorption of L-thyroxine. Ann Int Med 1994; 121: 152.

9.8.3.25. Liel Y, Sperber AD, Shany S. Nonspecific intestinal adsorption of levothyroxine by aluminium hydroxide. Am J Med 1994; 97: 363-365.

9.8.3.26. Mersebach H, Rasmussen AK, Kirkegaard L et al. Intestinal adsorption of levothyroxine by antacids and laxatives: case stories and in vitro experiments. Pharmacol Toxicol 1999; 84: 107-109.

9.8.3.27. Sperber AD, Liel Y: Evidence for interference with the intestinal absorption of levothyroxine sodium by aluminium hydroxide. Arch Int Med 1992; 152: 183-184.

9.8.3.28. Campbell NRC, Hasinoff BB, Stalts H, et al.: Ferrous sulfate reduces thyroxine efficacy in patients with hypothyroidism. Ann Int Med 1992; 117: 1010-1013.

9.8.3.29. Shakir KM, Chute JP, Aprill BS et al. Ferrous sulfate-induced increase in requirement for thyroxine in a patient with primary hypothyroidism. South Med J 1997; 90: 637-639.

9.8.3.30. Leger CS, Ooi TC. Ferrous fumarate-induced malabsorption of thyroxine. Endocrinologist 1999; 9: 493-495.

9.8.3.31. Singh N, Singh PN, Hershman JM. Effect of calcium carbonate on the absorption of levothyroxine. JAMA 2000; 283: 2822-2825.

9.8.3.32. Singh N, Weisler SL, Hershman JM. The acute effect of calcium carbonate on the intestinal absorption of levothyroxine. Thyroid 2001; 11: 967-971.

9.8.3.33. Csako G, McGriff NJ, Rotman-Pikielny P, Sarlis NJ, Pucino F. Exaggerated levothyroxine malabsorption due to calcium carbonate supplementation in gastrointestinal disorders. Ann Pharmacother 2001; 35: 1578-1583.

9.8.3.34. John-Kalarickal J, Pearlman G, Carlson HE. New medications which decrease levothyroxine adsorption. Thyroid 2007; 17: 763-765.

9.8.3.35. Weitzman SP, Ginsburg KC, Carlson HE. Colesevelam hydrochloride and lanthanum carbonate interfere with the absorption of levothyroxine. Thyroid 2009; 19: 77-79.

9.8.3.36. Diskin CJ, Stokes TJ, Dansby LM et al. Effect of phosphate binders upon TSH and L-thyroxine dose in patients on thyroid replacement. Int Urol Nephrol 2007; 39: 599-602.

9.8.3.37. Siraj ES, Gupta MK, Reddy SS. Raloxifene causing malabsorption of levothyroxine. Arch Int Med 2003; 163: 1367-1370.

9.8.3.38. Garwood CL, van Schepen KA, McDonough RP et al. Increased thyroid-stimulating hormone levels associated with concomitant administartion of levothyroxine and raloxifene. Pharmacotherapy 2006; 26: 881-885.

9.8.3.39. Madhava K, Hartley A. Hypothyroidism in thyroid carcinoma follow-up: orlistat may inhibit the absorption of thyroxine. Clin Oncol 2005; 17: 492.

9.8.3.40. Ain KB, Mori Y, Refetoff S. Reduced clearance rate of thyroxine-binding globulin (TBG) with increased sialylation: a mechanism for estrogen induced elevation of serum TBG concentration. J Clin Endocrinol Metab 1987; 65: 689-696.

9.8.3.41. Tahboub R, Arafah BM. Sex steroids and the thyroid. Best Pract Res Clin Endocrinol Metab 2009; 23: 769-780.

9.8.3.42. Sitruk-Ware R, Plu-Bureau G, Menard J et al. Effects of transvaginal ethinyl estradiol on hemostatic factors and hepatic proteins in a randomized, cross over study. J Clin Endocrinol Metab 2007; 92: 2074-2079.

9.8.3.43. Goebelsman IJ, Mashchak CA, Mishell Jr DR. Comparison of hepatic impact of oral and vaginal administration of ethinyl estradiol. Am J Obstetr Gynecol 1985; 151: 868-877.

9.8.3.44. Bisschop PH, Toorians AW, Endert E et al. The effects of sex-steroid administration on the pituitary-thyroid axis in transsexuals. Eur J Endocrinol 2006; 155: 11-16.

9.8.3.45. Arafah BM. Increased need for thyroxine in women with hypothyroidism during estrogen therapy. N Eng J Med 2001; 344:1743-1749.

9.8.3.46. Anker GB, Lonning PE, Aakvaag A et al. Thyroid function in postmenopausal breast cancer patients treated with tamoxifen. Scand J Clin Lab Invest 1998; 58: 103-107.

9.8.3.47. Marqusee E, Braverman LE, Lawrence J et al. The effect of droloxofene and estrogen on thyroid function in postmenopausal women. J Clin Endocrinol Metab 2000; 85: 4407-4410.

9.8.3.48. Van Bon AC, Wiersinga WM. Goserelin-induced transient thyrotoxicosis in a hypothyroid women on L-thyroxine replacement. Neth J Med 2008; 66: 256-258.

9.8.3.49. Mandell SJ, Larson PR, Seely EW, Brent GA: Increased need for thyroxine during pregnancy in women with rpimary hypothyroidism. New Engl J Med 1990; 323: 91-96.

9.8.3.50. Kaplan MM: Monitoring thyroxine treatment during pregnancy. Thyroid 1992; 2: 147-152.

9.8.3.51. Smallridge RC, Ladenson PW. Hypothyroidism in pregnancy: consequences to neonatal health. J Clin Endocrinol Metab 2001; 86: 2349-2353.

9.8.3.52.Alexander EK, Marqusee E, Lawrence J, Jarolim P, Fischer GA, Larsen PR. Timing and magnitude of increases in levothyroxine requirements during pregnancy in women with hypothyroidsm. N Engl J Med 2004; 351: 241-249.

9.8.3.53. Toft A. Increased levothyroxine requirements in pregnancy – Why, When, and How much? N Engl J Med 2004; 351: 292-294.

9.8.3.54. Yassa L, Marqusee E, Fawcett R, Alexander EK. Thyroid hormone early adjustment in pregnancy (the THERAPY) trial. J Clin Endocrinol Metab 2010; 95: 3234-3241.

9.8.3.55. Haddow JE, Palomaki GE, Allen WC et al. Maternal thyroid deficiency during pregnancy and subsequent neuropsychological development of the child. N Engl J Med 1999; 341: 549-555.

9.8.3.56. Pop VJ, Kuypens JL, Baar AL van, et al. Low maternal free thyroxine concentrations during early pregnancy are associated with impaired psychomotor development in infancy. Clin Endocrinol 1999; 50: 149-155.

9.8.3.57. Henrichs J, Bongers-Schokking JJ, Schenk JJ. Maternal thyroid function during early pregnancy and cognitive functioning in early childhood: the Generation R study. J Clin Endocrinol Metab 2010; 95 (doi:10.2010/jc.2010-0415) .

9.8.3.58. Morreale de Escobar G, Obregon MJ, Escobar del Rey F. Is neuropsychological development related to maternal hypothyroidism or maternal hypothyroxinemia? J Clin Endocrinol Metab 2000; 85: 3975-3987.

9.8.3.66.Illouz F, Laboureau-Soares S, Dubois S, Rohmer V, Rodien P. Tyrosine kinase inhibitors and modifications of thyroid function. Eur J Endocrinol 2009; 160: 331-336.

9.8.3.67. de Groot JW, Zonnenberg BA, Plukker JT, van der Graaf WT, Links TP, Imatinib induces hypothyroidism in patients receiving levothyroxine. Clin Pharmacol Ther 2005; 78: 433-438.

9.8.3.68. de Groot JW, Zonnenberg BA, van Ufford-Mannesse PQ et al. A phase II trial of imatinib therapy for metastatic medullary thyroid carcinoma. J Clin Endocrinol Metab 2007; 92: 3466-3469.

9.8.3.69. Pacini F, Sherman S, Schlumberger M et al. Exacerbation of postsurgical hypothyroidism during treatment of advanced differentiated (DTC) or medullary (MTC) thyroid carcinoma with AMG 706. Horm Res 2007; 68:29.

9.8.3.70. Sherman S, Wirth L, Droz JP et al. Motesanib diphosphate in progressive differentiated thyroid cancer. N Engl J Med 2008; 359: 31-42.

9.8.3.71. Gupta-Abramson V, Troxel AB, Nellore A et al. Phase II trial of sorafenib in advanced thyroid cancer. J Clin Oncol 2008; 26: 4714-4719.

9.8.3.72. Abdulrahman RM, Verloop H, Hoftijzer H et al. Sorafenib-induced hypothyroidism is associated with increased type 3 deiodination. J Clin Endocrinol Metab 2010; 95: 3758-3762..

9.8.3.59. Curran PG, DeGroot LJ: The effect of hepatic enzyme-inducing drugs on thyroid hormones and the thyroid gland. Endocr Rev 1991; 12:135.

9.8.3.60. Faber J, Lumholtz IB, Kirkegaard C et al.: The effects of phenytoin on the extrathyroidal turnover of thyroxine, 3,5,3'-triiodothyronine, 3,3',5'-triiodothyronine, and 3',5'-diiodothyronine in man. J Clin Endocrinol Metab 1985; 61: 1093.

9.8.3.61. DeLuca F, Arrigo T, Pandullo E et al.: Changes in thyroid function tests induced by 2 month carbamazepine treatment in L-thyroxine-substituted hypothyroid children. Eur J Pediatr 1986; 145: 77.

9.8.3.62. Isley WL: Effect of rifampin therapy on thyroid function tests in a hypothyroid patient on replacement L-thyroxine. Ann Int Med 1987; 107: 517-518

9.8.3.63. Figge J, Dluhy RG: Amiodarone-induced elevation of thyroid stimulating hormone in patients receiving levothyroxine for primary hypothyroidism. Ann Int Med 1990; 113: 563-565.

9.8.3.64. McCowen KC, Garber JR, Spark R: Elevated serum thyrotropin in thyroxine-treated patients with hypothyroidism given sertraline. New Engl J Med 1997; 337: 1010-1011.

9.8.3.65. Munera Y, Hugues FC, Le Jeunne C, Pays IF: Interaction of thyroxine sodium with antimalarial drugs. Brit Med J 1997; 314: 1593.

9.8.3.73. Arafah BM: Decreased levothyroxine requirement in women with hypothyroidism during androgen therapy for breast cancer. Ann Int Med 1994; 121: 247-251.

9.8.3.74. Cunningham JJ, Barzel NS: Lean body mass is a predictor of the daily requirement of thyroid hormone in older men and women. J Am Geriatr Soc 1984; 32: 204-207.

9.8.3.75. Rosenbaum RL, Barzel US. Levothyroxine replacement dose for primary hypothyroidism decreases with age. Ann Int Med 1982; 96: 53-55.

9.8.3.76. Griffin JE: Hypothyroidism in the elderly. Am J Med Sci 1990; 299: 334-345.

9.8.3.77. Thomas MC, Mathew TH, Russ GR. Changes in thyroxine requirements in patients with hypothyroidism undergoing renal transplantation. Am J Kidney Dis 2002; 39: 354-357.

9.8.3.78. Padwal R, Brocks D, Sharma AM. A systematic review of drug absorption following bariatric surgery and its theoretical implications. Obes Rev 2010; 11: 41-50.

9.8.3.79. Rubio IG, Gairao AL, Santo MA, Zanini AC, Medeiros-Neto G. Levothyroxine absorption in morbidly obese patients before and after Roux-en-Y gastric bypass (RYGB) surgery. Obes Surg 2012; 22: 253-258.

9.8.3.80. Pirola I, Formenti AM, Gandossi E et al. Oral liquid L-thyroxine (L-T4) may be better absorbed compared to L-T4 tablets. Obes Surg 2013; 23: 1493-1496.

9.3.8.81. Gkotsina M, Michelaki M, Mamali I et al. Improved levothyroxine pharmacokinetics after bariatric surgery. Thyroid 2013; 23: 414-419.

9.3.8.82. Braun D, Kim TD, le Coutre P, Kohrle J, Hershman JM, Schweizer U. Tyrosine kinase inhibitors noncompetitively inhibit MCT8-mediated iodothyronine transport. J Clin Endocrinol Metab 2012; 97: E100-E105.

9.8.3.83. de Carvalho GA, Bahls SC, Boeving A, Graf H. Effects of selective serotonin reuptake inhibitors on thyroid function in depressed patients with primary hypothyroidism or normal thyroid function. Thyroid 2009; 19: 691-697.

References for Section 9.8.4.

9.8.4.1, Murray JS, Jayarajasingh R, Perros P. Deterioration of symptoms after start of thyroid hormone replacement. Brit Med J 2001; 323: 332-333.

9.8.4.2. Topliss DJ, White EL, Stockigt JR. Significance of thyrotropin excess in untreated primary adrenal insufficiency. J Clin Endocrinol Metab 1980; 50: 52-56.

9.8.4.3. Shigemasa C, Kouchi T, Ueta Y et al. Evaluation of thyroid function in patients with isolated adrenocorticotropin deficiency. Am J Med Sci 1992; 304: 279-284.

9.8.4.4. Abdullatif HD, Ashraf AP. Reversible subclinical hypothyroidism in the presence of adrenal insufficiency. Endocr Pract 2006; 12: 572-575.

9.8.4.5. Samuels MH. Effects of variations in physiological cortisol levels on thyrotropin secretion in subjects with adrenal insufficiency: a Clinical Research Center Study. J Clin Endocrinol Metab 2000; 85: 1388-1393.

9.8.4.6. Santos AD, Miller RP, Puthenpurakal KM et al.: Echocardiographic characterization of the reversible cardiomyopathy of hypothyroidism. Am J Med 1980; 68: 675.

9.8.4.7. Polikar R, Burger AG, Scherrer U, Nicod P: The thyroid and heart. Circulation 1993; 87: 1435.

9.8.4.8. Ladenson PW: Recognition and management of cardiovascular disease related to thyroid dysfunction. Am J Med 1990; 88: 638.

9.8.4.9. Keating FR, Parkin TW, Selby JB, Dickinson LS: Treatment of heart disease associated with myxedema. Prog Cardiovasc Dis 1961; 3: 364.

9.8.4.10. Hay ID, Duick DX, Vlietstra RE et al.: Thyroxine therapy in hypothyroid patients undergoing coronary revascularization: a retrospective analysis. Ann Intern Med 1981: 95: 456.

9.8.4.11. Weinberg AD, Brennan MD, Gorman CA et al.: Outcome of anesthesia and surgery in hypothyroid patients. Arch Intern Med 1983; 143: 893.

9.8.4.12. Ladenson PW, Levin AA, Ridgway EC, Daniels GH: Complications of surgery in hypothyroid patients. Am J Med 1984; 77: 261.

9.8.4.13. Sherman SI, Ladenson PW: Percutaneous transluminal coronary angioplasty in hypothyroidism. Am J Med 1991; 90: 367-370.

9.8.4.14. Levine HD: Compromise therapy in the patient with angina pectoris and hypothyroidism: a clinical assessment. Am J Med 1980; 69: 411.

9.8.4.15. Behan LA, Monson JP, Agha A. The interaction between growth hormone and the thyroid axis in hypopituitary patients. Clin Endocrinol 2010; Apr 23 [Epub ahead of print].

9.8.4.16. Martins MR, Doin FC, Komatsu WR, Barros-Neto TL, Moises VA, Abucam J. Growth hormone replacement improves thyroxine biological effects: implications for management of central hypothyroidism. J Clin Endocrinol Metab 2007; 92: 4144-4153.

9.8.4.17. Porretti S, Giavoli C, Ronchi C, et al. Recombinant human GH replacement therapy and thyroid function in a large group of adult GH-deficient patients. When does L-T4 therapy becomes mandatory? J Clin Endocrinol Metab 2002; 87: 2042-2045.

9.8.4.18. Chonchol M, Lippi G, Salvagno G, Zoppini G, Muggeo M, Targher G. Prevalence of subclinical hypothyroidism in patients with chronic kidney disease. Clin J Am Soc Nephrol 2008; 3: 1296-1300.

References for Section 9.9.

9.9.1. Nicoloff JT, LoPresti JS: Myxedema coma. A form of decompensated hypothyroidism. Endocrinol Metab Clin North Am 1993; 22: 279-290.

9.9.2. Jordan RM: Myxedema coma. Pathophysiology, therapy, and factors affecting prognosis. Med Clin North Am 1995; 79: 185-194.

9.9.3. Fliers E, Wiersinga WM. Myxedema coma. Rev Endocr Metab Dis 2003; 4: 137-141.

9.9.4. Wartofsky L. Myxedema coma. Endocrinol Metab Clin N Am 2006; 35: 687-698.

9.9.5. Kwaku MO, Burman Kd. J Intensive Care Med 2007; 22: 224-231.

9.9.6. Dutta P, Bhansali A, Masoodi SR, Bhadada S, Sharma N, Rajput R. Predictors of outcome in myxedema coma: a study from a tertiary care centre. Crit Care 2008; 12: R1

9.9.7. Reinhardt W, Mann K. Incidence, clinical picture, and treatment of hypothyroid coma: results of a survey. Med Klin 1997; 92: 521-524.

9.9.8. Rodriguez I, Fluiters E, Perez-<emdez LF, Luna R, Paramo C, Garcia-Mayor RV. Factors associated with mortality of patients with myxedema coma: prospective study in 11 cases treated in a single institution. J Endocrinol 2004; 180: 347-350.

9.9.9. Holvey DN, Goodner CJ, Nicoloff JT, Dowling JT: Treatment of myxedema coma with intravenous thyroxine. Arch Intern Med 1964; 113: 139.

9.9.10. Blackburn CM, McConahey WM, Keating RF, Albert A: Calorigenic effect of single intravenous doses of L-triiodothyronine and L-thyroxine in myxedematous persons. J Clin Invest 1954; 33: 819.

9.9.11. MacKerrow SD, Osborn LA, Levy H et al.: Myxedema-associated cardiogenic shock treated with intravenous triiodothyronine. Ann Intern Med 1992; 117: 1014.

9.9.12. Hylander B, Rosenquist U: Treatment of myxedema coma - factors associated with fatal outcome. Acta Endocrinol 1985; 108: 65.

9.9.13. Arlot S, Debussche X, Lalau JD et al.: Myxedema coma: response of thyroid hormones with oral and intravenous high-dose L-thyroxine treatment. Intensive Care Med 1991; 17: 16.

9.9.14. Ridgway EC et al. Metabolic responses of patients with myxedema to large doses of intravenous L-thyroxine. Ann Int Med 1972; 77: 549-555.

9.9.15. Yamamoto T, Fukuyama J, Fujiyoshi A. Factors associated with mortality of myxedema coma: report of eight cases and literature survey. Thyroid 1999; 9: 1167-1174.

9.9.16. McKerrow SD, Osborn LA, Levy H, Eaton P, Economon O. Myxedema-associated cardiogenic shock treated with triiodothyronine. Ann Int Med 1992; 117: 1014-1015.

9.9.17. Ladenson PW, Levin AA, Ridgway EC, Daniels GH. Complications of surgery in hypothyroid patients. Am J Med 1984; 77: 261.

9.9.18. Sheu CC, Cheng MH, Tsai JR, Hwang JJ. Myxedema coma: a well-known but unfamiliar medical emergency. Thyroid 2007; 17: 371-372.9.9.19. Malliphedi A, Vali H, Okosieme O. Myxedema coma in a patient with subclinical hypothyroidism. Thyroid 2011; 21: 87-89.

9.9.20. Chu M, Seltzer TF. Myxedema coma induced by ingestion of raw bol choy. New Engl J Med 2010; 362: 1945-1946.

References for Section 9.10.1.

9.10.1.1. Hollowell JG, Staehling NW, Flanders WD et al. Serum TSH, T4 and thyroid antibodies in the United States population (1988 to 1994): National Health and Nutrition Examination Survey (NHANES III). J Clin Endocrinol Metab 2000; 87: 489-499.

9.10.1.2. National Academy of Clinical Biochemistry. NACB laboratory medicine practice guidelines. Available at: http://www.nacb.org/lmpg/main.stm

9.10.1.3. Wartofsky L, Dickey RA. The evidence for a narrower thyrotropin reference range is compelling. J Clin Endocrinol Metab 2005; 90: 5483-5488.

9.10.1.4. Surks MI, Goswami G, Daniels GH. The thyrotropoin reference range should remain unchanged. J Clin Endocrinol Metab 2005; 90: 5489-5496.

9.10.1.5. Brabant G, Beck-Peccoz P, Jarzab B et al. Is there a need to redefine the upper normal limit of TSH? Eur J Endocrinol 2006; 154: 633-637.

9.10.1. 6. Kratzsch J, Fielder GM, Leichtle A et al. New reference intervals for thyrotropin and thyroid hormones based on National Academy of Clinical Biochemistry criteria and regular ultrasonography of the thyroid. Clin Chem 2005; 51: 1480-1486.

9,10.1.7. Hamilton TE, Davis S, Onstad L. Kopecky KJ. Thyrotropin levels in a population with no clinical, antibody, or ultrasonographic evidence of thyroid disease: implications for the diagnosis of subclionical hypothyroidism. J Clin Endocrinol Metab 2008; 93: 1224-1230.

9.10.1.8. Surks MI, Hollowell JG. Age-specific distribution of serum thyrotropin and thyroid antibodies in the US population: implications for the prevalence of subclinical hypothyroidism. J Clin Endocrinol Metab 2007; 92: 4575-4582.

9.10.1.9. Karmisholt J, Andersen S, Laurberg P. Interval between tests and thyroxine estimation method influence outcome of monitoring of subclinical hypothyroidism. J Clin Endocrinol Metab 2008; 93: 1634-1640.

9.10.1.10. Karmisholt J, Laurberg P. Serum TSH and serum thyroid peroxidase antibody fluctuate in parallel and high urinary iodine excretion predicts subsequent thyroid failure in a 1-year study of patients with untreated subclinical hypothyroidism. Eur J Endocrinol 2008; 158: 209-215.

9.10.1.11. Brabant G, Dogu E, Kausche H, Prank K, von zur Mühlen A, Adriaanse R, Romijn JA, Wiersinga WM. Temporal pattern of TSH secretion and its significance in evaluating thyroid status from single TSH determinations. Exp Clin Endocrinol 1994; 102: 49-56.

9.10.1.12. De Marco G, Agretti P, Camilot M et al. Functional studies of new TSH receptor (TSHr) mutations identified in patients affected by hypothyroidism or isolated hyperthyrotrophinaemia. Clin Endocrinol 2009; 70: 335-338.

9.10.1.13. Nicoletti A, Bal M, De Marco G et al. Thyrotropin-stimulating hormone receptor gene analysis in pediatric patients with non-autoimmune subclinical hypothyroidism. J Clin Endocrinol Metab 2009; 94: 4187-4194.

9.10.1.14. Rapa A, Monzani A, Moia S et al. Subclinical hypothyroidism in children and adolescents: a wide range of clinical, biochemical and genetic factors involved. J Clin Endocrinol Metab 2009; 94: 2414-2420.

9.10.1.15. Chikunguwo S, Brethauer S, Niruyogi V et al. Influence of obesity and surgical weight loss on thyroid hormone levels. Surg Obes Relat Dis 2007; 6: 631-635.

9.10.1.16. Rotondi M, Leporati P, La Manna A et al. Raised serum TSH levels in patients with morbid obesity: is it enough to diagnose subclinical hypothyroidism? Eur J Endocrinol 2009; 160: 403-408.

9.10.1.17. Szabolcs I, Podoba J, Feldkamp J et al. Comparative screening for thyroid disorders in old age in areas of iodine deficiency, long-term iodine prophylaxis and abundant iodine intake. Clin Endocrinol 1997; 47: 87-92.

9.10.1.18. Laurberg P, Pedersen KM, Hreidarsson A, Sigfusson N, Iversen E, Knudsen PR. Iodine intake and the pattern of thyroid disorders: a comparative epidemiological study of thyroid abnormalities in the elderly in Iceland and in Jutland, Denmark. J Clin Endocrinol Metab 1998; 83: 765-769.

9.10.1.19. Tunbridge WMG, Evered DC, Hall R et al. The spectrum of thyroid disease in the community: the Whickham survey. Clin Endocrinol 1977; 7: 481-493.

9.10.1.20. Canaris GJ, Manowitz NR, Mayor G, Ridgway EC. The Colorado Thyroid Disease Prevalence Study. Arch Intern Med 2000; 160: 526-534.

9.10.1.21. Cooper DS, Biondi B. Subclinical thyroid disease. Lancet 2012; 379: 1142-1154.

References for Section 9.10.2.

9.10.2.1 Tunbridge WMG, Brewis M, French JM, et al.: Natural history of autoimmune thyroiditis. Brit Med J 1980; 282: 258-262.

9.10.2.2. Gordin A, Lamberg BA. Spontaneous hypothyroidism in symptomless thyroiditis. A long-term follow-up study. Clin Endocrinol 1981; 15: 537- 543.

9.10.2.3. Cooper DS, Halpern R, Wood LC, Levin AA, Ridgway EC. L-thyroxine therapy in subclinical hypothyroidism. A double-blind, placebo-controlled trial. Ann Int Med 1984; 101: 18-24.

9.10.2.4. Nyström E, Caidahl K, Fager G, Wikkelso C, Lundberg PA, Lindstedt G. A double-blind cross-over 12-month study of L-thyroxine treatment of women with “subclinical hypothyroidism”. Clin Endocrinol 1988; 29: 63-75.

9.10.2.5. Parle JV, Franklyn JA, Cross KW, et al.: Prevalence and follow-up of abnormal thyrotropin (TSH) concentrations in the elderly in the United Kingdom. Clin Endocrinol 1993; 34: 77-83.

9.10.2.6. Jaeschke R, Guyatt G, gerstein H et al. Does treatment with L-thyroxine influence health status in middle-aged and older adults with subclinical hypothyroidism? J Gen Intern Med 1996; 11: 744-749.

9.10.2.7. Kong WM, Sheikh MH, Lumb PJ et al. A 6-month randomized trial of thyroxine treatment in women with mild subclinical hypothyroidism. Am J Med 2002; 112: 348-354.

9.10.2.8. Huber G, Staub JJ, Meier C et al. Prospective study of the spontaneous course of subclinical hypothyroidism: prognostic factors of thyrotropin, thyroid reserve, and thyroid antibodies. J Clin Endocrinol Metab 2002; 87: 3221-3226.

9.10.2.9. Diez JJ, Iglesias P. Spontaneous subclinical hypothyroidism in patients older than 55 years: an analysis of natural course and risk factors for the development of overt thyroid failure. J Clin Endocrinol Metab 2004; 87: 4890-4897.

9.10.2.10. Gussekloo J, van Exel E, de Craen AJM, Meinders AE, FrohlichM, Westendorp RGJ. Thyroid status, disability and cognitive function, and survival in old age. JAMA 2004; 292: 2591-2599.

9.10.2.11. Diez JJ, Iglesias P, Burman KD. Spontaneous normalization of thyrotropin concentrations in patients with subclinical hypothyroidism. J Clin Endocrinol Metab 2005; 90: 4124-4127.

9.10.2.12. Demirbilek H, Kandemir N, Gonc EN, Ozon A, Alikasifoglu A. Assessment of thyroid function during the long course of Hashimoto’s thyroiditis in children and adolescents. Clin Endocrinol 2009; 71: 451-454.

9.10.2.13. Wasniewska M, Salermo M, Cassio A et al. Prospective evaluation of the natural course of idiopathic subclinical hypothyroidism in childhood and adolescence. Eur J Endocrinol 2009; 160: 417-421.

9.10.2.14. Geul KW, van Sluisveld ILL, Grobbee DE, et al.: The importance of thyroid microsomal antibodies in the development of elevated serum TSH in middle-aged women: Associations with serum lipids. Clin Endocrinol 1993; 39: 275-280.

9.10.2.15. Vanderpump MPJ, Tunbridge WMG, French JM, et al.: The incidence of thyroid disorders in the community: a twenty-year follow-up of the Whickham Survey. Clin Endocrinol 1995; 43: 55-68.

9.10.2.16. Strieder TG, Tijssen JG, Wenzel BE, Endert E, Wiersinga WM. Prediction of progression to overt hypothyroidism or hyperthyroidism in female relatives of patients with autoimmune thyroid disease using the Thyroid Events Amsterdam (THEA) score. Arch Int Med 2008; 168: 1657-1663.

9.10.2.17. Walsh JP, Bremner AP, Feddema P, Leedman PJ, Brown SJ, O’Leary P. Thyrotropin and thyroid antibodies as predictors of hypothyroidism: a 13-year, longitudinal study of a community-based cohort using current immunoassay techniques. J Clin Endocrinol Metab 2010; 95: 1095-1104.

9.10.2.18. Rosario PW, Bessa B, Valadao MM, Purisch S. Natural history of mild subclinical hypothyroidism: prognostic value of ultrasound. Thyroid 2009; 19: 9-12

9.10.2.19. Somwaru LL, Rariy CM, Arnold AM, Cappola AR. The natural history of subclinical hypothyroidism in the elderly: the cardiovascular health study. J Clin Endocrinol Metab 2012; 97: 1962-1969..

9.10.2.20. Sathyapalan T, Manuchehri AM, Thatcher NJ et al. The effect of soy phytoestrogen supplementation on thyroid status and cardiovascular risk markers in patients with subclinical hypothyroidism: a randomized, double-blind, crossover study. J Clin Endocrinol Metab 2011; 96: 1442-1449.

References for Section 9.10.3.

9.10.3.1. Biondi B, Cooper DS. The clinical significance of subclinical thyroid dysfunction. Endocr Rev 2008; 29: 76-131.

9.10.3.2. Staub JJ, Althaus BU, Engler H, et al.: Spectrum of subclinical and overt hypothyroidism: effect on thyrotropin, prolactin, and thyroid reserve, and metabolic impact on peripheral target tissues. Am J Med 1992; 92: 631-641.

  1. Cooper DS, Halpern R, Wood Lc, Levin AA, Ridgway EC. Thyroxine therapy in subclinical hypothyroidism. A double-blind, placebo-controlled trial. Ann Int Med 1984; 101: 18-24.
  1. Zulewski H, Muller B, Exer P, Miserez AR, Staub JJ. Estimation of tissue hypothyroidism by a new clinical score: evaluation of patients with various grades of hypothyroidism and controls. J Clin Endocrinol Metab 1997; 82: 771-776.
  1. Canaris GJ, Manowitz NR, Mayor G, Ridgway EC. The Colorado thyroid disease prevalence study. Arch Int Med 2000; 160: 526-534.

9.10.3.6. Grabe HJ, Volzke H, Ludemann J et al. Mental and physical complaints in thyroid disorders in the general population. Acta Psychiatr Scand 2006; 145: 573-581.

9.10.3.7. Haggerty JJ, Stern RA, Mason GA, et al.: Subclinical hypothyroidism: a modifiable risk factor for depression? Am J Psychiatry: 150; 508-510.

9.10.3.8. Gussekloo J, van Exel E, de Craen AJM, Meinders AE, FrohlichM, Westendorp RGJ. Thyroid status, disability and cognitive function, and survival in old age. JAMA 2004; 292: 2591-2599.

9.10.3.9. Roberts LM, Pattison H, Roalfe A et al. Is subclinical thyroid dysfunction in the elderly associated with depression or cognitive dysfunction? Ann Int Med 2006; 145: 573-581.

9.10.3.10. Jordes R, Waterloo K, Storhaug H et al. Neuropsychologic ial function and symptoms in subjects with subclinical hypothyroidism and the effect of thyroxine treatment. J Clin EndocrinolMetab 2006; 91: 145-153.

9.10.3.11. Bell RJ, Rivera-Woll L, Davidson SL, Topliss DJ, Donath S, Davis SR. Well-being, health-related quality of life and cardiovascular risk profile in women with subclinical thyroid disease – a community –based study. Clin Endocrinol 2007; 66: 548-556,

9.10.3.12. Park YJ, Lee EJ, Lee YJ et al. Subclinical hypothyroidism (SCH) is not associated with metabolic derangement, cognitive impairment, depression or poor quality of life (QoL) in elderly subjects. Arch Gerontol Geriatr 2010; 50: e68-e73.

9.10.3.13. Monzani F, Del Guerra P, Caraccio N, et al.: Subclinical hypothyroidism: neurobehavioral features and beneficial effect of L-thyroxine treatment. Clin Investig 1993; 71: 367-371.

9.10.3.14. Baldini IM, Vita A, Mauri MC et al. Psychopathological and cognitive features in subclinical hypothyroidism. Prog Neuropsychopharmacol Biol Psychiatry 1997; 21: 925-935.

9.10.3.15. Zhu DF, Wang ZX, Zhang DR et al. fMRI revealed neural substrate for reversible working memory dysfunction in subclinical hypothyroidism. Brain 2006; 129: 2923-2930.

9.10.3.16. Correia N, Mullally S, Cooke G et al. Evidence for a specific defect in hippocampal memory in overt and subclinical hypothyroidism. J Clin Endocrinol Metab 2009; 94: 3789-3797.

9.10.3.17. Ripoli A, Pingitore A, Favilli B et al. Does subclinical hypothyroidism affect cardiac pump performance? Evidence from a magnetic resonance imaging study. J Am Coll Cardiol 2005; 45: 439-445.

9.10.3.18. Vitale G, Galderisi M, Lupoli GA et al. Left ventricular myocardial impairment in subclinical hypothyroidism assessed by a new ultrasound tool: pulsed tissue Doppler. J Clin Endocrinol Metab 2002; 87: 4350-4355.

9.10.3.19. Aghini-Lombardi F, Di Bello Vitantonio, Enrica T et al. Early textural and functional alterations of left ventricular myocardium in mild hypothyroidsm. Eur J Endocrinol 2006; 155: 3-9.

9.10.3.20. Baycan S, Erdogan D, Caliskan M et al. Coronary flow reserve is impaired in subclinical hypothyroidism. Clin Cardiol 2007; 30: 562-566.

9.10.3.21. Biondi B, Galderisi M, Pagano L et al. Endothelial-mediated coronary flow reserve in patients with mild thyroid hormone deficiency. Eur J Endocrinol 2009; 161: 323-329.

9.10.3.22. Brenta G, Mutti LA, Schnitman M, Fretes O, Pezzone A, Matute ML. Assessment of left ventricular diastolic function by radionuclide ventriculography at rest and exercise in subclinical hypothyroidism, and its response to L-thyroxine therapy. Am J Cardiol 2003; 91: 1327-1330.

9.10.3.23. Kahaly GJ. Cardiovascular and atherogenic aspects of subclinical hypothyroidism. Thyroid 2000; 10: 665-679.

9.10.3.24. Nagasaki T, Inaba M, Kumeda Y et al. Increased pulse wave velocity in subclinical hypothyroidism. J Clin Endocrinol Metab 2006; 91: 154-158.

9.10.3.25. Nagasaki T, Inaba M, Kumeda Y et al. Central pulse wave velocity is responsible for increased brachial-ankle pulse wave velocity in subclinical hypothyroidism. Clin Endocrinol 2007; 66: 304-308.

9.10.3.26. Owen PJD, Rajiv C, Vinereanu D, Mathew T, Fraser AG, Lazarus JH. Subclinical hypothyroidsm, arterial stiffness and myocardial reserve. J Clin Endocrinol Metab 2006; 91: 2126-2132.

9.10.3.27. Lekakis J, Papamichael C, Alevizaki M, Piperingos G, Marafelia P, Mantzos J. Flow-mediated, endothelium-dependent vasodilatation is impaired in subjects with hypothyroidism, borderline hypothyroidism, and high-normal serum thyrotropin (TSH) values. Thyroid 1997; 7: 411-414.

9.10.3.28. Taddei S, Caraccio N, Virdis A et al. Impaired endothelium-dependent vasodilatation in subclinical hypothyroidism: beneficial effect of levothyroxine therapy. J Clin Endocrinol Metab 2003; 88: 3731-3737.

9.10.3.29. Cikim AS, Oflaz H, Ozbey N et al. Evaluation of endothelial function in subclinical hypothyroidism and subclinical hyperthyroidism. Thyroid 2004; 14: 605-609.

9.10.3.30. Taddei S, Caraccio N, Virdis A et al. Low-grade systemic inflammation causes endothelial dysfunction in patients with Hashimoto’s thyroiditis. J Clin Endocrinol Metab 2006; 91: 5076-5082.

9.10.3.31. Monzani F, Caraccio N, Kozakowa M et al. Effect of levothyroxine replacement on lipid profile and intima-media thickness in subclinical hypothyroidism: a double-blind, placebo-controlled study. J Clin Endocrinol Metab 2004; 89: 2009-2106.

9.10.3.32. Nagasaki T, Inaba M, Henmi Y et al. Decrease in carotid intima-media thickness in hypothyroid patients after normalization of thyroid function. Clin Endocrinol 2003; 59: 607-612.

9.10.3.33. Tunbridge WM, Evered DC, Hall R et al. Lipid profiles and cardiovascular disease in the Whickham area with particular reference to thyroid failure. Clin Endocrinol 1977; 7: 495-508.

9.10.3.34. Hueston WJ, Pearson WS. Subclinical hypothyroidism and the risk of hypercholesterolemia. Ann Fam Med 2004; 2: 351-355.

9.10.3.35. Walsh JP, Bremmer AP, Bulsara MK et al. Thyroid dysfunction and serum lipids: a community-based study. Clin Endocrinol 2005; 63: 670-675.

9.10.3.36. Kanaya AM, Harris F, Volpato S, Perez-Stable EJ, Harris T, Bauer DC. Association between thyroid dysfunction and total cholesterol level in an older biracial population: the health, aging and body composition study. Arch Int Med 2002; 162: 773-779.

9.10.3.37. Bindels AJ, Westendorp RG, Frolich M, Seidell JC, Blokstra A, Smelt AH. The prevalence of subclinical hypothyroidism at different total plasma cholesterol levels in middle aged men and women: a need for case-finding? Clin Endocrinol 1999; 50: 217-220.

9.10.3.38. Bauer DC, Ettinger B, Browner WS. Thyroid function and serum lipids in older women: a population-based study. Am J Med 1998; 104: 546-551.

9.10.3.39. Torun AN, Kulaksizoglu S,Kulaksizoglu M, Pamuk BO, Isbilen E, Tutuncu NB. Serum total antioxidant status and lipid peroxidation marker malondialdehyde levels in overt and subclinical hypothyroidism. Clin Endocrinol 2009; 70: 469-474.

9.10.3.40. Brenta G, Berg G, Zago V et al. Proatherogenic mechanisms in subclinical hypothyroidism: hepatic lipase activity in relation to the VLDL remnant IDL. Thyroid 2008; 18: 1233-1236.

9.10.3.41. Bakker SJ, ter Maaten JC, Popp-Snijders C, Slaets JP, Heine RJ, Gans RO. The relationship between thyrotropin and low density lipoprotein cholesterol is modified by insulin sensitivity in healthy euthyroid subjects. J Clin Endocrinol Metab 2001; 86: 1206-1211.

9.10.3.42. Muller B, Zulewski H, Huber P, Ratcliffe JG, Staub JJ. Impaired action of thyroid hormone associated with smoking in women with hypothyroidism. N Engl J Med 1995; 333: 964-969.

9.10.3.43. Tuzcu A, Bahceci M, Gokalp D, Tuzun Y, Gunes K. Subclinical hypothyroidism may be associated with elevated highsensitive C-reactive protein (low grade inflammation) and fasting hyperinsulinemia. Endocr J 2005; 52: 89-94.

9.10.3.44. Maratou E, Hadjidakis DJ, Kollias A et al. Studies of insulin resistance in patients with clinical and subclinical hypothyroidism. Eur J Endocrinol 2009; 160: 785-790.

9.10.3.45. Christ-Crain M, Meier C, Guglielmetti M et al. Eleveated C-reactive protein and homocysteine values: cardiovascular risk factors in hypothyroidism? A cross-sectional and a double-blind, placebo-controlled trial. Atherosclerosis 2003; 166: 379-386.

9.10.3.46. Kvetny J, Heldgaard PE, Bladbjerg EM, Gram J. Subclinical hypothyroidism is associated with a low-grade inflammation, increased triglyceride levels and predicts cardiovascular disease in males below 50 years. Clin Endocrinol 2004; 61: 232-238.

9.10.3.47. Hueston WJ, King DE, Geesey ME. Serum biomarkers for cardiovascular inflammation in subclinical hypothyroidism. Clin Endocrinol 2005; 63: 582-587.

9.10.3.48. Ozcan O, Cakir E, Yaman H, et al. The effects of thyroxine replacement on the levels of serum asymmetric dimethylarginine (ADMA) and other biochemical cardiovascular risk markers in patients with subclinical hypothyroidism. Clin Endocrinol 2005; 63: 203-206.

9.10.3.49. Jorde R, Figenschau Y, Hansen JB. Haemostatic function in subjects with mild subclinical hypothyroidism. The Tromso study. Thromb Haemost 2006; 95: 750-751.

9.10.3.50. Goulis DG, Tsimpiris N, Delaroudis S et al.: Stapedial reflex: A biological index found to be abnormal in clinical and subclinical hypothyroidism. Thyroid 1998; 8: 583.

9.10.3.51. Misiunas A, Niepominiscze H, Ravera B et al. Peripheral neuropathy in subclinical hypothyroidsm. Thyroid 1995; 5: 283-286.

9.10.3.52. Monzani F, Caraccio N, Siciliano G et al.: Clinical and biochemical features of muscle dysfunction in subclinical hypothyroidism. J Clin Endocrinol Metab 1997; 82: 3315-3318.

9.10.3.53. CaraccioN, Natali A, Sironi A et al. Muscle metabolism and exercise tolerance in subclinical hypothyroidsm: a controlled trial of levothyroxine. J Clin Endocrinol Metab 2005; 90: 4057-4062.

9.10.3.54. Reuters VS, Teixeira PF, Vigario PS et al. Functional capacity and muscular abnormalities in subclinical hypothyroidism. Am J Med Sci 2009; 338: 259-263.

9.10.3.55. Simonsick EM, Newman AB, Ferrucci L et al. Subclinical hypothyroidism and functional mobility in older adults. Arch Int Med 2009; 169: 2011-2017.

9.10.3.56. Laukkarinen J, Kiudelis G, Lempinen M et al. Increased prevalence of subclinical hypothyroidism in common bile duct stones. J Clin Endocrinol Metab 2007; 92: 4260-4264.

9.10.3.57. Squizzato A, Romualdi E, Piantanida E et al. Subclinical hypothyroidism and deep venous thrombosis. Thromb Haemost 2007; 97: 803-806.

9.10.3.58. Nagata M, Suzuki A, Sekiguchi S et al. Subclinical hypothyroidism is related to lower heel QUS in postmenopausal women. Endocr J 2007; 54: 625-630.

9.10.3.59, Texeira PF, Cabral MD, Silva NA et al. Serum leptin in overt and subclinical hypothyroidism: effect of levothyroxine treatment and relationship to menopausal status and body composition. Thyroid 2009; 19: 443-450.

9.10.3.60. Tanda ML, Lombardi V, Genovesi M et al. Plasma total and acylated ghrelin concentrations in patients with clinical and subclinical thyroid dysfunction. J Endocrinol Invest 2009; 32: 74-78.

9.10.3.61. Akin F, Yaylali GF, Turgut S, Kaptanoglu B. Growth hormone/insulin-like growth factor axis in patients with subclinical thyroid dysfunction. Growth Horm IGF Res 2009; 19: 252-255.

9.10.3.62. Cooper DS, Biondi B. Subclinical thyroid disease. Lancet 2012; 379: 1142-1154.

9.10.3.63.de Jongh RT, Lips P, van Schoor NM, et al. Endogenous subclinical thyroid disorders, physical and cogniitive function, depression, and mortality in older individuals. Eur J Endocrinol 2011; 165: 545-554.

9.10.3.64. Lindeman RD, Schade DS, LaRue A, et al. Subclinical hypothyroidism in a biethnic, urban community. J Am Geriatr Soc 1999; 47: 703-709.

9.10.3.65. Lee JS, Buzkova P, Fink HA, et al. Subclinical thyroid dysfunction and incident hip fracture in older adults. Arch Intern Med 2012; 170: 1876-1883.

References for Section 9.10.4.

9.10.4.1. Bastenie PA, Vanhaelst L, Bonnyns M et al.: Preclinical hypothyroidism: a risk factor for coronary heart disease. Lancet 1971; i: 203-204

9.10.4.2. Bastenie PA, Vanhaelst L, Goldstein J et al.: Asymptomatic autoimmune thyroiditis and coronary heart disease. Cross-sectional and prospective studies. Lancet 1977; ii: 155-158.

9.10.4.3. Vanderpump MPJ, Tunbridge WMG, French JM, et al.: The development of ischemic heart disease in relation to autoimmune thyroid disease in a 20-year follow-up study of an English community. Thyroid 1996; 6: 155-160.

9.10.4.4. Hak AE, Pols HA, Visser TJ, Drexhage HA, Hofman A, Witteman JC. Subclinical hypothyroidism is an independent risk factor for atheroslerosis and myocardial infarction in elderly women: the Rotterdam study. Ann Intern Med 2000; 132: 270-278.

9.10.4.5. Parle JV, Maisonneuve P, Sheppard MC, Boyle P, Franklyn JA. Prediction of all-cause and cardiovascular mortality in elderly people from one low serum thyrotropin result: a 10-year cohort study. Lancet 2001; 358: 861-865.

9.10.4.6. Imaizumi M, Akahoshi M, Ichimaru S et al. Risk for ischemic heart disease and all-cause mortality in subclinical hypothyroidism. J Clin Endocrinol Metab 2004; 89: 3365-3370.

9.10.4.7. Gussekloo J, van Exel E, de Craen AJ, Meinders AE, Frolich M, Westendorp RG. Thyroid status, disability and cognitive function, and survival in old age. JAMA 2004; 292: 2591-2599.

9.10.4.8. Walsh JP, Bremner AP, Bulsara MK et al. Subclinical thyroid dysfunction as a risk factor for cardiovascular disease. Arch Intern Med 2005; 165: 2467-2472.

9.10.4.9. Rodondi N, Newman AB, Vittinghoff E et al. Subclinical hypothyroidism and the risk of heart failure, other cardiovascular events, and death. Arch Intern Med 2005; 165: 2460-2466.

9.10.4.10. Cappola AR, Fried LP, Arnold AM et al. Thyroid status, cardiovascular risk, and mortality in older adults: the cardiovascular health study. JAMA 2006; 295: 1033-1041.

9.10.4.11. Iervasi G, Molinaro S, Landi P et al. Association between increased mortality and mild thyroid dysfunction in cardiac patients. Arch Intern Med 2007; 167: 1526-1532.

9.10.4.12. Iqbal A, Schimer H, Lunde P, Figenschau Y, Rasmussen K, Jorde R. Thyroid stimulating hormone and left ventricular function. J Clin Endocrinol Metab 2007; 92: 3504-3510.

9.10.4.13. Rodondi N, Bauer DC, Cappola AR et al. Subclinical thyroid dysfunction, and the risk of heart failure. The Cardiovascular Health Study. J Am Coll Cardiol 2008; 52: 1152-1159.

9.10.4.14. Boekholdt SM, Titan SM, Wiersinga WM et al. Initial thyroid status and cardiovascular risk factors; the EPIC-Norfolk propsective population study. Clin Endocrinol 2010; 72: 404-410.

9.10.4.15. Sgarbi JA, Matsumura LK, Kasamatsu TS, Ferreira SR, Maciel RM. Subclinical thyroid dysfunctions are independent risk factors for mortality in a 7.5-year follow-up: the Japanese-Brazilian thyroid study. Eur J Endocrinol 2010; 162: 569-577.

9.10.4.16. Ravzi S, Weaver JU, Vanderpump MP, Pearce SH. The incidence of ischemic heart disease and mortality in people with subclinical hypothyroidism: reanalysis of the Whickham Survey cohort. J Clin Endocrinol Metab 2010; 95: 1734-1740.

9.10.4.17. Rodondi N, Aujesky D, Vittinghoff, Cornuz J, Bauer DC. Subclinical hypothyroidism and the risk of coronary heart disease: a meta-analysis. Am J Med 2006; 119: 541-551.

9.10.4.18. Volzke H, Schwahn, Wallaschoski H, Dorr M. Review: The association of thyroid dysfunction with all-cause and circulatory mortality: is there a causal relationship? J Clin Endocrinol Metab 2007; 92: 2421-2429.

9.10.4.19. Ochs N, Auer R, Bauer DC et al. Meta-analysis: subclinical thyroid dysfunction and the risk for coronary heart disease and mortality. Ann Intern Med 2008; 148: 832-845.

9.10.4.20. Haentjens P, Van Meerhaeghe A, Poppe K, Velkeniers B. Subclinical thyroid dysfunction and mortality: an estimate of relative and absolute excess all-cause mortality based on time-to-event data from cohort studies. Eur J Endocrinol 2008; 159: 329-341.

9.10.4.21. Singh S, Duggal J, Molnar J, Maldonado F, Barsano CP, Arora R. Impact of subclinical disorders on coronary heart disease, cardiovasdcular and all-cause mortality: a meta-analysis. Int J Cardiol 2008; 125: 41-48.

9.10.4.22. Ravzi S, Shakoor A, Vanderpump M, Weaver JU, Pearce SHS. The influence of age on the relationship between subclinical hypothyroidism and ischemic heart disease: a metaanlysis. J Clin Endocrinol Metab 2008; 93: 2998-3007.

9.10.4.23. Rodondi N, Den Elzen W, Bauer DC et al. Subclinical hypothyroidism and the risk of coronary disease and mortality. JAMA 2010; 304: 1365-1374.

9.10.4.24. McQuade C, Skugor M, Brennan DM, Hoar B, Stevenson C, Hoogwerf BJ. Hypothyroidism and moderate subclinical hypothyroidism are associated with increased all-cause mortality independent of coronary heart disease risk factors: a PreCIS database study. Thyroid 2011; 21: 837-843.

9.10.4.25. Nanchen D, Gussekloo J, Westerdorp RG, et al. Subclinical thyroid dysfunction and the risk of heart failure in older persons at high cardiovascular risk. J Clin Endocrinol Metab 2012; 97: 852-861.

9.10.4.26. Waring AC, Harrison S, Samuels MH, et al. Thyroid function and mortality in older men: a prospective study. J Clin Endocrinol Metab 2012; 97: 862-870.

9.10.4.27. Tseng FY, Lin WY, Lin CC, et al. Subclinical hypothyroidism is associated with increased risk for all-cause and cardiovascular mortality in adults. J Am Coll Cardiol 2012; 60: 730-737.

9.10.4.28. Waring AC, Arnold AM, Newman AB, Buzkova P, Hirsch C, Cappola AR. Longitudinal changes in the oldest old and survival: the cardiovascular health study all-stars study. J Clin Endocrinol Metab 2012; 97: 3944-3950.

9.10.4.29. Hyland KA, Arnold AM, Lee Js, Cappola AR. Persisitent subclinical hypothyroidism and cardiovascular risk in the elderly: the cardiovascular health study. J Clin Endocrinol Metab 2013; 98: 533-540.

9.10.4.30. LeGrys VA, Funk MJ, Lorenz CE, et al. Subclincial hypothyroidism and risk for incident myocardial infarction among postmenopausal women. J Clin Endocrinol Metan 3013; 98: 2308-2317. 9.10.4.31. Rhee CM, Curhan GC, Alexander EK, Bhan I, Brunelli SM. Subclinical hypothyroidism and survival: the effects of heart failure and race. J Clin Endocrinol Metab 2013; 98: 2326-2336.

9.10.4.32. Gencer B, Collet TH, Virgini V, et al. Subclinical thyroid dysfunction and the risk of heart failure events: an individual participant data analysis from 6 prospective cohorts. Circulation 2012; 126: 1040-1049.

References for Section 9.10.5

9.10.5.1. Villar HC, Saconato H, Valente O, Atallah AN. Thyroid hormone replacement for subclinical hypothyroidism. Cochrane Database Syst Rev 2007; issue 3: CD003419.

9.10.5.2. Parle J, Roberts L, Wilson S et al. A randomized controlled trial of the effect of thyroxine replacement on cognitive function in community-living elderly subjects with subclinical hypothyroidism: the Birmingham Elderly Thyroid Study. J Clin Endocrinol Metab 2010; 95: May 25 (Epub ahead of print).

9.10.5.3. Baldini M, Colasanti A, Orsatti A, Airaghi L, Mauri MC, Cappellini MD. Neuropsychological functions and metabolic aspects in subclinical hypothyroidism: the effects of L-thyroxine. Prog Neuropsychopharmacol Biol Psychiatry 2009; 33: 854-859.

9.10.5.4. Correia N, Mullally S, Cooke G et al. Evidence for a specific defect in hippocampal memory in overt and subclinical hypothyroidism. J Clin Endocrinol Metab 2009; 94: 3789-3797.

9.10.5.5. Biondi B, Palmieri EA, Lombardi G, Fazio S. Effects of subclinical thyroid dysfunction on the heart. Ann Int Med 137:904-914, 2002.

9.10.5.6. Faber J, Petersen L, Wiinberg N et al. Hemodynamic changes after levothyroxine treatment in subclinical hypothyroidism. Thyroid 2002; 12: 319-324.

9.10.5.7. Adrees M, Gibney J, El-Saeity N, Boran G. Effects of L-T4 replacement in women with subclinical hypothyroidism. Clin Endocrinol 2009; 71: 298-303.

9.10.5.8. Nagasaki T, Inaba M, Yamada S et al. Decrease of brachial-ankle pulse wave velocity in female subclinical hypothyroid patients during normalization of thyroid function: a double-blind, placebo-controlled study. Eur J Endocrinol 2009; 160: 409-415.

9.10.5.9. Owen PJ, Rajiv C, Vinereanu D, Mathew T, Fraser AG, Lazarus JH. Subclinical hypothyroidism, arterial stiffness, and myocardial reserve. J Clin Endocrinol Metab 2006; 91: 2126-2132.

9.10.5.10. Peleg RK, Efrati S, Benbassat C, Fygenzo M, Golik A. The effect of arterial stiffness and lipid profile in patients with subclinical hypothyroidism. Thyroid 2008; 18: 825-830.

9.10.5.11. Shakoor SK, Aldibbiat A, Ingoe LE et al. Endothelial progenitor cells in subclinical hypothyroidism: the effect of thyroid hormone replacement therapy. J Clin Endocrinol Metab 2010; 95: 319-322.

9.10.5.12. Duman D, Demirtunc R, Sahin S, Esertas K. The effects of simvastatin and levothyroxine on intima-media thickness of carotid artery in female normolipemic patients with subcliniccal hypothyroidism: a prospective, randomized-controlled study. J Cardiovasc Med 2007; 8: 1007-1011.

9.10.5.13. Kim SK, Kim SH, Park KS, Park SW, Cho YW. Regression of increased common carotid artery-intima media thickness in subclinical hypothyroidism after thyroid hormone replacement. Endocr J 2009; 56: 753-758.

9.10.5.14. Kebapcilair L, Comlekci A, Tuncel P et al. Effect of levothyroxine replacement therapy on paraoxonase-1 and carotid intima-media thickness in subclinical hypothyroidism. Med Sci Monit 2010; 16: CR41-47.

9.10.5.15. Teixeira PF, Reuters VS, Ferreira MM et al. Treatment of subclinical hypothyroidism reduces atherogenic lipid levels in a placebo-controlled double-blind trial. Horm Metab Res 2008; 40: 50-55.

9.10.5.16. Tanis BC, Westendorp RGJ, Smelt AHM: Effect of thyroid substitution on hypercholesterolaemia in patients with subclinical hypothyroidism: a reanalysis of intervention studies. Clin Endocrinol 1996; 44: 643-649.

9.10.5.17. Danese MD, Ladenson PW, Meinert CL, Powe NR. Effect of thyroxine therapy on serum lipoproteins in patients with mild thyroid failure: a quantitative review of the literature. J Clin Endocrinol Metab 2000; 85: 2993-3001.

9.10.5.18. Duntas LH, Wartofsky L. Cardiovascular risk factors and subclinical hypothyroidism: focus on lipids and new emerging risk factors. What is the evidence? Thyroid 2007; 17: 1075-1084.

9.10.5.19. Meier C, Staub J-J, Roth C-B, et al. TSH-controlled L-thyroxine therapy reduces cholesterol levels and clinical symptoms in subclinical hypothyroidism: a double-blind, placebo-controlled trial (Basel Thyroid Study). J Clin Endocrinol Metab 2001; 86: 4860-4866

9.10.5.20. Mainenti MR, Vigario PS, Teixeira PF, Maia MD, Oliveira FP, Vaisman M. Effect of levothyroxine replacement on exercise performance in subclinical hypothyroidism. J Endocrinol Invest 2009; 32: 470-473.

9.10.5.21. Simonsick EM, Newman AB, Ferrucci L et al. Health ABC study. Subclinical hypothyroidism and functional mobility in older adults. Arch Intern Med 2009; 169: 2011-2017.

9.10.5.22. Cinemre H, Bilir C, Gokosmanoglu F, Bahcebasi T. Hematologic effects of levothyroxine in iron-deficient subclinical hypothyroid patients: a randomized, double-blind, controlled study. J Clin Endocrinol Metab 2009; 94: 151-156.

9.10.5.23. Abalovich M, Amino N, Barbour LA et al. Thyroid dysfunction during pregnancy and and postpartum: an Endocrine Society clinical practice guideline. J Clin Endocrinol Metab 2007; 92: Supplement S1-S47.

9.10.5.24. Rahman AH, Abbassy HA, Elatif Abbassy AA. Improved IVF outcomes after treatment of subclinical hypothyroidism in infertle women. Endocr Pract 2010; 29: 1-17.

9.10.5.25. Dermott MT, Ridgway EC. Subclinical hypothyroidism is mild thyroid failure and should be treated. J Clin Endocrinol Metab 2001; 86: 4585-4590.

9.10.5.26. Owen PJD, Lazarus JH. Subclinical hypothyroidism: the case for treatment. Trends Endocrinol Metab 2003; 14: 257-261.

9.10.5.27. Chu JW, Crapo LM. The treatment of subclinical hypothyroidism is seldom necessary. J Clin Endocrinol Metab 2001; 86: 4591-4599.

9.10.5.28. Vanderpump M. Subclinical hypothyroidism: the case against treatment. Trends Endocrinol Metab 2003; 14: 262-266.

9.10.5.29. Surks MI, Ortiz E, Daniels GH et al. Subclinical thyroid disease. Scientific review and guidelines for diagnosis and management. JAMA 2004; 291: 228-238.

9.10.5.30. Ladenson PW. Cardiovascular consequences of subclinical thyroid dysfunction: more smoke but no fire. Ann Intern Med 2008; 148; 880-881.

9.10.5.31. Klubo-Gwiezdzinska J, Wartofsky L. Thyrotropin blood levels, subclinical hypothyroidism, and the elderly patient. Arch Intern Med 2009; 169: 1949-1951.

9.10.5.32.. Cooper DS. Subclinical hyperthyroidism. N Engl J Med 2001; 345: 260-265.

9.10.5.33. Gussekloo J, van Exel E, de Craen AJ, Meinders AE, Frolich M, Westendorp RG. Thyroid status, disability and cognitive function, and survival in old age. JAMA 2004; 292: 2591-2599.

9.10.5.34. Razvi S, Weaver JU, Butler TJ, Pearce SH. Levothyroxine treatment of subclinical hypothyroidism, fatal and nonfatal cardiovascular events, and mortality. Arch Intern Med 2012; 172: 811-817.

9.10.5.35. Ravzi S, Ingoe L, Keeka G, Outes C, McMillan C, Weaver JU. The beneficial effect of L-thyroxine on cardiovascular risk factors, endothelial function, and quality of life in subclinical hypothyroidism: randomized, crossover trial. J Clin Endocrinol Metab 2007; 92: 1715-1723.

9.10.5.36 Garber JR, Cobin RH, Gharib H, et al: Clinical practice guidelines for hypothyroidism in adults: cosponsored by the American Association of Clinical Endocrinologists and the American Thyroid association. Thyroid 2012; 22:1200-1235.

9.10.5.37 Pearce SHS, Brabant G, Duntas LH, et al: 2013 ETA guideline: Management of subclinical hypothyroidism. Eur Thyroid , 2013; 2:

References for Section 9.11.

9.11.1. Wang C, Crapo LM: The epidemiology of thyroid disease and implication for screening. Endocrinol Metab Clin North Am 1997; 26: 189-218.

9.11.2. Kågedal B, Månson JC, Norr A, et al.: Screening for thyroid disorders in middle-age women by computer-assisted evaluation of a thyroid hormone panel. Scand J Clin Lab Invest 1981; 41: 403-408.

9.11.3. Meier Ch, Staub JJ, Roth CB, et al.: Screening for new thyroid dysfunction in women attending a primary care unit using a third generation TSH assay. Schweiz Med Wochenschr 1998; 128: 128: 250-253

9.11.4. Helfand M, Capro LM: Screening for thyroid disease. Ann Int Med 1990; 112: 840-849.

9.11.5. Danese MD, Powe NR, Sawin CT, Ladenson PW: Screening for mild thyroid failure at the periodic health examination: A decision and cost-effectiveness analysis. JAMA 1996; 276: 285-292.

9.11.6. Powe NR, Danese MD, Ladenson PW: Decision analysis in endocrinology and metabolism. Endocrinol Metab Clin North Am 1997; 26: 89-112.

9.11.7. Helfand M, Redfern CC: Screening for thyroid disease: an update. Ann Intern Med 1998; 129: 144-158

9.11.8. Weetman AP: Hypothyroidism: screening and subclinical disease. Brit Med J 1997; 314: 1175-1178.

9.11.9. Vanderpump MPJ, Ahlquist JAO, Franklyn JA, Clayton RN: Consensus statement for good practice and audit measures in the management of hypothyroidism and hyperthyroidism. BMJ 1996; 313: 539-544.

9.11.10. Oomen HAPC, Schipperijn AMJ, Drexhage HA: The prevalence of affective disorder and in particular of a rapid cycling of bipolar disorder in patients with abnormal thyroid function tests. Clin Endocrinol 1996; 45: 215-223.

9.11.11. Giani C, Fierabracci P, Bonacci R, et al.: Relationship between breast cancer and thyroid disease: relevance of autoimmune thyroid disorders in breast malignancy. J Clin Endocrinol Metab 1996; 81: 990-994.

9.11.12. Negro R, Schwartz A, Gismondi R, Tinelli A, Mangieri T, Stagnaro-Green A. Universal sreening versus case finding for detection and treatment of thyroid hormonal dysfunction during pregnancy. J Clin Endocrinol Metab 2010; 95: 1699-1707.

Insulin Biosynthesis, Secretion, Structure, and Structure-Activity Relationships

TAKE-HOME POINTS

  1. The structure of insulin contains determinants of foldability, trafficking, self-assembly, and receptor binding.
  2. Insulin is the biosynthetic product of a single-chain precursor, preproinsulin, whose proteolytic processing is coupled to trafficking between cellular compartments.
  3. The connecting (C) domain of proinsulin is removed by a specialized set of endoproteases and a carboxypeptidase activity, acting mainly within maturing secretory granules.
  4. Insulin is stored as microcrystalline arrays of zinc insulin hexamers within specialized glucose-regulated secretory vescicles.
  5. Regulation of insulin secretion is coupled to metabolism and electrophysiologic events involving plasma membrane depolarization and calcium-ion homeostasis.
  6. The insulin receptor is a transmembrane protein containing an extracellular hormone-binding domain and intracellular tyrosine kinase domain.
  7. Binding of insulin to the insulin receptor (an ( αβ ) 2 dimer) is mediated by side chains in both the A- and B chains of insulin.
  8. The primary hormone-binding site (Site 1) in the extracellular domain of the insulin receptor contains the L1 β -helix of one α -subunit and the C-terminal α CT α -helix of the other α -subunit.
  9. Dominant mutations in the insulin gene cause monogenic syndromes of diabetes mellitus, prominently including permanent neonatal-onset diabetes, due to toxic misfolding of proinsulin variants.

INTRODUCTION

Insulin plays a central role in the regulation of human metabolism. The hormone is a 51-residue anabolic protein that is secreted by the β-cells in the Islets of Langerhans. Containing two chains (A and B) connected by disulfide bonds, the mature hormone is the post-translational product of a single-chain precursor, designated proinsulin . Extensive studies of the three-dimensional structure of insulin, pioneered by D. C. Hodgkin , have enabled the development of therapeutic analogs for the treatment of the metabolic disorder diabetes mellitus (DM) . The insulin gene is the site of dominant mutations associated with DM . Although such mutations are uncommon, their molecular analysis has provided important insights into the biochemical bases of the hormone’s pathway of biosynthesis and mechanism of receptor binding . The largest class of mutations is associated with the impaired folding of proinsulin, which leads in turn to progressive endoplasmic-reticular (ER) stress, β-cell death and DM, usually with onset in the neonatal period .

Key complementary functions of insulin are (a) stimulation of glucose uptake from the systemic circulation and (b) suppression of hepatic gluconeogenesis, together regulating glucose homeostasis . DM is characterized by decreased glucose tolerance resulting from a relative deficiency of insulin or a lack of sensitivity to the endogenous hormone. Insufficient insulin, or decreased insulin sensitivity, results in hyperglycemia. Long-term exposure of tissues to elevated ambient glucose concentrations is associated with the development of complications, including macro- and microvascular disease. Of particular concern are coronary heart disease, cerebrovascular disease, and the characteristic retinopathy, nephropathy, and neuropathy of this disorder .

The history of the discovery of insulin and its therapeutic utility defined a paradigm for the integration of physiologic and biochemical approaches in experimental medicine . At the end of the 19 th century Von Mering and Minkowski noted that removal of the pancreas led to the development of DM in dogs . In 1916 Schafer first speculated that an antidiabetic hormone, which he named “insuline,” was secreted from pancreatic islets . Barron noted in 1920 that ligation of the pancreatic duct, with destruction of the exocrine pancreas, only resulted in DM if the islets, so named by Langerhans in 1869 , were also destroyed . Subsequently, the work of Banting, Best, Collip and MacCleod in the early 1920′s resulted in the identification of a substance in extracts of pancreas that had the remarkable ability to reduce blood glucose levels in diabetic animals . By 1923 these pancreatic extracts were employed to successfully treat patients with DM. The dramatic clinical utility of insulin encouraged broad public support for medical research .

The modern view of insulin is as a ligand that activates a specific cellular receptor, designated the insulin receptor (IR) . The IR belongs to a superfamily of receptor tyrosine kinase whose activation modulates multiple post-receptor signaling pathways . Insulin thus regulates a host of other cellular processes, such as protein and fat synthesis, RNA and DNA synthesis, as well as cell growth and differentiation . It is, however, the regulation of glucose uptake that is of primary concern in the clinical manifestations of diabetes; therefore, this chapter will begin with a brief description of how plasma glucose homeostasis is achieved.

REGULATION OF PLASMA GLUCOSE BY INSULIN

Specific membrane transporters facilitate the movement of glucose into cells to reduce plasma glucose concentrations in response to insulin stimulation. The transported glucose is subsequently used as metabolic fuel or stored as a complex polymeric structure, designated glycogen. Two major types of glucose transporters are known: Na + -dependent and Na + -independent. Only the Na + -independent transporters possess an insulin-responsive isoform. We describe each in turn.

The Na + -dependent glucose transporter family has been identified in several tissues, particularly in small intestinal epithelium (SGLT1) and the renal proximal tubule (SGLT2) as well as in other kidney tubule cells . These transporters are located on the lumenal side of intestinal and kidney cells and act to absorb glucose against its concentration gradient by coupling the movement of glucose into these cells with the concomitant movement of Na + into the cell. Since Na + is moving down its electrochemical gradient this energy can be used to cotransport glucose into the cells. Thus, this transporter is dependent on the concentrations of extracellular and intracellular sodium ions, which are maintained by a Na + /K + -ATPase ion pump.

The Na + -independent glucose transporter family, consisting of several isoforms, facilitates the movement of glucose down its concentration gradient across a plasma membrane. Although seven isoforms have been identified (designated Glut1-7) , only one will be discussed in detail here, Glut4, because it is the transporter that is in highest concentration in insulin-sensitive tissues, including skeletal muscle, cardiac muscle, and adipose tissue (fat) . Glut4, and to a lesser extent Glut1 enable these cells to increase their uptake of glucose, thereby lowering circulating concentrations. Because the intracellular concentration of glucose is low due to the rapid phosphorylation of glucose to glucose-6-phosphate (not a substrate of the Glut transporters) and its conversion to other metabolic products, the presence of active transporters in the plasma membrane favors the movement of glucose into cells.

Insulin enhances glucose uptake by increasing the number of transporters in the plasma membrane of target cells. This was first demonstrated in adipocytes and subsequently in skeletal and cardiac muscle . Insulin stimulation of such cells mobilizes transporters from intracellular compartments to the plasma membrane to facilitate glucose transport. Translocation of receptors to the plasma membrane has been demonstrated to occur within 30 seconds of insulin stimulation ; as the stimulus dissipates the decrease in the number of plasma membrane receptors declines coincident with a decline in glucose transport . Whereas glucose transport via Glut4 is a passive process (limited only by the chemical potential of the glucose gradient and the V max of the transporters), translocation and reverse reinternalization of receptors are energy-dependent processes . The impaired ability of insulin, on binding and activation of the IR, to signal Glut4 translocation from intracellular stores contributes to postprandial hyperglycemia in Type 2 DM . Animal studies have also demonstrated that insulin resistance is associated with a decreased translocation of glucose transporters to the plasma membrane in muscle cells . In fact, decreased insulin levels in animal models of DM have been shown not only to decrease transporter translocation, but also to attenuate expression of Glut4 in muscle cells. Thus, it appears that insulin provides both a short-term signal to increase glucose-transporter translocation and a long-term signal to maintain a basal level of expression of such transporters in target cells. The combination of acute and basal actions provides a common mechanism in Type 1 DM (characterized by low or vanishing endogenous insulin levels) or Type 2 DM (characterized by insulin resistance) could cause pathologically high plasma glucose levels: loss of regulation and expression of transmembrane glucose transporters. Glut2, expressed on surface of β-cells, contributes to the regulation of insulin secretion . Accordingly, a β-cell specific IR knock-out (KO) model indicated that insulin likely positively regulates its own secretion from the β-cell .

Since its purification and clinical application in 1923 , the central importance of insulin in regulating glucose metabolism and the prevention of DM has stimulated research in attempts to understand the mechanism of action of this peptide hormone. This work has lead to the determination of the three-dimensional structure of insulin , identification of its precursor and the processing and secretion mechanisms that underlie its production . Complementary advances have seen the identification of the IR and of its mechanisms of signal transduction . Recent studies have defined three-dimensional structures of proinsulin and elements of the IR involved in hormone binding . In addition to their fundamental importance, these discoveries have deepened our understanding of the molecular basis of DM and its treatment as discussed below.

INSULIN BIOGENESIS AND MECHANISM OF RELEASE

Insulin was the first peptide hormone discovered. Before Abel crystallized insulin in 1926 and Jensen and Evans in 1935 identified the N-terminal phenylalanine of the B-chain , proving that insulin was indeed a protein, all hormones were believed to be small molecules. With the elucidation of the amino-acid sequence of insulin by Sanger in the mid 1950′s (see Figure 1), it became known that insulin was a two-chain heterodimer consisting of a 21-residue A-chain linked to a 30-residue B chain by two disulfide bonds derived from cysteine residues (A7-B7 and A20-B19). An intrachain disulfide bond also exists within the A-chain (A6-A11).

Figure 1 . Primary structures of porcine insulin and porcine proinsulin. The primary sequence of porcine insulin (a) as determined by Sanger and co-workers ; and proinsulin . The sequence of human insulin is identical to that of porcine insulin except for the change of Ala B30 to Thr B30 in human insulin.

Although this primary structure provided valuable information regarding the amino-acid composition and size (ca. 6,000 Daltons) of the insulin molecule , questions concerning the processes of insulin biosynthesis and secretion were not resolved until the late 1960′s with the discovery of proinsulin . This precursor protein (ca. 9,000 D) contains both the A- and B-chain of insulin in a continuous single chain joined through an intervening segment, designated the C domain . The C domain varies in length among vertebrate species (typically 30-35 residues) and is flanked at each end by dibasic residues (Arg-Arg and Lys-Arg) . Proinsulin is cleaved at those dibasic links by a trypsin-like enzyme to release the mature hormone and a free C-peptide (which lacks the dibasic residues).

Figure 2. (a) Diagrammatic illustration of the processing of insulin. The transcription and translation of the human insulin gene, as well as processing of preproinsulin to insulin is illustrated. (b) Conversion of proinsulin to insulin demonstrating the secondary and tertiary structures of insulin; the flexible C domain in proinsulin and split proinsulins is shown in dotted line. Thick blue arrows indicated predominant path.

Chan et al. subsequently discovered that there was an additional and larger precursor of insulin, preproinsulin. This single-chain polypeptide (ca. 12,000 Daltons) consists of proinsulin extended at the amino-terminus by a 24-residue signal peptide of hydrophobic residues . Such a signal sequence is characteristic of proteins that enter the secretory pathway . The signal peptide of preproinsulin is cleaved coincident with its translocation into the ER , and thus proinsulin itself (together with the oxidative folding machinery of the ER) mediates proper disulfide pairing and three-dimensional protein folding .

The steps involved in the conversion of the information encoded within the 1500 bases of the human insulin gene (as sequenced by Bell and colleagues in 1980 ) into preproinsulin and its subsequent proteolytic conversion to insulin are illustrated in Figure 2A and 2B. The first set of steps occur at the nucleic-acid level: the initial mRNA transcript is modified via excision of the two intervening sequences, the 5′ terminus is capped by 7-methyl guanosine, and the 3’-terminus undergoes polyadenylation to produce a mature mRNA product. This mRNA product encodes preproinsulin, which is translated on the rough endoplasmic reticulum (rER) and subsequently translocated into the RER lumen via a series of interactions of the signal peptide with the signal recognition particle (SRP) and SRP-receptor in the rER membrane . The next set of steps occurs at the protein level. The signal peptide is cleaved in the lumen of the rER by a signal peptidase (located on the lumenal side of the rER membrane). Within the cisternae of the rER, proinsulin undergoes rapid folding and disulfide bond formation to generate the native tertiary structure, the direct precursor of insulin. The signal peptide is rapidly degraded in the rER and is therefore not a normal secretory product of the β-cells .

In the final series of steps proinsulin is transported to the Golgi apparatus where it is packaged into secretory granules and converted to native insulin and C-peptide. The conversion process may begin in the trans Golgi network but continues in the condensing vacuoles (early secretory granules), and the products are stored in mature secretory vesicles, and secreted in equimolar amounts along with small amounts (ca. 2-3%) of proinsulin and intermediate cleavage products . Glucose, in addition to stimulating insulin secretion by β-cells, also activates insulin gene transcription, enhances insulin mRNA stability, and stimulates its translation .

Proinsulin, despite its larger size, shares many of the physical properties of insulin. Proinsulin has been shown, for example, to form dimers and Zn 2+ -coordinated hexamers in a manner similar to insulin , has a comparable isoelectric point and solubility , and cross-reacts with insulin antisera . These findings motivated the hypothesis that the structure of the insulin moiety in proinsulin is similar, if not identical, to that of native insulin . Although the crystal structure of proinsulin has not been determined , presumably due to the flexibility of the C domain, its solution structure has recently been determined by multidimensional NMR methods . The insulin moiety indeed retains the conformation of insulin whereas the C domain is flexible (but not completely disordered). The solution structure rationalizes why proinsulin is a full agonist of insulin and displays 3-5% biological activity ; the binding regions of the insulin moiety are accessible even in the presence of the C domain .

In 1969 it was demonstrated via pulse-chase studies that proteolytic processing of proinsulin occurs in the Golgi apparatus (GA) and/or early secretory vesicles of the β-cells ; and subsequent studies have identified the trans Golgi as the initial compartment wherein proinsulin and its converting enzymes are brought together to form secretory granules . Monoclonal antibodies specific for intact proinsulin also demonstrated that proinsulin is transferred from the rER to the cis- and trans Golgi, where the precursor is concentrated to form prosecretory vesicles . Several studies have demonstrated that shuttling of proinsulin vesicles from the rER to the cis Golgi and through the GA is an energy-requiring process requiring ATP . Subsequent conversion of proinsulin to insulin, initiated in the trans Golgi, accelerates within prosecretory granules as they acidify and mature in the cytosol over a period of 1-3 hours in preparation for secretion. Residual proinsulin and intermediate cleavage products then comprise only 2-3 percent of total stored insulin-related protein. Insulin secretory granules turn over at a much slower rate of many hours or several days normally. In the intracellular pathway taken by proinsulin on budding from the rER, distinct times are required for individual stages of transfer. In rat islets the conversion of proinsulin to insulin begins about 30 min after ribosomal synthesis of preproinsulin and resembles first-order reaction kinetics with half-times of approximately 30-60 minutes .

The conversion of proinsulin to insulin occurs through the joint action of two types of proteases: one with trypsin-like endoprotease activity which cleave after the dibasic residues pairs at each end of the C domain, and another with exopeptidase activity resembling that of carboxypeptidase B to remove the basic residues left after tryptic-like cleavage . Previous studies have also demonstrated that mixtures of pancreatic trypsin and carboxypeptidase B could convert proinsulin to insulin in vitro . Two endoproteases were found within insulinoma secretory granules . Initially called Type I and Type II converting enzymes, each of these acidic endoproteases was found to be dependent on Ca 2+ ions. Type I was active in 1 mM Ca 2+ and cleaved at Arg31-Arg32 in proinsulin (the first two positions of the C domain) whereas Type II required 0.1 mM Ca 2+ and cleaved predominantly at Lys64-Arg65 (the last two positions of the C domain) . Each was also found to have an acidic pH optimum near 6.0.

The discovery in yeast Saccharomyces cerevisciae of an endoprotease, designated Kex2, that cleaves at dibasic residue sites in the yeast α -mating factor and in microbial toxin precursor polypeptides facilitated the search for homologous endoproteases in mammalian β-cells . Kex2 is a homologue of subtilisin, a bacterial serine protease. In yeast this integral membrane protein is localized in the trans Golgi network. Analysis of human insulinoma cDNA via PCR techniques lead to the discovery of PC2, an enzyme having an homologous catalytic domain . PC2 shares 49% amino-acid identity with Kex2, but importantly lacks its transmembrane (TM) segment, indicating that it was likely to be a soluble protease and thus was a candidate for one of the processing endoproteases in β-cell secretory vesicles. Similar screening methods led to the discovery of PC1/3, the second secretory-granule convertase . Subsequent work demonstrated that both PC2 and PC1/3 display optimal activity at pH 5.5, consistent with the internal pH of β-cell secretory granules and are also expressed in the brain and many other neuroendocrine cells. These two proprotein convertases are members of a larger 7-member family of kexin/subtilisin-like endoproteases, which normally function within the secretory pathway in most eukaryotic cells .

Immunocytochemical studies of pancreatic islets β-cells provided evidence for the presence of both PC2 and PC1/3 in β-cells pancreatic islets , and their ability to convert proinsulin to insulin was demonstrated by co-infection with vaccinia viruses expressing PC2 and PC1/3 in Cos7 cells expressing proinsulin . Whereas PC2 mediated cleavage only at the C-peptide/A-chain junction, PC1/3 cleaved at both dibasic sites with a preference for the B-chain/C-peptide site. Subsequent studies also established the identity of PC2 and PC1/3 with the calcium-dependent Type II and Type I insulinoma endoproteases, respectively, as discovered by Davidson and coworkers . The carboxypeptidase B-like exopeptidase, which removes COOH-terminal basic residues after cleavage by PC2 and PC3/PC1, was also found and is known as carboxypeptidase E . Although structurally homologous to pancreatic carboxypeptidases A and B , carboxypeptidase E has several unique features that differentiate it from other carboxypeptidases . In the maturing secretory granule of the B cell, PC2, PC1/3, and carboxypeptidase E work together to convert proinsulin to mature insulin and C-peptide. Studies have demonstrated that PC1/3 first cleaves proinsulin at the B chain-C peptide junction, generating an intermediate product that is preferentially cleaved by PC2 to yield insulin . Thus knockout of PC1/3 results in very high circulating proinsulin levels while PC2 nulls exhibit much lower levels .

The PC family of convertases are all synthesized as inactive precursors with a lengthy N-terminal propeptide that is autocatalytically cleaved at a tetrabasic cleavage site by the proenzyme as it passes from the slightly alkaline ER to the neutral and mildly acidic conditions of the Golgi apparatus and is then removed by a second cleavage within the propeptide which assists in its release and disposal, yielding the active full-length enzyme . In the case of PC1/3 the full length form is active , but undergoes C-terminal autocatalytic truncation to a smaller, more active form within the secretory granules . In the case of PC2 the removal of the propeptide is more complex and requires the presence of neuroendocrine protein 7B2 . Activation of PC2 thus does not occur in the Golgi but rather within the secretory granules where it also functions. The full length enzyme is the active form in the granules within the beta cells and other neuroendocrine tissues. Knockout of 7B2 thus provides a phenocopy of the PC2 null when on the same genetic background, or inbred mouse strain (for review, see Ref ).

Secretory granules in the β-cell undergo maturation in the cytosol. Electron-microscopic (EM) studies have demonstrated that mature granules have a dense crystalline-appearing core with a spacing similar to that of 2-Zn insulin crystals . Newly synthesized insulin is likely to forms crystals with zinc ions that are transported into the maturing secretory granules as demonstrated by a knockout of the zinc transporter (ZnT8) . The crystals reside in the dense core of the β-granules whereas the soluble C-peptide resides in the less dense or clear periphery of the granule . Proinsulin is also known to crystallize with insulin in small amounts, probably as mixed hexamers . Both proinsulin and insulin have the ability to bind zinc and form zinc-coordinated hexamers (two Zn 2+ axial atoms per hexameric unit); the side chain of histidine B10 coordinates axial Zn 2+ ions . The self-assembly and micro-crystallization of zinc-insulin hexamers may be regulated by compartmental pH. The secretory granules possess an intrinsic proton pump, which serves to lower the pH within the granule to pH 5.0-5.5: this is optimal for both prohormone processing and crystallization in vitro . The pH within the rER is less acidic, which promotes thiol-disulfide exchange and hence proinsulin folding with native disulfide pairing.

High ambient glucose concentration in the islets promotes insulin biosynthesis and is the primary regulator of secretion. Elevated glucose concentrations cause an increase in cAMP levels by a mechanism that does not appear to involve activation of adenylate cyclase . cAMP then exerts its effects via a mechanism involving protein kinase A (PKA), leading to the phosphorylation and activation of certain key proteins . Through this complex chain of events, glucose and cAMP (and possibly contributions from the rise in intracellular free calcium and IP3) rapidly increase translation and transcription of insulin mRNA . Insulin mRNA normally turns over slowly, with a half life of approximately 30 hours at normal or below normal levels. However, elevated ambient glucose concentrations increase the half life of insulin mRNA as much as threefold . Calcium dependent exocytosis of secretory granules is the main mechanism of secretion in both glucose-stimulated and basal states . Little or no direct secretion of proinsulin occurs from the rER to the plasma membrane by way of unregulated pathways . Other hormones and chemical substances also play an important role in the regulation of insulin secretion , including glucagon, which is secreted by α -cells in pancreatic islets as discussed in detail in another chapter; glucagon-like peptide (GLP-1: Ref. ); cholecystokinin ; and gastric inhibitory peptide all acting via specific receptors on the β-cell. Inhibitors of insulin secretion include catecholamines (adrenaline and noradrenaline) which interact with adrenergic receptors on the β-cell membrane , and somatostatin which is secreted by  -cells of the pancreatic islets . Amylin is also secreted by the β -cell although the regulatory mechanisms for amylin co-secretion are not well understood.

The integrated regulation of insulin secretion in pancreatic β -cells provides an opportunity to develop quantitative data-based computer models relating metabolic sensing to electrophysiological events and to intracellular Ca 2+ dynamics, ultimately leading to exocytosis of the secretory granules . Such analyses provides fundamental insight into the biochemical and biophysical determinants of cytoplasmic transport of organelles .

BIOCHEMISTRY AND ELECTROPHYSIOLOGY OF INSULIN SECRETION

Secretion of insulin from β -cells is not only an important step in the regulation of glucose homeostasis in healthy individuals, but has also been demonstrated to be impaired (for different reasons) in both Type 1 and Type 2 DM . In fact, in the prediabetic state of Type 1 DM as well as in various forms of Type 2 DM, abnormalities in insulin secretion are an integral component of the pathophysiology . In light of its extensive characterization the β-cell also serves as a model of the secretory process for other cell types.

Insulin is stored in large dense core vesicles (LDCV) and released by exocytosis as described above. Such release is a multistep process that consists of the transport of the secretory vesicles to the plasma membrane, then docking, priming, and finally fusion of the vesicle with the plasma membrane. This process is regulated cooperatively by nutrients, other hormones, and neurotransmitters in association with electrical depolarization of the β-cell and release of insulin. Only a small portion of the insulin stored in vesicles in the β-cell is released, however, even under maximum stimulation. This suggests that systemic insulin levels are regulated by secretion rather than by biosynthesis and is not ordinarily limited by the size of storage pools. However the mechanisms that regulate the directed transport of the insulin granules to the plasma membrane are also not well understood.

The best characterized mechanism of coupling glucose metabolism to insulin secretion resides in the electrical excitability of the β-cell. A large number of ion channels, pumps, and transporters contribute to intracellular calcium concentration, as well as other ions, to set the membrane potential (V m ) of the β-cell; this set point is near -70 mV when extracellular glucose is ~3 mM (see Figure 3). In 1968 Dean and Mathews demonstrated that β-cells are electrically excitable and that glucose controlled this excitability . They also showed that the action potentials of β-cells were increased by sulfonylureas. The use of patch-clamp techniques enabled the electrical activity of the β-cell to be studied in detail, and such studies by Ashcroft, Rorsman and others elucidated the key role of ATP-sensitive potassium (KATP) channels in the resting membrane potential of the β-cell as well as the importance of these channels in the mechanism of insulin secretion (for a recent perspective, see ).

Figure 3 . β -cell ion channels. KATP channel conductance predominates in the resting β-cell, maintaining the resting Vm near the electrochemical potential (EK) for K+ (~ -80 mV). These potassium channels belong to the inward rectifier (Kir) subfamily. They obtain their name since these channels conduct K + current into the cell more readily than to the outside of the cell (inward rectification, as a diode). KATP channels are, however, weak inward rectifiers because they pass a significant amount of current in the outward direction. At -70 mV open KATP channels carry a small amount of outward current, which maintains the hyperpolarized resting potential in the β-cells.

A myriad of biochemical and biophysical structure-function studies of recombinant Kir channels has led to a more complete understanding of these channels. The crystal structure of a bacterial Kir analog, the Streptomyces lividens KcSA channel, has been determined ; and the inner pore of a mammalian Kir has likewise been crystallized, and its structure determined . The structures revealed that Kir channels consist of four subunits: each folds into the membrane to define two transmembrane domains (M1 and M2) surrounding a pore loop (P). The four P-loops line the central ion-conducting pore with the M1 and M2 subunits providing outer supports (Figure 4).

Elevation of glucose concentration to >8-10 mM results in depolarization of the β-cell. Glucose is taken up into the β-cell by the GLUT2 transporter and metabolized via glucokinase, glycolysis and in mitochondria to generate ATP. This alters the ATP/ADP ratio, which causes closure of KATP channels and depolarization of the cell via the decreased K + permeability. ATP inhibits KATP channels and ADP opens them . Other nucleotides generated by glucose metabolism (Ap3A: diadenosine triphosphate, and Ap4A: diadenosine tetraphosphate) have been implicated as second messengers mediating the closure of KATP channels, but their significance remains uncertain. Mutations in either the Kir, or SUR1, can result in persistent activation, leading to neonatal hyperinsulinemia and hypoglycemia . It is also known that some Kir channels in β-cells are activated via G-protein coupled receptors as reviewed below.

Figure 4 . Kir channels. KATP channels are unique in the inward rectifier family because they require an auxiliary subunit, the sulfonylurea receptor (SUR1), to function. The SUR1 is a member of the ATP binding cassette (ABC) family of membrane proteins, which includes the cystic fibrosis transmembrane conductance regulator (CFTR) chloride channel among others . It was named due to its binding to iodinated glyburide but clearly it is not actually an sufonylurea receptor. KATP channels in the β-cell consists of Kir6.2 subunits surrounded by their accompanying SUR1 subunit (not shown).

Pioneering studies by Katz, Miledi, and Douglas first established intracellular Ca 2+ concentrations as a general coupling factor between membrane depolarization and vesicular exocytosis ; this mechanism operates in β-cells (for review, see ). Voltage dependent Ca 2+ channels (Cav) open on membrane depolarization as caused by Kir channel closure; and it is this Ca 2+ influx that leads in turn to exocytosis of the glucose-regulated secretory granules and insulin secretion.

Cav channels are classified one the basis of a low-voltage threshold (LV: activated at more negative potentials) or high-voltage threshold (HV: activated at relatively depolarized potentials). HV channels can be further divided into subclasses: L, N, P, Q, and R . Insulin secretion is inhibited by dihydropyridine-based calcium-channel blocking agents, which inhibit L-type Cav. Although activators of L-type Cav can stimulate insulin secretion, Cav1.3 KO mice exhibited perturbed islet function resulting in glucose intolerance; these mice were smaller than controls and yet maintained glucose-dependent insulin secretion . This phenotype is likely to be secondary either to upregulation of other Cav1.2 channels or to the existence of other mechanisms of insulin secretion unreliant on voltage gated channels as discussed next. It has been suggested that the neuronal type of Ca channels play a direct role in exocytosis .

Several hormones and neurotransmitters regulate insulin secretion in addition to the voltage-sensitive pathways. Molecules such as epinephrine, galanin, somatostatin, acetylcholine, and glucagon-like peptide (GLP) each contribute to the regulation of insulin secretion by binding to cognate receptors . Cholecystokinin (CCK receptor pathway) and acetylcholine (M3 receptor in β-cells) potentiate insulin secretion via phosphoinositide catabolism with the subsequent mobilization of intracellular calcium ions. These ligands bind to G-protein coupled receptors (GPCRs) that can activate phospholipase-C (PLC) . PLC-catalyzed hydrolysis of phosphatidylinositol 4,5-bisphosphate (PIP 2 ) produces inositol 1,4,5-triphosphate (IP 3 ) and diacylglycerol (DAG). Two families of Ca 2+ channels are present on the ER and rER: IP3 receptors and ryanodine receptors (RyR). Each is capable of causing the release of Ca 2+ ions stored in the ER. DAG concomitantly activates protein kinase (PKC). Other potentiators of insulin secretion, such as GLP-1, and glucose-dependent insulinotropic polypeptide (GIP), bind to their respective cognate GPCRs to activate adenyl cyclase and increase intracellular cAMP, which in turn activates protein kinase A (PKA). Stimulation of either PKC or PKA alters second-messenger systems in the β-cell and can chemically modify ion channels to influence insulin secretion. In fact, insulin exocytosis can be induced independently from Ca 2+ fluxes in in vitro studies through intracellular application of GTP or nucleotide analogs GppNHp and GTP  S.

Specific proteins are also likely to be involved in the interaction of secretory vesicles with the plasma membrane. Pairing proteins on the vesicle membrane (v-SNARES) have been found to tightly interact with cognate proteins on the target membrane (t-SNARES). Cytosolic cofactors such as N-ethylmaleimide-sensitive factor (NSF) and α /β-SNAP may assist with ATP binding and hydrolysis to cause exocytosis .

In addition to voltage-sensitive pathways and GPCRs, the monomeric G-proteins (such as Rab27 and Rab3 ) may activate second-messenger cascades resulting in exocytosis; and multiple Ca 2+ -binding proteins (such as synaptotagmin ) may further regulate vesicle fusion. It is clear that the process of insulin secretion is highly complex; and research continues to provide new insights into β-cell molecular biology and electrophysiology. Such studies promise to provide a better understanding of insulin action and its deregulation in DM.

INSULIN STRUCTURE

The three-dimensional structure of insulin has been studied in great detail in crystals by X-ray diffraction and in solution by NMR spectroscopy . Such studies have yielded valuable information regarding the folding of proinsulin and function of insulin . As previously mentioned, insulin was first crystallized in rhombohedral form in 1926 ; and almost 10 years later Scott elucidated the importance of zinc ions and other divalent cations in crystallization . In 1969 the structure of hexameric 2-Zn insulin (designated T 6 in modern nomenclature ) was determined by Dorothy C. Hodgkin and coworkers using X-ray methods ; this structure and was later refined to atomic resolution . Currently there are several crystal forms of insulin, defining three structural families of hexamers (T 6 , T 3 R f 3 , and R 6 ) , zinc-free dimers (T 2 ) , and monomeric fragments . These families are shown in schematic form in Figure 5A and as ribbon models in Figure 5B; models of the component T-state protomer and R-state protomer are shown in Figure 6. The solution structure of engineered insulin monomers and dimers resembles the crystallographic T state .

Figure 5 . Structural families of insulin hexamers. A, Schematic representation of the three types of zinc insulin hexamers, designated T 6 , T 3 R f 3 , and R 6 . B, Corresponding ribbon representation of wild-type crystal structures. Axial zinc ions are shown in blue-gray . Coordinates were obtained from Protein Databank entries 4INS , 1TRZ, and 1ZNJ, respectively. Residues B1-B8 exhibit a change in secondary structure as shown in black . T-state protomers are otherwise shown in red , and R-state protomers in blue . For cylinder models of the T- and R-state protomers, see Figure 6 below. This figure is reprinted by Ref with permission of the authors.

The T 6 insulin hexamer (MW ~36000 Da) consists of six molecules of insulin arranged as three dimeric units related by a threefold symmetry axis . The dimers (MW ~12000) possess a pseudo twofold symmetry axis, which is perpendicular to the threefold axis of rotation. Although each protomer within the dimers has similar main-chain structure, they are not identical in the arrangement of certain side chains, breaking the twofold symmetry. The most obvious difference is that the side chain of Phe B25 is folded in towards the hydrophobic core in one protomer but outwards in the other . Two axial Zn 2+ atoms lie on the threefold symmetry axis of the hexamer; each exhibits octahedral coordination by three His B10 residues and three water molecules.

The T 6 insulin structure defines hydrophobic, solvent-exposed, and potential binding surfaces of insulin. This characterization has been supported by NMR-based solution structures of the insulin hexamer , engineered dimer , and engineered monomer . Many additional X-ray structures of insulin , insulin derivatives and insulins of other species, such as the Atlantic hagfish ( Myxine glutinosa ), a variant insulin containing a substitution (His B10  Asn) that prevents zinc binding and hexamer formation . In most of these instances the insulin, or derivative, maintains an overall tertiary structure that corresponds well with the protomers in the T 6 structure. For this reason the T 6 insulin hexamer widely been employed as the prototypic insulin structure. An emerging theme of such studies is that classical crystal structure of the free hormone (T state) depicts spatial relationships pertinent to the folding pathway of proinsulin whereas a key subset of such relationships are altered or broken on receptor binding . The relevance of the crystallographic R state to receptor binding continues to be a source of speculation .

Figure 6 . Classical T and R structures of insulin. Ribbon models of TR dimer based on crystal structures of zinc insulin hexamers. The B chain is shown in black and A chain in green. The position of Gly B8 is shown as a red ball in T-state-specific β -turn (left) or R-state-specific α helix (right). D-amino-acid substitutions at B8 stabilize the T-state but block receptor binding whereas L-amino-acid substitutions destabilize the T-state but can be highly active . The structure of insulin as an engineered monomer in solution resembles the T state. This figure is reprinted from the Supplement to Ref with permission of the author.

Insulin exists primarily as a monomer at low concentrations (~10 -6 M) and forms dimers at higher concentrations at neutral pH . At high concentrations and in the presence of zinc ions, insulin forms hexameric complexes . We shall begin with a discussion of the insulin monomer, which is the circulating state of the molecule in plasma, and then discuss its self-assembly. This section culminates with a description of the IR and evidence for a novel receptor-bound conformation of insulin .

THE INSULIN MONOMER

The A- and B chains of insulin (individually illustrated in Figures 7a and b, respectively) exhibit extensive secondary structure despite their limited lengths (Figure 3). The A chain consists of two α -helical segments (A1-A8 and A12-A19) that are nearly antiparallel. These helices are connected by a non-canonical turn (residues A9-A12), bringing into proximity the N- and C-chain termini. The B chain (Figure 7b) contains central α -helix (residues B9-B19) flanked by disulfide bridges (cystines A7-B7 and A20-B19) and β -turns (B7-B10 and B20-B23); residues B1-B5 are extended in the T state . Each β -turn contains at least one conserved Gly with positive  dihedral angle (residues B8, B20, and B23). The B7-B10 β-turn enables the side chain of His B5 to interact with the central region of the A chain I association with cysteine A7-B7. The B20-B23 β -turn orients the C-terminal segment of the B-chain (residues B23-B30) in close proximity and antiparallel to the central B-chain α -helix. Residues B24-B28 have a β-strand structure. The conserved aromatic side chains of Phe B24 and Tyr B26 are in contact with Leu B11 , Val B12 and Leu B15 of the central B-chain α -helix, defining an α -turn- β supersecondary structure.

Figure 7 . The structures of insulin A- and B-chains. The figure illustrates insulin A-chain (a) and B-chain (b) as determined from the three-dimensional X-ray analysis of the T 6 hexamer (2-Zn insulin). Both chains are viewed perpendicular to the threefold symmetry axis of the insulin hexamer (see text).

Joining of the A- and B-chain buries the sulfur atoms of cystines A6-A11 and A20-B19 as well as the nonpolar side chains of Ile A2 , Val A3 , Leu A16 , Tyr A19 , Leu B6 , Leu B11 , and Leu B15 . The edge of the hydrophobic core in part exposes Val A3 , Ile A13 , Val B12 , Val B18 , Phe B24 , and Tyr B26 . Together, these conserved residues contribute to the stability of the native structure. The latter residues also contribute to the surface of the molecule that is partially exposed to solvent in the monomer and involved in dimerization or hexamer assembly discussed below.

The conserved Phe at position B25 is of special interest. Whereas in the T state its main chain amide group hydrogen bonds to the carbonyl oxygen of Tyr A19 , the side chain can adopt different orientations (Figure 8). In molecule I Phe B25 is folded against the hydrophobic core of the same protomer but in molecule II the aromatic ring is displace outward. The actual orientation of Phe B25 in solution is likely an intermediate conformation .Molecules I and II also display notable differences in A-chain structure. Subtle differences between the two molecules of the dimer are also seen in the pattern of hydrogen bonds in the N-terminal A-chain α -helix (A1-A8). These differences are accentuated in the TR transition (below) and may foreshadow the mechanism of induced fit on receptor binding . Such findings illustrate the general principle that insulin (like other globular proteins) exhibits highly organized structure that may nonetheless undergo adjustments on assembly or interactions with ligands or other proteins.

Figure 8 . Structural illustration of the monomer-monomer interface in the insulin dimer. The dimer is viewed along the crystallographic 2-fold axis. The side chains of residues Val B12 , Leu B15 , Phe B24 , Phe B25 , and Tyr B26 which form the core of the insulin dimer are illustrated in the figures (not labeled: Pro B28 ). Four main-chain hydrogen bonds are formed from the main-chain atoms of Phe B24 and Tyr B26 are illustrated as dotted lines. In the panel (b) is a magnified view of the dimer interface in panel (a).

INSULIN SELF-ASSEMBLY

Dimerization. The T 6 insulin hexamer contains three dimers in which molecules I and II form an extensive nonpolar interface (Figure 8a). The C-terminal segments of each B-chain come together to form an antiparallel β-sheet (residues B24-B28 and its dimer-related mates). This sheet, containing four intermolecular main-chain hydrogen bonds, is further stabilized by hydrophobic interactions involving the side chains of Val B12 , Tyr B16 , Phe B24 , Tyr B26 , Pro B28 , and to some extent, Phe B25 (Figure 8b). These residues are shielded from contact with solvent (with the exception of Phe B25 ).Although dimerization is associated with local and non-local damping of conformational fluctuations within the protein (relative to the isolated monomer) , an entropic drive is obtained from desolvation of non-polar surfaces, predicted to liberate bound water molecules into the bulk solution . Dimerization does not require zinc ions and exhibits a dissociation constant K d of approximately 10 -5 M .

Hexamer Formation . In the presence of Zn 2+ , insulin dimers associate to form hexameric units coordinated by two zinc ions within the central axis of the hexamer (Figure 6). These axial Zn 2+ ions are coordinated to the imidazole groups of His B10 (three per zinc ion) and in a particular instance to His B5 . Several hexameric forms of insulin have been crystallized of which the classical T 6 structure (“2-Zn insulin” ) remains the prototype . In the this hexamer three dimers are related by a 3-fold symmetry axis, which is located in the hydrophilic pore at the center of the hexameric unit that connects the two Zn 2+ ions . Each zinc ion is octahedrally coordinated to three His B10 imidazole nitrogens and three water molecules. The three-fold symmetry axis is perpendicular to the approximate two-fold symmetry axis of the dimers. Contacts between dimers in the hexamer are less extensive than contacts between protomers within the dimer . The T 6 hexamer is approximately 50 Å in diameter and 35 Å in height .

Allostery among Hexamers . In crystals and in solution insulin forms three structural families of hexamers (T 6 , T 3 R f 3 , and R 6 ). The equilibrium between these structures is modulated by salt concentration and the binding of phenolic ligands (which favors the R state or frayed R f state). The T 3 R f 3 hexamer (formerly designated 4-Zn insulin with rhombohedral crystal form) and R 6 hexamer are arranged similarly to the classical T 6 hexamer in overall aspects. The local and non-local structural rearrangements among these three families of hexamers are collectively designated the TR transition . Molecular analysis of this transition has provided an influential biophysical model for the propagation of conformational change in protein assemblies . Because elements of the TR transition may also pertain to the mechanism of receptor binding (below), we shall describe salient features of the T 3 R f 3 , and R 6 hexamers in turn.

T 3 R f 3 Hexamers (4-Zn Insulin). The classical “rhombohedral transition” of zinc insulin crystals was induced by high concentrations of sodium chloride . The structural basis of this transition was elucidated by D. C. Hodgkin and coworkers in 1976 . Each dimeric unit consists of one molecule I and one molecule II monomer. Whereas in the hexamer the molecule I trimer (T 3 ) has the same octahedral zinc-ion coordination as in the T 6 hexamer, the molecule II trimer (R f 3 ) exhibits substantial, however, displays structural reorganization. The N-terminal B-chain residues B3-B8 (with “fraying” of flexible terminal residues B1 and B2) forms a continuous extension of the central B9-B19 α -helix. This transition in secondary structure, which entails a movement of more than 25 Å at B1 , is coupled to a change in coordination of the second axial zinc ion from octahedral to tetrahedral; the stoichiometry of zinc ions per hexamer is on average 2.67 . The TR transition also causes a rotation of the A1-A8 α -helix (thus requiring a reorganization of the details of side-chain packing in the hydrophobic core and change in conformation of the A7-B7 disulfide bridge) and small displacement of the B24-B28 β -strand away from the A chain, breaking the main-chain hydrogen bond between Phe B25 and Tyr A19 . Similar T 3 R f 3 hexamers may be induced at lower salt concentrations by phenolic ligands wherein the R f 3 trimer contains three bound phenolic molecules .

R 6 Hexamers. High concentrations of phenolic ligands induce a further conformation change to form the R 6 hexamer . The hexamer contains six (or uncommonly seven) bound phenolic ligands. Crystal forms exist which exhibit rigorous sixfold symmetry or which contain six independent protomers in the asymmetric unit with only quasi-sixfold symmetry. In the R 6 hexamer each protomer contains a continuous B1-B19 α -helix and breakage of the B25-A19 main-chain hydrogen bond association with a small displacement of the B-chain C-terminus segment from the A chain. The specific binding site for the phenolic ligand does not pre-exist in the T 6 structure but is may occur in nascent form as part of an extended conformational equilibrium among the three hexamer types. In this R-state-specific binding pocket two hydrogen bonds engage the phenolic hydroxyl group from the A6 carbonyl oxygen and A11 amide hydrogen. The side chain of His B5 packs against each phenolic molecule. Tetrahedral coordination of the zinc ions resembles that of the salt-induced R f 3 trimer of 4-Zn insulin (above) .

Although the TR transition was originally defined in the crystalline state, spectroscopic studies have verified that an analogous equilibrium exists in solution . This conformational transition of B-chain secondary structure has been studied extensively in solutions of hexameric insulin by monitoring changes in the coordination of Co 2+ (Ni 2+ and Co 2+ also form hexamers) from octahedral to tetrahedral using visible-absorbance spectroscopy , changes in α -helical content detected using circular dichroism , or 2D-NMR . The solution structure of the phenol-stabilized R 6 hexamer resembles the crystal structure .

In addition to the clues provided by the TR transition with respect to the mechanism of receptor binding (next section), the phenol-stabilized R 6 hexamer exhibits augmented thermodynamic and kinetic stability relative to the T6 hexamer. Retarding physical- and chemical degradation of the polypeptide chains, these favorable biophysical properties have been exploited in pharmaceutical formulations to increase the shelf-life of insulin products . Because phenolic ligands were traditionally employed in insulin formulations due to their bacteriostatic properties , their additional role as protein-stabilizing agents and their elegant structural role in the hexamer represents the value of serendipity as a source of therapeutic advance .

INSULIN STRUCTURE-FUNCTION RELATIONSHIPS

It is the insulin monomer that binds to and triggers the insulin receptor. A key unresolved issue is the extent to which the monomer undergoes a change in conformation on binding to the insulin receptor. It is likely that the molecular understanding of how insulin binds were be deepened in the next five years through advances in structural biology of the insulin receptor .

The predominance of structural information as described above pertains to insulin hexamers as described above. Although of biological relevance (insulin undergoes zinc-dependent hexamer self-assembly and micro-crystallization in the secretory granule of the β -cell ), it is unlikely that a substantial amount of hexamer exists in the plasma wherein the concentration of zinc ions is low. Further, although zinc-free dimers are be present in the portal circulation, progressive dilution of the secreted insulin in the systemic circulation would lead to a predominance of monomeric molecules. NMR studies confirm that the conformation of the free monomer in solution resembles that of the T-state crystallographic protomer , but its flexibility raises the possibility that receptor binding is associated with induced fit.

With this caveat in mind, the three-dimensional crystal structure of insulin has nonetheless allowed specific residue positions and side-chain orientations to be related to biological activity. Such analogs have been obtained by synthetic methods , comparison of species variants , and site-directed mutagenesis . Together, such analyses of structure-activity relationships in insulin have yielded an understanding of which residues and positions are necessary for receptor binding . Although such data may be confounded by indirect effects of amino-acid substitutions on the structure of the hormone, overall aspects of the long-sought structure of the hormone-receptor complex have been inferred from photo-cross-linking studies and recently been confirmed in a low-resolution co-crystal structure of insulin bound to a fragment of the receptor ectodomain .

Several assays have been used to determine the binding potency of insulin analogs such as (a) the in vivo mouse convulsion assay, (b) in vitro receptor binding studies of analogs in competition with radio-iodinated insulin, and (c) by the ability of insulin analogs to enhance 14 C-glucose oxidation, or conversion of 3 H-glucose into lipids in adipocytes . Most of these studies show a strong correlation between receptor binding and biological activity , except for high-binding affinity analogs (>120%) which show only 100% activity in vivo probably due to rapid clearance or very low-binding analogs which may accumulate at the cell surface and generate a higher-than-expected activity due to decreased clearance .

Three conserved regions in insulin have been of particular interest in the primary receptor-binding surface of insulin: (i) the N-terminal and C-terminal segments of the A chain (Gly A1 -Ile A2 -Val A3 -Glu A4 and Tyr A19 -Cys A20 -Asn A21 ), (ii) the central α -helix of the B chain (especially Val B12 ) and (iii) and the C-terminal segment of B chain (Phe B24 -Phe B25 -Tyr B26 ). All of these residues are located on or near the surface of insulin and therefore may interact with insulin receptor . This surface is notable for clinical mutations associated with a monogenic syndrome of adult-onset diabetes mellitus. The substitutions are Val A3  Leu (Insulin Wakayama), Phe B24  Ser (Insulin Los Angeles), and Phe B25  Leu (Insulin Chicago) . Whereas the A3 and B25 mutations markedly impair receptor binding, Ser B24 impairs binding by less than tenfold as will be discussed in the final section of this chapter . Evidence for the proximity of these three surfaces to the insulin receptor has been obtained by residue-specific photo-cross-linking studies .

The above surfaces of insulin are classified as its “Site-1” related binding surface by DeMeyts and colleagues in relation to a proposed “Site-2” related surface . Sites 1 and 2 pertain to a proposed architecture and mode of binding of the insulin receptor . The putative Site-2 related surface of insulin, although not rigorously established in the hormone-receptor complex, is proposed to correspond to its hexamer-forming surface, including residues His B10 , Leu B17 , Val B18 , Ser A12 , Leu A13 and Glu A17 . Substitutions in Site 2 affect the kinetic properties of hormone binding disproportionately to effects on affinity. Such kinetic properties (related to the residence time of the hormone-receptor complex) correlate with relative post-receptor signaling pathways; prolonged residence times favor mitogenic signaling relative to metabolic signaling . Although the location of Site 2 in the ectodomain of the insulin receptor is not well defined, such interactions are likely to be of pharmacological interest in relation to the risk of cancer in patients exposed to high doses of insulin .

We discuss in turn structure-activity relationships in the A- and B chains and conclude this section with a brief summary of structural relationships in the ectodomain of the insulin receptor.

A-Chain Analysis. The low-resolution structure of insulin bound to a fragment of the receptor strongly suggests that the A chain retains its native secondary structure (with two α -helices) and tertiary U-shaped structure on receptor binding .

The N-terminal residues of the A chain are conserved among vertebrate insulins and have been extensively investigated for their relevance in ligand receptor interactions. N-acetylation of the A-chain N-terminus results in a reduction of receptor binding to approximately 30%, suggesting the importance of a positively charged free amino group at A1 . Deletion of Gly A1 also results in a reduction in receptor binding to 15% of native hormone. Substitution of Gly A1 by diverse L-amino acids results in analogs having reduced binding of 2-20% whereas D-substitutions are well tolerated . The precise size, shape and hydrophobicity of Ile A2 and Val A3 are stringently required for high-affinity receptor binding . The A2 and A3 side chains adjoin the C-terminal α -helix of the receptor α -subunit (designated α CT; see below) in the low-resolution structure of insulin bound to a receptor fragment . Although the details of side-chain packing were not resolved, a nonpolar interface is implied by the predicted registry of the respective α -helices at this interface. Such contacts are in accord with photo-cross-linking studies .

The invariant C-terminal A-chain residues (Tyr A19 -Cys A20 -Asn A21 ) have also been studied and may have dual roles in structure and function. Whereas the primary role of Tyr A19 is likely to be structural through its long-range packing with the side chain of Ile A2 , the para -OH group of Tyr A19 is exposed to solvent and may contact the receptor. Substitution by Phe or Trp or modification of the ring by mono- or diiodination impairs activity. (The other tyrosine in the A-chain (Tyr A14 ) is not conserved and may be modified with little change in activity .) Removal also impairs activity but substitutions are well tolerated. In the crystal structure the A21 main-chain amide donates a hydrogen bond to the main-chain carbonyl of Gly B23 , and Katsoyannis and colleagues have provided elegant evidence (based on the inductive effect of fluoro-substitutions) that the strength of this hydrogen bond contributes to the efficiency of disulfide pairing in chain combination . The side chain of Asn A21 was not well defined in the low-resolution structure of insulin bound to a receptor fragment .

B-Chain Analysis . The B chain of insulin has been more extensively studied than the A chain, particularly with respect to the TR transition (B1-B8) and receptor-binding determinants in the C-terminal β -strand (B24-B28) . Although neither of these segments was well visualized in the low-resolution structure of insulin bound to a receptor fragment , their absence is likely to reflect technical features of the model system (such as disorder in the co-crystals or absence of the fibronectin-homology receptor domains; see below). Nonetheless, the relative positions of structural elements in the low-resolution co-crystal structure suggests that the B chain (unlike the A chain) undergoes a change in conformation involving its unseen N- and C-terminal segments.

The immediate N-terminal B-chain residues (Phe B1 -Val B2 -Asn B3 -Gln B4 ) can be deleted with only modest reductions in biological activity (60-70% that of insulin in lowering blood glucose in rabbits) . These results suggested that the N-terminal four residues are of limited significance in the hormone-receptor complex. By contrast, successive removal of His B5 results in an analog, des -pentapeptide(B1-B5)insulin, that possesses only 15% activity . Further deletion of Leu B6 results in a compound with < 1% binding affinity; and substitution of Leu B6 by other amino acids (Gly, Ala, and Phe) in full-length insulin results in reduced binding [< 0.1%-10%] in the order presented . Cystine A7-B7 lies on the surface of the free insulin monomer and may in principle contribute to receptor binding.

Evidence for the importance of Gly B8 to the biological activity of insulin has been obtained by non-standard mutagenesis . This position represents the crux of the TR transition. In the T-state Gly B8 lies at the junction between the central α -helix of the B-chain (as part of the B7-B10 β -turn) and its N-terminal segment whereas in the R state Gly B8 (like His B5 ) is part of an extended α -helix. Gly B8 exhibits different  dihedral angles in the two states: positive in the T-specific β -turn and negative in the R-specific α -helix. Substitution of Gly B8 by D- (or L-) amino acids leads to stereospecific stabilization (or destabilization) of the T-state. Remarkably, the stabilizing D-substitutions markedly impair receptor binding . This impairment is associated with a shift in the conformational equilibrium among T 6 , T 3 R f 3 , and R 6 hexamers favoring the T-state . The low biological activity of such nonstandard analogs is ascribed to stabilization of a native-like but inactive T-state conformation . In four different low-resolution structures of insulin or insulin analogs bound to a receptor fragment , the inferred  dihedral angle of Gly B8 appears to be R-like in three of the structures and T-like in the remaining structure. It is possible that Gly B8 is a site of conformational change in the hormone-receptor complex but not coupled to an R-like α -helical transition of residues B1-B7. An alternative possibility is that a Gly B8 “switch” functions to adjust the conformation of cystine A7-B7 and in turn optimize the spatial relationship between the central B-chain α -helix and N-terminal A-chain α -helix.

Importance of the B-Chain COOH-terminus. The B-chain C-terminal β -strand is the most extensively investigated region of insulin. Deletion analysis has revealed that the C-terminal residues Tyr B26 , Thr B27 , Pro B28 , Lys B29 , and Thr B30 may be removed; the resulting analog des -pentapeptide(B26-B30)-insulin-amide (i.e., lacking a C-terminal carboxylate) is fully active . This and related truncated templates have been widely employed in studies of structure-activity relationships and for synthesis of a pioneering B25-specific photo-cross-linking reagent by Katsoyannis and colleagues . The C-terminal five residues are nonetheless necessary for dimerization and hexamer formation . Crystal and NMR-based structures of such truncated analogs retain native-like structures in the α -helical core of the protein . Despite the dispensability of residues B26-B30 for receptor binding, successive removal of Phe B25 and Phe B25 (with analogous C-terminal amidation) progressively impairs receptor binding (relative affinities 6% and 0.2% binding, respectively) . Photo-cross-linking studies have provided evidence that these side chains contact the insulin receptor and may also provide sites of conformational change .

Phe B24 . The aromatic ring of Phe B24 stabilizes the B20-B23 β -turn and seals one edge of the hydrophobic core adjoining cysteine A20-B19. The functional importance of Phe B24 is indicated by the efficient photo-cross-linking of a para -azido-Phe B24 probe to insulin receptor and by the low activities of diverse analogs, including Tyr B24 and Ala B24 (each impaired by >50-fold) . Surprisingly, however, substitution of Phe B24 by Gly is well tolerated , and D-amino-acid substitutions can even enhance receptor binding . Although a range of NMR and fluorescence studies of such anomalous analogs have been reported with varying results (which may in part reflect protein flexibility rather than stable elements of structure) , NMR studies of a super-active engineered monomer containing D-Ala B24 have suggested that the main-chain at position B24 may be a site of a conformational change on receptor binding , leading to detachment of the B24-B28 β -strand from the α -helical core of the hormone as long envisaged . Such detachment is supported by the low-resolution co-crystal structure of a model hormone-receptor complex . Interestingly, D-amino-acid substitutions at position B24 impair (rather than enhance) the binding of truncated insulin analogs lacking residues B26-B30 . A possible explanation for this finding is that chiral perturbation is not needed to displace the truncated β -strand whereas the D-side chain itself is less favorably oriented in such analogs.

Phe B25 . Photo-reactive insulin analogs containing para -azido-Phe or para -benzoyl-Phe at position B25 efficiently cross-link to the receptor . For this reason and due to the marked inactivity of Phe B25  Leu (Insulin Chicago) , substitutions of Phe B25 has been extensively studied. High activity requires a trigonal (sp2 hybridized)  -carbon as in aromatic side chains rather than a tetrahedral (sp3 hybridized)  -carbon as in Leu . Tolerance of β -napthyl-alanine nonetheless suggests that there is room at the hormone-receptor interface for a larger aromatic side chain. Interestingly, the activities of certain low-binding analogs, such as [homo-Phe B25 ]-insulin, are in part rescued when B26-B30 is removed. These experiments suggest that PheB25 may make two contributions to receptor binding: the specific docking of its side chain and as a further site (in addition to B24) of main-chain conformational change leading to detachment of the B-chain C-terminal segment.

STRUCTURE OF THE INSULIN RECEPTOR

Nearly forty-five years after the structure of insulin was first solved, it is known only in part how insulin binds to its receptor, progress being slowed by the biochemical and structural complexity of the insulin receptor, a heavily-glycosylated disulfide-linked ( αβ ) 2 dimer.

In the biosynthesis of the insulin receptor each pro-receptor monomer is proteolytically cleaved into an N-terminal α -chain and C-terminal β -chain linked by a single disulfide bond. The extracellular portion of the ( αβ ) 2 dimer includes both α -chains as part of each transmembrane β-chain. Each receptor monomer consists of several structural domains: from the N-terminus, a leucine‑rich repeat domain L1 (residues 1-157), a cysteine‑rich region (CR, residues 158-310) a second leucine‑rich repeat domain L2 (residues 311-470), and three fibronectin type-III domains: FnIII‑1 (residues 471-595), FnIII‑2 (residues 596-808) and FnIII‑3 (residues 809-906). FnIII‑2 contains a ~120‑residue insert domain (ID, residues 638-756) which contains the α / β cleavage site. C‑terminal of the FnIII‑3 domain lies a single transmembrane α -helix, followed by a ~40‑residue intracellular juxtamembrane region (JM), a tyrosine kinase (TK) catalytic domain and a ~100‑residue C‑tail.

Crystallographic studies of the free ectodomain by C. Ward and colleagues have shown that the extracellular domain assembles as an inverted “V”; in each protomer the L1-CR-L2 domains forming one leg and the three FnIII domains the other (Figure 9). One protomer is related to the other in the dimer by a two-fold rotation about the axis of the inverted “V”, resulting in the L1-CR-L2 leg of one monomer being packed against the three FnIII domains of the other. As discussed above, current models envisage that hormone binding is mediated by two adjoining structural elements termed Site 1 and Site 2 . Both sites are required for high-affinity hormone binding and negative cooperativity .

Figure 9 . Structure of the free ectodomain of the insulin receptor. The original structure was determined by Ward and colleagues and refined to include the α CT element (magenta within dashed ovals) by Lawrence and colleagues . The disulfide-linked dimer of intact a subunits and truncated β subunits (( αβ ) 2 ) was stabilized by F ab immunoglobulin fragments. The dimeric structure adopts an inverted-V conformation. Abbreviations: α CT, C-terminal α -helical element of the α -subunit; CR, cysteine-rich domain; FnIII-1,2, and 3, fibronectin homology domains 1, 2 and 3; ID α , the portion of the insert domain within the α -subunit; and L1 and L2, large domains 1 and 2. We thank M. C. Lawrence for preparation of the figure (reprinted from the web-based Supplement to Ref with permission).

The primary hormone-binding surface is provided by Ste 1, containing two distinct regions: (i) the central of the three β ‑sheets that make up the L1 domain, and (ii) the last 16 residues of the α -chain (the so-called α CT segment). In the intact ectodomain the L1 and α CT elements belong to different α -subunits in the ( αβ ) 2 dimer. Co-crystals diffracting to 3.9 Å have been obtained of insulin or truncated insulin analogs bound to an L1-CR fragment of the ectodomain together with a synthetic α CT peptide . Interpretation of the electron-density map was aided by modeling based on the known structure of the L1-CR-L2 fragment and the lack of significant structural changes in the complex. Three tubes of electron density were observed fitting the known lengths and orientations of the three α -helices in insulin (the central B-chain α -helix and two A-chain α -helices). The resulting model (Figure 10) is remarkable for its unexpected features. Whereas extensive interactions had been predicted between insulin and L1, the major interface is between insulin and α CT. The role of L1 is to bind and orient the opposite face of α CT. Further, the position and orientation of α CT on the nonpolar surface of L1 differed from its position and orientation in the free ectodomain . Such repositioning is significant as the bound location of α CT is such as to displace residues B26-B30 from their positions in classical structures of insulin. As a further surprise, with the exception of the B-chain C-terminal segment (not visualized in the reported structure), it is the A chain (and not the B chain) that provides the more extensive receptor-binding surface. The key A-chain recognition element comprises A1-A4 as anticipated by mutagenesis and chemical modification as discussed above.

Figure 10 . Structure of the “micro-receptor” complex between insulin and a fragment of the insulin receptor . This ternary complex contains an L1-CR fragment of the receptor α -subunit, a synthetic α CT peptide, and insulin. Abbreviations: α CT, C-terminal α -helical element of the α -subunit; CR, cysteine-rich domain; InsA and InsB, respective A- and B chains of insulin; and L1, large domain 1. We thank M. C. Lawrence for preparation of the figure (reprinted from Ref with permission).

The current low-resolution structure rationalizes a wealth of prior biochemical data but leaves unanswered many questions. How and why the B chain changes conformation on receptor binding and how such changes may be propagated to effect signal transduction are not clear. Further, because residues B1-B7 were not visualized, the relevance of an R-like transition in secondary structure could not be definitively evaluated. A possible location for Site 2 was suggested when the structure of the insulin receptor ectodomain dimer was determined : loops at the junction of the FnIII‑1 and FnIII‑2 domains from the protomer opposite to that contributing the L1 domain to Site 1. This model is consistent with the "cross-linking" trans binding mechanism of receptor activation . To date no co-crystals have been reported with receptor fragments containing Site 2.

MUTATIONS IN THE INSULIN GENE

The classical insulopathies (Insulins Chicago, Los Angeles and Wakayama) accumulated in the blood stream as mutant insulins. These all have point mutations that lower receptor binding and biological activity. Rare clinical syndromes result from mutations elsewhere affecting key steps in biosynthesis within the β -cell .

A monoallelic genetic mutation resulting in substitution of His B10 by Asp leads to the secretion of a mutant proinsulin (319): Asp B10 -proinsulin (Proinsulin Providence) due to its selectively increased export via constitutive pathways of a significant proportion of the prohormone (approx. 1/3); this fraction thus escapes processing but also undergoes increased endosomal pathway-mediated degradation, resulting in circulating hyperproinsulinemia in a mouse model . No defect was observed in the processing of the fraction of the Asp B10 proinsulin that remained within the regulated secretory pathway. Asp B10 -insulin also exhibits enhanced binding affinity and this property, in its prohormone may be related to its possible mis-sorting, i.e. via insulin receptor mediated intracellular trafficking. Expression in cultured β-cells has confirmed that the mutant proinsulin can be fully processed at normal rates, but a large fraction of unprocessed mutant prohormone is released via an unregulated-constitutive pathway . Proinsulin processing is also impaired when Arg65 of the C-domain dibasic pair (Lys64-Arg65) is mutated to His65 or Leu65 (Proinsulin Boston/Denver/Tokyo and Proinsulin Kyoto, respectively) . In all these cases the defect lies in the inability of processing enzymes (PC3/PC1 and PC2) to cleave at the mutated cleavage site, which leads to the secretion of a partially cleaved intermediate form of proinsulin.

Figure 11 . Mutations in the insulin moiety of proinsulin associated with neonatal diabetes mellitus. (A) Such mutations may be classified as cysteine-related or non-cysteine-related. The latter provide probes for structural determinants of the efficiency of disulfide pairing in the redox-coupled folding of nascent proinsulin (for review, see ).

It has recently been found that the majority of dominant negative mutations in the insulin gene affecting biosynthesis and folding of the hormone lead to permanent neonatal-onset DM with impaired secretion of the variant or wild-type insulin or proinsulin . Impaired β -cell function often develops prior to maturation of the immune system and so presents as an auto-antigen negative form of apparent Type 1 DM. Dominant mutations in the insulin gene define the second most common genetic cause of permanent neonatal DM (relative to heterozygous activating mutation in a subunit of the β -cell voltage-gated potassium channel, either KCNJ11 (encoding the Kr6.2 subunit) or ABCC8 (encoding the Sur1 subunit) ). Rarely the β-cells survive for years or even decades in the cases of apparently less damaging insulin mutations, as mentioned below.

Mutations in the insulin gene causing neonatal diabetes occur in each region of preproinsulin, including its signal peptide, A-, B- and C domains. The majority result in addition or removal of a cysteine, leading in either case to an odd number of potential pairing sites: the resulting imbalance leads in general to misfolding and aggregation . Mutations identified in the insulin moiety of proinsulin are shown in Figure 11A (sequence) and Figure 11B (structure). One human mutation encodes the same “Akita” substitution (Cys A7  Tyr) as in the Ins2 gene of the Mody4 mouse ; this dominant murine substitution thus provides a relevant model of progressive β -cell failure . The variant murine proinsulin in vitro undergoes partial unfolding with increased aggregation . Analogous perturbations were characterized in human insulin- and proinsulin analogs lacking cystine A7-B7 . Heterozygous expression of a variant Ins2 allele encoding Cys A6  Ser (identified in an N-ethyl-N-nitrosourea screen) likewise induces DM . The identification of identical human and murine mutations at position A7 suggests that the pathogenesis of neonatal DM in these patients is likewise similar to that of the Akita mouse . Biochemical studies of clinical variants in β -cells or neurosecretory cell lines have revealed perturbations in disulfide pairing, which range from severe or mild depending on the site of mutation and the properties of the substituted side chain .

Akita β -cells exhibit an early defect in the folding and trafficking of both wild-type and variant proinsulins with elevated markers of ER stress, electron-dense deposits in abnormal ER and GA, mitochondria swelling, and progressive loss of β -cell mass . Extension of such findings to human variants is supported by mouse models of human mutant proinsulin constructs .

Figure 12 . Structural features of insulin contributing to foldability. Stereo stick representations of selected regions of the insulin protomer (2-Zn molecule 1; Protein Databank identifier 4INS). Unless otherwise indicated, A-chain is shown in gray and B-chain in black. Selected A-chain side chains are highlighted in color (below); golden balls indicate sulfur atoms of disulfide bridges. (A) Environment of solvent-exposed cystine A7-B7 showing key role of the phi dihedral angle of Gly B8 (arrow), predicted to influence orientation of thiolate B7 to favor disulfide pairing. The side chains of His B5 and Leu B6 are shown in red. (B) Environment of internal cystine A6-A11 highlighting packing of Leu B6 and Ile B11 (red). (C) Environment of internal cystine A20-B19 highlighting packing of A- and B-chain side chains in hydrophobic core: (magenta) Leu A16 and Tyr A19 ; (red) Val B12 , Leu B15 , Gly B23 (ball at C α position), and Phe B24 . Important structural sites not shown: the inter-chain A3-related crevice (lined by Ile A2 , Tyr B26 and Pro B28 ) and T-state-specific tertiary contacts between the N-terminal arm of the B-chain (residues B1-B5) and the A-chain. This figure is adapted from Ref with permission of the author.

Whereas an odd number of cysteines presumably induces a severe block to folding, non-cysteine-related mutations are also observed in such patients (Figure 11). Such sites of mutation identify residues whose side chain or main-chain conformation contributes to the efficiency of folding. Selected structural relationships in the native state presumed to direct specific disulfide pairing in the oxidative folding pathway of proinsulin are illustrated in Figure 12. The folding process may be visualized as a series of trajectories on free-energy landscapes of progressive steepness with successive disulfide pairing (Figure 13A). This pathway highlights the importance of cysteine A20-B19 as a key initial folding intermediate associated with partial formation of native-like super-secondary structure as shown at left in Figure 13B . Competing subsequent routes to the native state are likely to yield a variety of partial folds with accessible unpaired cysteines . Aberrant disulfide bond formation between such partial folds may account for the interruption of wild-type proinsulin folding in the presence of a non-foldable clinical variant.

Figure 13 . Energy landscape view of proinsulin folding and disulfide pairing. (A) Formation of successive disulfide bridges may be viewed as enabling a sequence of folding trajectories on a succession of steeper funnel-shaped free-energy landscapes. (B) Preferred pathway of disulfide pairing begins with cystine A20-B19 (left), whose pairing is directed by a nascent hydrophobic core formed by the central B-domain α -helix (residues B9-B19), part of the C-terminal B-chain β -strand (B24-B26), and part of the C-terminal A-domain α -helix (A16-A20). Alternative pathways mediate formation of successive disulfide bridges (middle panel) en route to the native state (right). The mechanism of disulfide pairing is perturbed by clinical mutations associated with misfolding of proinsulin. Sites of non-cysteine-related mutations causing neonatal DM (Figure 11 above) highlight native structural features critical to foldability. Figure is reprinted from Ref with permission of the author.

Among these, a few better-tolerated mutations present later in life as auto-antibody-negative presumed Type 1 DM or Type 2 DM. One such mutation, presenting in the second decade as maturity-onset diabetes of the young (MODY), is due to substitution of Arg B22 by Gln. This mutation alters a solvent-exposed site in the B21-B24 β -turn not required for receptor binding . DM presenting in the third decade of life is associated with substitution of Phe B24 by Ser . This mutation causes only a mild impairing of receptor binding but like Gln B22 imposes ER stress at a level intermediate between over-expression of wild-type proinsulin and neonatal variants.

Chronic elevation of more moderate ER stress in β -cells due to subtle mutations in proinsulin (such as substitution of Arg B22 by Gln and the classical SerB24 allele of Insulin Chicago) presumably leads to a slower loss of β -cell mass . Such milder mutations are associated with onset of diabetes later in childhood or early adulthood. Evidence is also accumulating that subtle perturbations of wild-type insulin biosynthesis may contribute to the pathogenesis of non-syndromic Type 2 diabetes as in the Akita mouse . A newly recognized biophysical contribution to such β -cell dysfunction is “molecular crowding” in the ER due to over-expression of nascent proinsulin in the face of peripheral insulin resistance , a concept that may have therapeutic implications .

REFERENCES

1. Steiner, D.F. 1967. Evidence for a precursor in the biosynthesis of insulin. Trans. N. Y. Acad. Sci. 30:60-68.

2. Adams, M.J., Blundell, T.L., Dodson, E.J., Dodson, G.G., Vijayan, M., Baker, E.N., Hardine, M.M., Hodgkin, D.C., Rimer, B., and Sheet, S. 1969. Structure of rhombohedral 2 zinc insulin crystals. Nature 224:491-495.

3. Blundell, T.L., Cutfield, J.F., Cutfield, S.M., Dodson, E.J., Dodson, G.G., Hodgkin, D.C., Mercola, D.A., and Vijayan, M. 1971. Atomic positions in rhombohedral 2-zinc insulin crystals. Nature 231:506-511.

4. Brange, J., Ribel, U., Hansen, J.F., Dodson, G., Hansen, M.T., Havelund, S., Melberg, S.G., Norris, F., Norris, K., and Snel, L. 1988. Monomeric insulins obtained by protein engineering and their medical implications. Nature 333:679-682.

5. Brange, J., and Vølund, A. 1999. Insulin analogs with improved pharmacokinetic profiles. Adv. Drug Deliv. Rev. 35:307-335.

6. Hirsch, I.B. 2005. Insulin analogues. N. Engl. J. Med. 352:174-183.

7. Bell, G.I., Pictet, R.L., Rutter, W.J., Cordell, B., Tischer, E., and Goodman, H.M. 1980. Sequence of the human insulin gene. Nature 284:26-32.

8. Steiner, D.F., Tager, H.S., Chan, S.J., Nanjo, K., Sanke, T., and Rubenstein, A.H. 1990. Lessons learned from molecular biology of insulin-gene mutations. Diabetes Care 13:600-609.

9. Stoy, J., Edghill, E.L., Flanagan, S.E., Ye, H., Paz, V.P., Pluzhnikov, A., Below, J.E., Hayes, M.G., Cox, N.J., Lipkind, G.M., et al. 2007. Insulin gene mutations as a cause of permanent neonatal diabetes. Proc. Natl. Acad. Sci. U. S. A. 104:15040-15044.

10. Weiss, M.A. 2013. Diabetes mellitus due to the toxic misfolding of proinsulin variants. FEBS Lett. 587:1942-1950.

11. Greeley, S.A., Tucker, S.E., Naylor, R.N., Bell, G.I., and Philipson, L.H. 2010. Neonatal diabetes mellitus: a model for personalized medicine. Trends Endocrinol. Metab. 21:464-472.

12. Liu, M., Hodish, I., Haataja, L., Lara-Lemus, R., Rajpal, G., Wright, J., and Arvan, P. 2010. Proinsulin misfolding and diabetes: mutant INS gene-induced diabetes of youth. Trends Endocrinol. Metab. 21:652-659.

13. De Meyts, P., and Whittaker, J. 2002. Structural biology of insulin and IGF1 receptors: implications for drug design. Nat. Rev. Drug Discov. 1:769-783.

14. Greeley, S.A., Naylor, R.N., Philipson, L.H., and Bell, G.I. 2011. Neonatal diabetes: an expanding list of genes allows for improved diagnosis and treatment. Curr. Diab. Rep. 11:519-532.

15. Scott, E.L. 1912. On the influence of intravenous injections of an extract of the pancreas on experimental pancreatic diabetes. Am. J. Physiol. 29:306-310.

16. Duvillie, B., Cordonnier, N., Deltour, L., Dandoy-Dron, F., Itier, J.M., Monthioux, E., Jami, J., Joshi, R.L., and Bucchini, D. 1997. Phenotypic alterations in insulin-deficient mutant mice. Proc. Natl. Acad. Sci. U. S. A. 94:5137-5140.

17. Jonsson, J., Carlsson, L., Edlund, T., and Edlund, H. 1994. Insulin-promoter-factor 1 is required for pancreas development in mice. Nature 371:606-609.

18. Schӓfer, E.A. 1916. The Endocrine Organs: An introduction to the study of internal secretion . London: Longmans, Green, and Co. 27 pp.

19. Bliss, M., editor. 1982. The discovery of insulin . Chicago, Illinois: University of Chicago Press. 93-99 pp.

20. Nandi, A., Kitamura, T., Kahn, C.R., and Accili, D. 2004. Mouse models of insulin resistance. Physiol. Rev. 84:623-647.

21. Leslie, R.D.G., and Robbins, D.C. 1995. Diabetes: Clinical Science in Practice : Cambridge University Press. 492 pp.

22. Von Mering, J., and Minkowski, O. 1890. Diabetes mellitus nach pankreas exterpation. Arch. Exp. Path. Pharmacol. 26:371.

23. Barron, M. 1920. The relation of the islets of Langerhans to diabetes with special reference to cases of pancreatic lithiasis. In Surgery, gynecology, and obstetrics . Chicago: The Surgical Publishing Company of Chicago. 437-448.

24. Langerhans, P. 1869. Beitrage zur Mikroskopischen Anatomie der Bauchspeicheldruse. In Institute of Pathology . Berlin: University of Berlin. 32.

25. Banting, F.G., and Best, C.H. 1922. The internal secretion of the pancreas. J. Lab. Clin. Med. 7.

26. Ullrich, K.J. 1979. Sugar, amino acid, and Na + cotransport in the proximal tubule. Annu. Rev. Physiol. 41:181-195.

27. Kahn, C.R., Neville, D.M., and Roth, J. 1973. Insulin receptor interaction in the obese hyperglycemic mouse. A model of insulin resistance. J. Biol. Chem. 248:244-250.

28. Petruzzelli, L., Herrera, R., and Rosen, O.M. 1984. Insulin receptor is an insulin-dependent tyrosine protein kinase: copurification of insulin-binding activity and protein kinase activity to homogeneity from human placenta. Proc. Natl. Acad. Sci. U. S. A. 81:3327-3331.

29. Ullrich, A., Bell, J.R., Chen, E.Y., Herrera, R., Petruzelli, L.M., Dull, T.J., Gray, A., Coussens, L., Liao, Y.C., Tsubokawa, M., et al. 1985. Human insulin receptor and its relationship to the tyrosine kinase family of oncogenes. Nature 313:756-761.

30. Ward, C.W., and Lawrence, M.C. 2011. Landmarks in insulin research. Front. Endocrin. 2:1-11.

31. De Meyts, P. 2008. The insulin receptor: a prototype for dimeric, allosteric membrance receptors? Trends Biochem. Sci. 33:376-384.

32. Hopfer, U. 1987. Membrane transport mechanisms for hexoses and and amino acids in the small intestine. In Physiology of the Gastrointestinal Tract . L.R. Johnson, editor. New York: Raven Press. 1499-1526.

33. Haase, W., and Koepsell, H. 1989. Electron microscopic immunohistochemical localization of components of Na+-cotransporters along the rat nephron. Eur. J. Cell Biol. 48:360-374.

34. Rosholt, M.N., and King, P.A. 1995. Diabetes: Clinical science in practice. In Diabetes: Clinical science in practice . R.D.G. Leslie, and D.C. Robbins, editors. Cambridge, MA: Cambridge University Press. 77-95.

35. Fukumoto, H., Kayano, T., Buse, J.B., Edwards, Y., Pilch, P.F., Bell, G.I., and Seino, S. 1989. Cloning and characterization of the major insulin-responsive glucose transporter expressed in human skeletal muscle and other insulin-responsive tissues. J. Biol. Chem. 264:7776-7779.

36. James, D.E., Strube, M., and Mueckler, M. 1989. Molecular cloning and characterization of an insulin-regulatable glucose transporter Nature (Lond) 338:83-87

37. Charron, M.J., and Kahn, B.B. 1990. Divergent molecular mechanisms for insulin-resistant glucose transport in muscle and adipose cells in vivo . J. Biol. Chem. 265:7994-8000.

38. Cushman, S.W., and Wardzala, L.J. 1980. Potential mechanism of insulin action on glucose transport in the isolated rat adipose cell. Apparent translocation of intracellular transport systems to the plasma membrane. J. Biol. Chem. 255:4758-4762.

39. Kono, T., Suzuki, K., Dansey, L.E., Robinson, F.W., and Blevins, T.L. 1981. Energy-dependent and protein synthesis-independent recycling of the insulin-sensitive glucose transport mechanism in fat cells. J. Biol. Chem. 256:6400-6407.

40. Rosholt, M.N., King, P.A., and Horton, E.S. 1994. High-fat diet reduces glucose transporter responses to both insulin and exercise. Am. J. Physiol. 266:R95-101.

41. King, P.A., Betts, J.J., Horton, E.D., and Horton, E.S. 1993. Exercise, unlike insulin, promotes glucose transporter translocation in obese Zucker rat muscle. Am. J. Physiol. 265:R447-452.

42. Wheeler, T.J. 1988. Translocation of glucose transporters in response to anoxia in heart. J. Biol. Chem. 263:19447-19454.

43. Karnieli, E., Chernow, B., Hissin, P.J., Simpson, I.A., and Foley, J.E. 1986. Insulin stimulates glucose transport in isolated human adipose cells through a translocation of intracellular glucose transporters to the plasma membrane: a preliminary report. Horm. Metab. Res. 18:867-868.

44. Clark, A.E., Holman, G.D., and Kozka, I.J. 1991. Determination of the rates of appearance and loss of glucose transporters at the cell surface of rat adipose cells. Biochem. J. 278 ( Pt 1):235-241.

45. Baron, A.D., Brechtel, G., Wallace, P., and Edelman, S.V. 1988. Rates and tissue sites of non-insulin- and insulin-mediated glucose uptake in humans. Am. J. Physiol. 255:E769-774.

46. Barnard, R.J., Lawani, L.O., Martin, D.A., Youngren, J.F., Singh, R., and Scheck, S.H. 1992. Effects of maturation and aging on the skeletal muscle glucose transport system. Am. J. Physiol. 262:E619-626.

47. Bourey, R.E., Koranyi, L., James, D.E., Mueckler, M., and Permutt, M.A. 1990. Effects of altered glucose homeostasis on glucose transporter expression in skeletal muscle of the rat. J. Clin. Invest. 86:542-547.

48. Klip, A., Ramlal, T., Bilan, P.J., Cartee, G.D., Gulve, E.A., and Holloszy, J.O. 1990. Recruitment of GLUT-4 glucose transporters by insulin in diabetic rat skeletal muscle. Biochem. Biophys. Res. Commun. 172:728-736.

49. Unger, R.H. 1991. Diabetic hyperglycemia: link to impaired glucose transport in pancreatic β cells. Science 251:1200-1205.

50. Thorens, B., Wu, Y.J., Leahy, J.L., and Weir, G.C. 1992. The loss of GLUT2 expression by glucose-unresponsive β cells of db/db mice is reversible and is induced by the diabetic environment. J. Clin. Invest. 90:77-85.

51. Kulkarni, R.N., Bruning, J.C., Winnay, J.N., Postic, C., Magnuson, M.A., and Kahn, C.R. 1999. Tissue-specific knockout of the insulin receptor in pancreatic β cells creates an insulin secretory defect similar to that in type 2 diabetes. Cell 96:329-339.

52. Kitamura, T., Kahn, C.R., and Accili, D. 2003. Insulin receptor knockout mice. Annu. Rev. Physiol. 65:313-332.

53. Braun, M., Ramracheya, R., and Rorsman, P. 2012. Autocrine regulation of insulin secretion. Diabetes Obes. Metab. 14 Suppl 3:143-151.

54. Baker, E.N., Blundell, T.L., Cutfield, J.F., Cutfield, S.M., Dodson, E.J., Dodson, G.G., Hodgkin, D.M., Hubbard, R.E., Isaacs, N.W., and Reynolds, C.D. 1988. The structure of 2Zn pig insulin crystals at 1.5 Å resolution. Philos. Trans. R. Soc. Lond. B Biol. Sci. 319:369-456.

55. Steiner, D.F. 1977. The Banting Memorial Lecture 1976. Insulin today. Diabetes 26:322-340.

56. Steiner, D.F. 1978. On the role of the proinsulin C-peptide. Diabetes 27 Suppl 1:145-148.

57. Dodson, G., and Steiner, D. 1998. The role of assembly in insulin's biosynthesis. Curr. Opin. Struct. Biol. 8:189-194.

58. Steiner, D.F. 2000. New aspects of proinsulin physiology and pathophysiology. J. Pediatr. Endocrinol. Metab. 13:229-239.

59. Seidah, N.G., and Prat, A. 2007. The proprotein convertases are potential targets in the treatment of dyslipidemia. J. Mol. Med. 85:685-696.

60. Yang, Y., Hua, Q.X., Liu, J., Shimizu, E.H., Choquette, M.H., Mackin, R.B., and Weiss, M.A. 2010. Solution structure of proinsulin: connecting domain flexibility and prohormone processing. J. Biol. Chem. 285:7847-7851.

61. Smith, B.J., Huang, K., Kong, G., Chan, S.J., Nakagawa, S., Menting, J.G., Hu, S.-Q., Whittaker, J., Steiner, D.F., Katsoyannis, P.G., et al. 2010. Structural resolution of a tandem hormone-binding element in the insulin receptor and its implications for design of peptide agonists. Proc. Natl. Acad. Sci. U. S. A. 107:6771-6776.

62. Menting, J.G., Whittaker, J., Margetts, M.B., Whittaker, L.J., Kong, G.K., Smith, B.J., Watson, C.J., Zakova, L., Kletvikova, E., Jiracek, J., et al. 2013. How insulin engages its primary binding site on the insulin receptor. Nature 493:241-245.

63. Abel, J.J. 1926. Crystalline Insulin. Proc. Natl. Acad. Sci. U. S. A. 12:132-136.

64. Murnaghan, J.H., and Talalay, P. 1967. John Jacob Abel and the crystallization of insulin. Perspect. Biol. Med. 10:334-380.

65. Jensen, H., and Evans, J., E. A. . 1935. Studies on crystalline insulin: xviii. the nature of the free amino groups in insulin and the isolation of phenylalanine and proline from crystalline insulin J. Biol. Chem. 108 1-9.

66. Ryle, A.P., Sanger, F., Smith, L.F., and Kitai, R. 1955. The disulphide bonds of insulin. Biochem. J. 60:541-556.

67. Sanger, F. 1959. Chemistry of insulin; determination of the structure of insulin opens the way to greater understanding of life processes. Science 129:1340-1344.

68. Chance, R.E., Ellis, R.M., and Bromer, W.W. 1968. Porcine proinsulin: characterization and amino acid sequence. Science 161:165-167.

69. Harfenist, E.J., and Craig, L.C. 1952. The molecular weight of insulin. J. Am. Chem. Soc. 74 J. Am. Chem. Soc.,.

70. Steiner, D.F., and Oyer, P.E. 1967. The biosynthesis of insulin and a probable precursor of insulin by a human islet cell adenoma. Proc. Natl. Acad. Sci. U. S. A. 57:473-480.

71. Steiner, D.F., Cunningham, D., Spigelman, L., and Aten, B. 1967. Insulin biosynthesis: evidence for a precursor. Science 157:697-700.

72. Steiner, D.F., Clark, J.L., Nolan, C., Rubenstein, A.H., Margoliash, E., Aten, B., and Oyer, P.E. 1969. Proinsulin and the biosynthesis of insulin. Recent Prog. Horm. Res. 25:207-282.

73. Clark, J.L., Cho, S., Rubenstein, A.H., and Steiner, D.F. 1969. Isolation of a proinsulin connecting peptide fragment (C-peptide) from bovine and human pancreas. Biochem. Biophys. Res. Commun. 35:456-461.

74. Brown, H., Sangar, F., and Kitai, R. 1955. The structure of pig and sheep insulins. Biochem. J. 60:555-565.

75. Chan, S.J., Keim, P., and Steiner, D.F. 1976. Cell-free synthesis of rat preproinsulins: characterization and partial amino acid sequence determination. Proc. Natl. Acad. Sci. U. S. A. 73:1964-1968.

76. Lomedico, P.T., Chan, S.J., Steiner, D.F., and Saunders, G.F. 1977. Immunological and chemical characterization of bovine preproinsulin. J. Biol. Chem. 252:7971-7978.

77. Steiner, D.F., Chan, S.J., Welsh, J.M., Nielsen, D., Michael, J., Tager, H.S., and Rubenstein, A.H. 1986. Models of peptide biosynthesis: the molecular and cellular basis of insulin production. Clin. Invest. Med. 9:328-336.

78. Sanders, S.L., and Schekman, R. 1992. Polypeptide translocation across the endoplasmic reticulum membrane. J. Biol. Chem. 267:13791-13794.

79. Kreil, G. 1981. Transfer of proteins across membranes. Annu. Rev. Biochem. 50:317-348.

80. Steiner, D.F., editor. 2001. The prohormone convertases and precursor processing in protein biosynthesis : Academic Press, New York. 163-198 pp.

81. Liu, M., Lara-Lemus, R., Shan, S.O., Wright, J., Haataja, L., Barbetti, F., Guo, H., Larkin, D., and Arvan, P. 2012. Impaired cleavage of preproinsulin signal peptide linked to autosomal-dominant diabetes. Diabetes 61:828-837.

82. Walter, P., Gilmore, R., and Blobel, G. 1984. Protein translocation across the endoplasmic reticulum. Cell 38:5-8.

83. Patzelt, C., Chan, S.J., Duguid, J., Hortin, G., Keim, P., Heinrikson, R.L., and Steiner, D.F. 1978. In Regulatory proteolytic enzymes and their inhibitors . S. Magnusson, M. Ottensen, B. Foltmann, K. Dano, and H. Neurath, editors. New York: Pergamon. p 69 Pergaman, New Yor.

84. Steiner, D.F., Clark, J.L., Nolan, C., Rubenstein, A.H., Margoliash, E., Melani, F., and Oyer, P.E., editors. 1970. The biosynthesis of insulin and some speculation regarding the pathogenesis of human diabetes . New York: John Wiley & Sons. 57 pp.

85. Steiner, D.F., Kemmler, W., Clark, J.L., Oyer, P.E., and Rubenstein, A.H., editors. 1972. The biosynthesis of insulin . Baltimore: Williams & Wilkins. 175-198 pp.

86. Frank, B.H., and Veros, A.J. 1968. Physical studies on proinsulin-association behavior and conformation in solution. Biochem. Biophys. Res. Commun. 32:155-160.

87. Rubenstein, A.H., Melani, F., Pilkis, S., and Steiner, D.F. 1969. Proinsulin. Secretion, metabolism, immunological and biological properties. Postgrad. Med. J. 45:Suppl:476-481.

88. Fullerton, W.W., Potter, R., and Low, B.W. 1970. Proinsulin: Crystallization and preliminary x-ray diffraction studies. Proc. Natl. Acad. Sci. U. S. A. 66:1213-1219.

89. Freychet, P., Brandenburg, D., and Wollmer, A. 1974. Receptor-binding assay of chemically modified insulins. Comparison with in vitro and in vivo bioassays. Diabetologia 10:1-5.

90. Orci, L., Ravazzola, M., Amherdt, M., Madsen, O., Vassalli, J.D., and Perrelet, A. 1985. Direct identification of prohormone conversion site in insulin-secreting cells. Cell 42:671-681.

91. Orci, L., Glick, B.S., and Rothman, J.E. 1986. A new type of coated vesicular carrier that appears not to contain clathrin: its possible role in protein transport within the Golgi stack. Cell 46:171-184.

92. Wattenberg, B.W., and Rothman, J.E. 1986. Multiple cytosolic components promote intra-Golgi protein transport. Resolution of a protein acting at a late stage, prior to membrane fusion. J. Biol. Chem. 261:2208-2213.

93. Nagamatsu, S., Bolaffi, J.L., and Grodsky, G.M. 1987. Direct effects of glucose on proinsulin synthesis and processing during desensitization. Endocrinology 120:1225-1231.

94. Kemmler, W., Peterson, J.D., and Steiner, D.F. 1971. Studies on the conversion of proinsulin to insulin. I. Conversion in vitro with trypsin and carboxypeptidase B. J. Biol. Chem. 246:6786-6791.

95. Nolan, C., Margoliash, E., Peterson, J.D., and Steiner, D.F. 1971. The structure of bovine proinsulin. J. Biol. Chem. 246:2780-2795.

96. Oyer, P.E., Cho, S., Peterson, J.D., and Steiner, D.F. 1971. Studies on human proinsulin. Isolation and amino acid sequence of the human pancreatic C-peptide. J. Biol. Chem. 246:1375-1386.

97. Busse, W.D., and Gattner, H.G. 1973. Selective cleavage of one disulfide bond in insulin: preparation and properties of insulin A7-B7-di-S-sulfonate. Hoppe Seylers Z Physiol. Chem. 354:147-155.

98. Tager, H.S., Emdin, S.O., Clark, J.L., and Steiner, D.F. 1973. Studies on the conversion of proinsulin to insulin. II. Evidence for a chymotrypsin-like cleavage in the connecting peptide region of insulin precursors in the rat. J. Biol. Chem. 248:3476-3482.

99. Davidson, H.W., Rhodes, C.J., and Hutton, J.C. 1988. Intraorganellar calcium and pH control proinsulin cleavage in the pancreatic β cell via two distinct site-specific endopeptidases. Nature 333:93-96.

100. Rhodes, C.J., Lincoln, B., and Shoelson, S.E. 1992. Preferential cleavage of des-31,32-proinsulin over intact proinsulin by the insulin secretory granule type II endopeptidase. Implication of a favored route for prohormone processing. J. Biol. Chem. 267:22719-22727.

101. Julius, D., Brake, A., Blair, L., Kunisawa, R., and Thorner, J. 1984. Isolation of the putative structural gene for the lysine-arginine-cleaving endopeptidase required for processing of yeast prepro-α-factor. Cell 37:1075-1089.

102. Fuller, R.S., Sterne, R.E., and Thorner, J. 1988. Enzymes required for yeast prohormone processing. Annu. Rev. Physiol. 50:345-362.

103. Mizuno, K., Nakamura, T., Ohshima, T., Tanaka, S., and Matsuo, H. 1988. Yeast KEX2 genes encodes an endopeptidase homologous to subtilisin-like serine proteases. Biochem. Biophys. Res. Commun. 156:246-254.

104. Smeekens, S.P., and Steiner, D.F. 1990. Identification of a human insulinoma cDNA encoding a novel mammalian protein structurally related to the yeast dibasic processing protease Kex2. J. Biol. Chem. 265:2997-3000.

105. Smeekens, S.P., Avruch, A.S., LaMendola, J., Chan, S.J., and Steiner, D.F. 1991. Identification of a cDNA encoding a second putative prohormone convertase related to PC2 in AtT20 cells and islets of Langerhans. Proc. Natl. Acad. Sci. U. S. A. 88:340-344.

106. Seidah, N.G., Marcinkiewicz, M., Benjannet, S., Gaspar, L., Beaubien, G., Mattei, M.G., Lazure, C., Mbikay, M., and Chretien, M. 1991. Cloning and primary sequence of a mouse candidate prohormone convertase PC1 homologous to PC2, Furin, and Kex2: distinct chromosomal localization and messenger RNA distribution in brain and pituitary compared to PC2. Mol. Endocrinol. 5:111-122.

107. Shennan, K.I., Smeekens, S.P., Steiner, D.F., and Docherty, K. 1991. Characterization of PC2, a mammalian Kex2 homologue, following expression of the cDNA in microinjected Xenopus oocytes. FEBS Lett. 284:277-280.

108. Bailyes, E.M., Shennan, K.I., Seal, A.J., Smeekens, S.P., Steiner, D.F., Hutton, J.C., and Docherty, K. 1992. A member of the eukaryotic subtilisin family (PC3) has the enzymic properties of the type 1 proinsulin-converting endopeptidase. Biochem. J. 285 (Pt 2):391-394.

109. Zhou, A., Webb, G., Zhu, X., and Steiner, D.F. 1999. Proteolytic processing in the secretory pathway. J. Biol. Chem. 274:20745-20748.

110. Seidah, N.G., and Prat, A. 2012. The biology and therapeutic targeting of the proprotein convertases. Nat Rev Drug Discov 11:367-383.

111. Smeekens, S.P., Montag, A.G., Thomas, G., Albiges-Rizo, C., Carroll, R., Benig, M., Phillips, L.A., Martin, S., Ohagi, S., Gardner, P., et al. 1992. Proinsulin processing by the subtilisin-related proprotein convertases furin, PC2, and PC3. Proc. Natl. Acad. Sci. U. S. A. 89:8822-8826.

112. Tanaka, S., Kurabuchi, S., Mochida, H., Kato, T., Takahashi, S., Watanabe, T., and Nakayama, K. 1996. Immunocytochemical localization of prohormone convertases PC1/PC3 and PC2 in rat pancreatic islets. Arch. Histol. Cytol. 59:261-271.

113. Skidgel, R.A. 1988. Basic carboxypeptidases: regulators of peptide hormone activity. Trends Pharmacol. Sci. 9:299-304.

114. Fricker, L.D., Evans, C.J., Esch, F.S., and Herbert, E. 1986. Cloning and sequence analysis of cDNA for bovine carboxypeptidase E. Nature 323:461-464.

115. Furuta, M., Carroll, R., Martin, S., Swift, H.H., Ravazzola, M., Orci, L., and Steiner, D.F. 1998. Incomplete processing of proinsulin to insulin accompanied by elevation of Des-31,32 proinsulin intermediates in islets of mice lacking active PC2. J. Biol. Chem. 273:3431-3437.

116. Steiner, D.F. 2011. Adventures with insulin in the islets of Langerhans. J Biol Chem 286:17399-17421.

117. Zhu, X., Zhou, A., Dey, A., Norrbom, C., Carroll, R., Zhang, C., Laurent, V., Lindberg, I., Ugleholdt, R., Holst, J.J., et al. 2002. Disruption of PC1/3 expression in mice causes dwarfism and multiple neuroendocrine peptide processing defects. Proc. Natl. Acad. Sci. U. S. A. 99:10293-10298.

118. Furuta, M., Yano, H., Zhou, A., Rouille, Y., Holst, J.J., Carroll, R.J., Ravazzola, M., Orci, L., Furuta, H., and Steiner, D.F. 1997. Defective prohormone processing and altered pancreatic islet morphology in mice lacking active SPC2. Proc. Natl. Acad. Sci. U. S. A. 94:6646-6651.

119. Muller, L., Zhu, X., and Lindberg, I. 1997. Mechanism of the facilitation of PC2 maturation by 7B2: involvement in ProPC2 transport and activation but not folding. J Cell Biol 139:625-638.

120. Westphal, C.H., Muller, L., Zhou, A., Zhu, X., Bonner-Weir, S., Schambelan, M., Steiner, D.F., Lindberg, I., and Leder, P. 1999. The neuroendocrine protein 7B2 is required for peptide hormone processing in vivo and provides a novel mechanism for pituitary Cushing's disease. Cell 96:689-700.

121. Greider, M.H., Howell, S.L., and Lacy, P.E. 1969. Isolation and properties of secretory granules from rat islets of Langerhans. II. Ultrastructure of the β granule. J. Cell Biol. 41:162-166.

122. Lange, R.H. 1974. Crystalline islet B-granules in the grass snake (Natrix natrix (L.)): tilting experiments in the electron microscope. J. Ultrastruct. Res. 46:301-307.

123. Michael, J., Carroll, R., Swift, H.H., and Steiner, D.F. 1987. Studies on the molecular organization of rat insulin secretory granules. J. Biol. Chem. 262:16531-16535.

124. Lemaire, K., Ravier, M.A., Schraenen, A., Creemers, J.W., Van de Plas, R., Granvik, M., Van Lommel, L., Waelkens, E., Chimienti, F., Rutter, G.A., et al. 2009. Insulin crystallization depends on zinc transporter ZnT8 expression, but is not required for normal glucose homeostasis in mice. Proc. Natl. Acad. Sci. U. S. A. 106:14872-14877.

125. Steiner, D.F. 1973. Cocrystallization of proinsulin and insulin. Nature 243:528-530.

126. Frank, B.H., and Veros, A.J. 1970. Interaction of zinc with proinsulin. Biochem. Biophys. Res. Commun. 38:284-289.

127. Grant, P.T., Coombs, T.L., and Frank, B.H. 1972. Differences in the nature of the interaction of insulin and proinsulin with zinc. Biochem. J. 126:433-440.

128. Howell, S.L., Tyhurst, M., Duvefelt, H., Andersson, A., and Hellerstrom, C. 1978. Role of zinc and calcium in the formation and storage of insulin in the pancreatic β-cell. Cell Tissue Res. 188:107-118.

129. Kemmler, W., Steiner, D.F., and Borg, J. 1973. Studies on the conversion of proinsulin to insulin. 3. Studies in vitro with a crude secretion granule fraction isolated from rat islets of Langerhans. J. Biol. Chem. 248:4544-4551.

130. Steiner, D.F., and Rubenstein, A. 1973. In Proceedings of the 8th Midwest Conference on Endocrinology and Metabolism . F.M. Matschinsky, R. Fertel, J. Kotler-Brajtburg, S. Stillings, J. Ellerman, F. Raybaud, J.H. Thurston, R.P. Breitenbach, and X.J. Mussachia, editors: University of Missouri. 43-59.

131. Prentki, M., and Matschinsky, F.M. 1987. Ca 2+ , cAMP, and phospholipid-derived messengers in coupling mechanisms of insulin secretion. Physiol Rev 67:1185-1248.

132. Hughes, S.J., and Ashcroft, S.H.J. 1992. In Nutrient regulation of insulin secretion . P.R. Flatt, editor. Colchester, England: Portland Press. 271-289

133. Nielsen, D.A., Welsh, M., Casadaban, M.J., and Steiner, D.F. 1985. Control of insulin gene expression in pancreatic β-cells and in an insulin-producing cell line, RIN-5F cells. I. Effects of glucose and cyclic AMP on the transcription of insulin mRNA. J. Biol. Chem. 260:13585-13589.

134. Welsh, M., Nielsen, D.A., MacKrell, A.J., and Steiner, D.F. 1985. Control of insulin gene expression in pancreatic β-cells and in an insulin-producing cell line, RIN-5F cells. II. Regulation of insulin mRNA stability. J. Biol. Chem. 260:13590-13594.

135. Rabinovitch, A., Blondel, B., Murray, T., and Mintz, D.H. 1980. Cyclic adenosine-3',5'-monophosphate stimulates islet B cell replication in neonatal rat pancreatic monolayer cultures. J. Clin. Invest. 66:1065-1071.

136. Swenne, I. 1982. Effects of cyclic AMP on DNA replication and protein biosynthesis in fetal rat islets of Langerhans maintained in tissue culture. Biosci. Rep. 2:867-876.

137. Tanese, T., Lazarus, N.R., Devrim, S., and Recant, L. 1970. Synthesis and release of proinsulin and insulin by isolated rat islets of Langerhans. J. Clin. Invest. 49:1394-1404.

138. Rhodes, C.J., and Halban, P.A. 1987. Newly synthesized proinsulin/insulin and stored insulin are released from pancreatic B cells predominantly via a regulated, rather than a constitutive, pathway. J. Cell Biol. 105:145-153.

139. Kelly, R.B. 1985. Pathways of protein secretion in eukaryotes. Science 230:25-32.

140. Valverde, I., Garcia-Morales, P., Ghiglione, M., and Malaisse, W.J. 1983. The stimulus-secretion coupling of glucose-induced insulin release. LIII. Calcium-dependency of the cyclic AMP response to nutrient secretagogues. Horm. Metab. Res. 15:62-68.

141. Holst, J.J., Orskov, C., Nielsen, O.V., and Schwartz, T.W. 1987. Truncated glucagon-like peptide I, an insulin-releasing hormone from the distal gut. FEBS Lett. 211:169-174.

142. Rajan, S., Torres, J., Thompson, M.S., and Philipson, L.H. 2012. SUMO downregulates GLP-1-stimulated cAMP generation and insulin secretion. Am. J. Physiol. Endocrinol. Metab. 302:E714-723.

143. Howell, S.L., Jones, P.M., and Persaud, S.J. 1994. Regulation of insulin secretion: the role of second messengers. Diabetologia 37 Suppl 2:S30-35.

144. Berggren, P.O., Rorsman, P., and Efendic, S. 1992. In Nutrient regulation of insulin secretion . P.R. Flatt, editor. Colchester, England: Portland Press. 289-318

145. Persaud, S.J., Jones, P.M., and Howell, S.L. 1989. Effects of Bordetella pertussis toxin on catecholamine inhibition of insulin release from intact and electrically permeabilized rat islets. Biochem. J. 258:669-675.

146. Rorsman, P., and Braun, M. 2013. Regulation of insulin secretion in human pancreatic islets. Annu. Rev. Physiol. 75:155-179.

147. Fridlyand, L.E., and Philipson, L.H. 2011. Coupling of metabolic, second messenger pathways and insulin granule dynamics in pancreatic b -cells: a computational analysis. Prog. Biophys. Mol. Biol. 107:293-303.

148. Fridlyand, L.E., Jacobson, D.A., and Philipson, L.H. 2013. Ion channels and regulation of insulin secretion in human b -cells: a computational systems analysis. Islets 5:1-15.

149. Tabei, S.M., Burov, S., Kim, H.Y., Kuznetsov, A., Huynh, T., Jureller, J., Philipson, L.H., Dinner, A.R., and Scherer, N.F. 2013. Intracellular transport of insulin granules is a subordinated random walk. Proc. Natl. Acad. Sci. U. S. A. 110:4911-4916.

150. Polonsky, K.S. 1995. Lilly Lecture 1994. The β-cell in diabetes: from molecular genetics to clinical research. Diabetes 44:705-717.

151. Taylor, S.I., Accili, D., and Imai, Y. 1994. Insulin resistance or insulin deficiency. Which is the primary cause of NIDDM? Diabetes 43:735-740.

152. Turner, R.C., Hattersley, A.T., Shaw, J.T., and Levy, J.C. 1995. Type II diabetes: clinical aspects of molecular biological studies. Diabetes 44:1-10.

153. Dean, P.M., and Matthews, E.K. 1968. Electrical activity in pancreatic islet cells. Nature 219:389-390.

154. Dean, P.M., and Matthews, E.K. 1970. Glucose-induced electrical activity in pancreatic islet cells. J. Physiol. 210:255-264.

155. Ashcroft, F.M., Harrison, D.E., and Ashcroft, S.J. 1984. Glucose induces closure of single potassium channels in isolated rat pancreatic β-cells. Nature 312:446-448.

156. Cook, D.L., and Hales, C.N. 1984. Intracellular ATP directly blocks K + channels in pancreatic B-cells. Nature 311:271-273.

157. Cook, D.L., Ikeuchi, M., and Fujimoto, W.Y. 1984. Lowering of pHi inhibits Ca2+-activated K+ channels in pancreatic B-cells. Nature 311:269-271.

158. McTaggart, J.S., Clark, R.H., and Ashcroft, F.M. 2010. The role of the KATP channel in glucose homeostasis in health and disease: more than meets the islet. J. Physiol. 588:3201-3209.

159. Doyle, D.A., Morais Cabral, J., Pfuetzner, R.A., Kuo, A., Gulbis, J.M., Cohen, S.L., Chait, B.T., and MacKinnon, R. 1998. The structure of the potassium channel: molecular basis of K+ conduction and selectivity. Science 280:69-77.

160. Nishida, M., and MacKinnon, R. 2002. Structural basis of inward rectification: cytoplasmic pore of the G protein-gated inward rectifier GIRK1 at 1.8 A resolution. Cell 111:957-965.

161. Ashcroft, F.M., and Rorsman, P. 1990. ATP-sensitive K + channels: a link between B-cell metabolism and insulin secretion. Biochem. Soc. Trans. 18:109-111.

162. Seino, S., Iwanaga, T., Nagashima, K., and Miki, T. 2000. Diverse roles of K ATP channels learned from K ir 6.2 genetically engineered mice. Diabetes 49:311-318.

163. Huopio, H., Shyng, S.L., Otonkoski, T., and Nichols, C.G. 2002. K ATP channels and insulin secretion disorders. Am. J. Physiol. Endocrinol. Metab. 283:E207-216.

164. Stoffel, M., Espinosa, R., 3rd, Powell, K.L., Philipson, L.H., Le Beau, M.M., and Bell, G.I. 1994. Human G-protein-coupled inwardly rectifying potassium channel (GIRK1) gene (KCNJ3): localization to chromosome 2 and identification of a simple tandem repeat polymorphism. Genomics 21:254-256.

165. Reuveny, E., Slesinger, P.A., Inglese, J., Morales, J.M., Iniguez-Lluhi, J.A., Lefkowitz, R.J., Bourne, H.R., Jan, Y.N., and Jan, L.Y. 1994. Activation of the cloned muscarinic potassium channel by G protein β γ subunits. Nature 370:143-146.

166. Philipson, L.H., Kuznetsov, A., Toth, P.T., Murphy, J.F., Szabo, G., Ma, G.H., and Miller, R.J. 1995. Functional expression of an epitope-tagged G protein-coupled K + channel (GIRK1). J. Biol. Chem. 270:14604-14610.

167. Aguilar-Bryan, L., and Bryan, J. 1999. Molecular biology of adenosine triphosphate-sensitive potassium channels. Endocr. Rev. 20:101-135.

168. Katz, B., and Miledi, R. 1965. The effect of calcium on acetylcholine release from motor nerve terminals. Proc. R. Soc. Lond. B Biol. Sci. 161:496-503.

169. Douglas, W.W. 1968. Stimulus-secretion coupling: the concept and clues from chromaffin and other cells. Br. J. Pharmacol. 34:451-474.

170. Wollheim, C.B., and Sharp, G.W. 1981. Regulation of insulin release by calcium. Physiol. Rev. 61:914-973.

171. Rorsman, P., Braun, M., and Zhang, Q. 2012. Regulation of calcium in pancreatic a - and b -cells in health and disease. Cell Calcium 51:300-308.

172. Tsien, R.W., Ellinor, P.T., and Horne, W.A. 1991. Molecular diversity of voltage-dependent Ca 2+ channels. Trends Pharmacol. Sci. 12:349-354.

173. Namkung, Y., Skrypnyk, N., Jeong, M.J., Lee, T., Lee, M.S., Kim, H.L., Chin, H., Suh, P.G., Kim, S.S., and Shin, H.S. 2001. Requirement for the L-type Ca 2+ channel α 1D subunit in postnatal pancreatic β cell generation. J. Clin. Invest. 108:1015-1022.

174. Braun, M., Ramracheya, R., Bengtsson, M., Zhang, Q., Karanauskaite, J., Partridge, C., Johnson, P.R., and Rorsman, P. 2008. Voltage-gated ion channels in human pancreatic beta-cells: electrophysiological characterization and role in insulin secretion. Diabetes 57:1618-1628.

175. Straub, S.G., and Sharp, G.W. 2012. Evolving insights regarding mechanisms for the inhibition of insulin release by norepinephrine and heterotrimeric G proteins. Am. J. Physiol. Cell Physiol. 302:C1687-1698.

176. Sutton, R.B., Fasshauer, D., Jahn, R., and Brunger, A.T. 1998. Crystal structure of a SNARE complex involved in synaptic exocytosis at 2.4 Å resolution. Nature 395:347-353.

177. Regazzi, R., Ravazzola, M., Iezzi, M., Lang, J., Zahraoui, A., Andereggen, E., Morel, P., Takai, Y., and Wollheim, C.B. 1996. Expression, localization and functional role of small GTPases of the Rab3 family in insulin-secreting cells. J. Cell Sci. 109 ( Pt 9):2265-2273.

178. Mizuta, M., Kurose, T., Miki, T., Shoji-Kasai, Y., Takahashi, M., Seino, S., and Matsukura, S. 1997. Localization and functional role of synaptotagmin III in insulin secretory vesicles in pancreatic β-cells. Diabetes 46:2002-2006.

179. Peking, I.R.G. 1971. Insulin's crystal structure at 2.5A resolution. Peking Rev. 40:11-16.

180. Blundell, T.L., Dodson, G.G., Hodgkin, D.C., and Mercola, D.A. 1972. Insulin: the structure in the crystal and its reflection in chemistry and biology. Adv. Protein Chem. 26:279-402.

181. Bentley, G., Dodson, E., Dodson, G., Hodgkin, D., and Mercola, D. 1976. Structure of insulin in 4-zinc insulin. Nature 261:166-168.

182. Derewenda, U., Derewenda, Z., Dodson, E.J., Dodson, G.G., Reynolds, C.D., Smith, G.D., Sparks, C., and Swenson, D. 1989. Phenol stabilizes more helix in a new symmetrical zinc insulin hexamer. Nature 338:594-596.

183. Smith, G.D., Pangborn, W.A., and Blessing, R.H. 2003. The structure of T6 human insulin at 1.0 A resolution. Acta Crystallogr. D Biol. Crystallogr. 59:474-482.

184. Hua, Q.X., Shoelson, S.E., Kochoyan, M., and Weiss, M.A. 1991. Receptor binding redefined by a structural switch in a mutant human insulin. Nature 354:238-241.

185. Knegtel, R.M., Boelens, R., Ganadu, M.L., and Kaptein, R. 1991. The solution structure of a monomeric insulin. A two-dimensional 1 H-NMR study of des -(B26-B30)-insulin in combination with distance geometry and restrained molecular dynamics. Eur. J. Biochem. 202:447-458.

186. Jørgensen, A.M., Kristensen, S.M., Led, J.J., and Balschmidt, P. 1992. Three-dimensional solution structure of an insulin dimer. A study of the B9(Asp) mutant of human insulin using nuclear magnetic resonance, distance geometry and restrained molecular dynamics. J. Mol. Biol. 227:1146-1163.

187. Olsen, H.B., Ludvigsen, S., and Kaarsholm, N.C. 1996. Solution structure of an engineered insulin monomer at neutral pH. Biochemistry 35:8836-8845.

188. Hua, Q.X., Hu, S.Q., Frank, B.H., Jia, W., Chu, Y.C., Wang, S.H., Burke, G.T., Katsoyannis, P.G., and Weiss, M.A. 1996. Mapping the functional surface of insulin by design: structure and function of a novel A-chain analogue. J. Mol. Biol. 264:390-403.

189. Sorensen, M.D., Bjorn, S., Norris, K., Olsen, O., Petersen, L., James, T.L., and Led, J.J. 1997. Solution structure and backbone dynamics of the human α3-chain type VI collagen C-terminal Kunitz domain. Biochemistry 36:10439-10450.

190. Ludvigsen, S., Olsen, H.B., and Kaarsholm, N.C. 1998. A structural switch in a mutant insulin exposes key residues for receptor binding. J. Mol. Biol. 279:1-7.

191. Xu, B., Hua, Q.X., Nakagawa, S.H., Jia, W., Chu, Y.C., Katsoyannis, P.G., and Weiss, M.A. 2002. Chiral mutagensis of insulin's hidden receptor-binding surface: structure of an allo -isoleucine A2 analogue. J. Mol. Biol. 316:435-441.

192. Hua, Q.X., Xu, B., Huang, K., Hu, S.Q., Nakagawa, S., Jia, W., Wang, S., Whittaker, J., Katsoyannis, P.G., and Weiss, M.A. 2009. Enhancing the activity of insulin by stereospecific unfolding. Conformational life cycle of insulin and its evolutionary origins. J. Biol. Chem. 284:14586-14596.

193. Žáková, L., Kletvíková, E., Veverka, V., Lepsík, M., Watson, C.J., Turkenburg, J.P., Jirácek, J., and Brzozowski, A.M. 2013. Structural integrity of the B24 site in human insulin is important for hormone functionality. J. Biol. Chem. 288:10230-10240.

194. Liang, D.C., Chang, W.R., and Wan, Z.L. 1994. A proposed interaction model of the insulin molecule with its receptor. Biophys. Chem. 50:63-71.

195. Scott, D.A. 1934. Crystalline insulin. Biochem. J. 28:1592-1602 1591.

196. Scott, D.A., and Fisher, A.M. 1935. Crystalline insulin. Biochem. J. 29:1048-1054.

197. Brader, M.L., and Dunn, M.F. 1991. Insulin hexamers: new conformations and applications. Trends Biochem. Sci. 16:341-345.

198. Badger, J., Harris, M.R., Reynolds, C.D., Evans, A.C., Dodson, E.J., Dodson, G.G., and North, A.C. 1991. Structure of the pig insulin dimer in the cubic crystal. Acta Crystallogr. B Struct. Sci. 47:127-136.

199. Badger, J. 1992. Flexibility in crystalline insulins. Biophys. J. 61:816-819.

200. Bi, R.C., Dauter, Z., Dodson, E., Dodson, G., Giordano, F., and Reynolds, C. 1984. Insulin structure as a modified and monomeric molecule. Biopolymers 23:391-395.

201. Liang, D.C., Stuart, D., B., D.J., Todd, R., You, J.M., and Luo, M.Z. 1985. X-ray studies of des-pentapeptide (B26-B30) insulin at 2.4 Å resolution: The DPI molecule and its structural relationship with insulin. Sci. Sin. 28:472-484.

202. Bao, S.J., Xie, D.L., Zhang, J.P., Chang, W.R., and Liang, D.C. 1997. Crystal structure of des heptapeptide(B24-B30)insulin at 1.6 Å resolution: implications for receptor binding. Proc. Natl. Acad. Sci. U. S. A. 94:2975-2980.

203. Wan, Z., Huang, K., Whittaker, J., and Weiss, M.A. 2008. The structure of a mutant insulin uncouples receptor binding from protein allostery. An electrostatic block to the TR transition. J. Biol. Chem. 283:21198-21210.

204. Derewenda, U., Derewenda, Z.S., Dodson, G.G., and Hubbard, R.E. 1990. Insulin structure. In Insulin—the handbook of experimental pharmacology . P. Cuatrecasas, and S. Jacobs, editors. Berlin: Springer-Verlag. 23-39.

205. Jacoby, E., Hua, Q.X., Stern, A.S., Frank, B.H., and Weiss, M.A. 1996. Structure and dynamics of a protein assembly. 1 H-NMR studies of the 36 kDa R 6 insulin hexamer. J. Mol. Biol. 258:136-157.

206. Chang, X., Jorgensen, A.M., Bardrum, P., and Led, J.J. 1997. Solution structures of the R6 human insulin hexamer. Biochemistry 36:9409-9422.

207. Jørgensen, L.N., Dideriksen, L.H., and Drejer, K. 1992. Carcinogenic effect of the human insulin analog B-10 Asp in female rats. Diabetologia 35 Suppl 1:A3.

208. Hua, Q.X., and Weiss, M.A. 1990. Toward the solution structure of human insulin: sequential 2D 1 H NMR assignment of a des -pentapeptide analogue and comparison with crystal structure. Biochemistry 29:10545-10555.

209. Derewenda, U., Derewenda, Z., Dodson, E.J., Dodson, G.G., Bing, X., and Markussen, J. 1991. X-ray analysis of the single chain B29-A1 peptide-linked insulin molecule. A completely inactive analogue. J. Mol. Biol. 220:425-433.

210. Smith, G.D., and Dodson, G.G. 1992. Structure of a rhombohedral R6 insulin/phenol complex. Proteins 14:401-408.

211. Dodson, E.J., Dodson, G.G., Lewitova, A., and Sabesan, M. 1978. Zinc-free cubic pig insulin: crystallization and structure determination. J. Mol. Biol. 125:387-396.

212. Cutfield, J.F., Cutfield, S.M., Dodson, E.J., Dodson, G.G., Emdin, S.F., and Reynolds, C.D. 1979. Structure and biological activity of hagfish insulin. J. Mol. Biol. 132:85-100.

213. Nakagawa, S.H., Zhao, M., Hua, Q.X., Hu, S.Q., Wan, Z.L., Jia, W., and Weiss, M.A. 2005. Chiral mutagenesis of insulin. Foldability and function are inversely regulated by a stereospecific switch in the B chain. Biochemistry 44:4984-4999.

214. Hua, Q.X., Nakagawa, S.H., Hu, S.Q., Jia, W., Wang, S., and Weiss, M.A. 2006. Toward the active conformation of insulin. Stereospecific modulation of a structural switch in the B chain. J. Biol. Chem. 281:24900-24909.

215. Nakagawa, S.H., Hua, Q.X., Hu, S.Q., Jia, W., Wang, S., Katsoyannis, P.G., and Weiss, M.A. 2006. Chiral mutagenesis of insulin. Contribution of the B20-B23 β-turn to activity and stability. J. Biol. Chem. 281:22386-22396.

216. Weiss, M.A. 2009. Proinsulin and the genetics of diabetes mellitus. J. Biol. Chem. 284:19159-19163.

217. Frank, B.H., Pekar, A.H., and Veros, A.J. 1972. Insulin and proinsulin conformation in solution. Diabetes 21:486-491.

218. Goldman, J., and Carpenter, F.H. 1974. Zinc binding, circular dichroism, and equilibrium sedimentation studies on insulin (bovine) and several of its derivatives. Biochemistry 13:4566-4574.

219. Milthorpe, B.K., Nichol, L.W., and Jeffrey, P.D. 1977. The polymerization pattern of zinc(II)-insulin at pH 7.0. Biochim. Biophys. Acta 495:195-202.

220. McKern, N.M., Lawrence, M.C., Streltsov, V.A., Lou, M.Z., Adams, T.E., Lovrecz, G.O., Elleman, T.C., Richards, K.M., Bentley, J.D., Pilling, P.A., et al. 2006. Structure of the insulin receptor ectodomain reveals a folded-over conformation. Nature 443:218-221.

221. Kruger, P., Strassburger, W., Wollmer, A., van Gunsteren, W.F., and Dodson, G.G. 1987. The simulated dynamics of the insulin monomer and their relationship to the molecule's structure. Eur. Biophys. J. 14:449-459.

222. Wodak, S.J., Alard, P., Delhaise, P., and Renneboog-Squilbin, C. 1985. Simulation of conformational changes in 2 Zn insulin. J. Mol. Biol. 181:317-322.

223. Weiss, M.A., Hua, Q.X., Lynch, C.S., Frank, B.H., and Shoelson, S.E. 1991. Heteronuclear 2D NMR studies of an engineered insulin monomer: assignment and characterization of the receptor-binding surface by selective 2 H and 13 C labeling with application to protein design. Biochemistry 30:7373-7389.

224. Tidor, B., and Karplus, M. 1994. The contribution of vibrational entropy to molecular association. The dimerization of insulin. J. Mol. Biol. 238:405-414.

225. Rupley, J.A. 1967. The binding and cleavage by lysozyme of N-acetylglucosamine oligosaccharides. Proc. R. Soc. Lond. B Biol. Sci. 167:416-428.

226. Jeffrey, P.D., Milthorpe, B.K., and Nichol, L.W. 1976. Polymerization pattern of insulin at pH 7.0. Biochemistry 15:4660-4665.

227. Pocker, Y., and Biswas, S.B. 1981. Self-association of insulin and the role of hydrophobic bonding: a thermodynamic model of insulin dimerization. Biochemistry 20:4354-4361.

228. Cutfield, J.F., Cutfield, S.M., Dodson, E.J., dodson, G.G., Reynolds, C.D., and Vallely, D. 1981. In Structural studies on molecules of biological interest . G.G. Dodson, J. Glusker, and D. Sayre, editors. Oxford: Oxford University Press. 527-546.

229. Chothia, C., Lesk, A.M., Dodson, G.G., and Hodgkin, D.C. 1983. Transmission of conformational change in insulin. Nature 302:500-505.

230. Ciszak, E., and Smith, G.D. 1994. Crystallographic evidence for dual coordination around zinc in the T 3 R 3 human insulin hexamer. Biochemistry 33:1512-1517.

231. Schlichtkrull, J. 1958. Insulin Crystals: chemical and biological studies on insulin crystals and insulin zinc suspensions. Copenhagen, Munksgaard: Kobenhavns universitet. 139.

232. Roy, M., Brader, M.L., Lee, R.W., Kaarsholm, N.C., Hansen, J.F., and Dunn, M.F. 1989. Spectroscopic signatures of the T to R conformational transition in the insulin hexamer. J. Biol. Chem. 264:19081-19085.

233. Wollmer, A., Rannefeld, B., Johansen, B.R., Hejnaes, K.R., Balschmidt, P., and Hansen, F.B. 1987. Phenol-promoted structural transformation of insulin in solution. Biol. Chem. Hoppe Seyler 368:903-911.

234. Coffman, F.D., and Dunn, M.F. 1988. Insulin-metal ion interactions: the binding of divalent cations to insulin hexamers and tetramers and the assembly of insulin hexamers. Biochemistry 27:6179-6187.

235. Wood, S.P., Blundell, T.L., Wollmer, A., Lazarus, N.R., and Neville, R.W. 1975. The relation of conformation and association of insulin to receptor binding; a-ray and circular-dichroism studies on bovine and hystricomorph insulins. Eur. J. Biochem. 55:531-542.

236. Renscheidt, H., Strassburger, W., Glatter, U., Wollmer, A., Dodson, G.G., and Mercola, D.A. 1984. A solution equivalent of the 2Zn→4Zn transformation of insulin in the crystal. Eur. J. Biochem. 142:7-14.

237. Williamson, K.L., and Williams, R.J. 1979. Conformational analysis by nuclear magnetic resonance: insulin. Biochemistry 18:5966-5972.

238. Ramesh, V., and Bradbury, J.H. 1986. 1 H-NMR studies of insulin. Reversible transformation of 2Zn→4Zn insulin hexamer. Int. J. Pept. Protein Res. 28:146-153.

239. Brange, J., editor. 1987. Galenics of Insulin: The Physico-chemical and Pharmaceutical Aspects of Insulin and Insulin Preparations . Berlin: Springer Berlin Heidelberg. 1-103 pp.

240. Brange, J., and Langkjoer, L. 1993. Insulin structure and stability. Pharm. Biotechnol. 5:315-350.

241. Weiss, M.A. 2013. Design of ultra-stable insulin analogues for the developing world. Journal of Health Specialties 1:59-70.

242. Ciszak, E., Beals, J.M., Frank, B.H., Baker, J.C., Carter, N.D., and Smith, G.D. 1995. Role of C-terminal B-chain residues in insulin assembly: the structure of hexameric Lys B28 Pro B29 -human insulin. Structure (Lond.) 3:615-622.

243. Howell, S.L., Young, D.A., and Lacy, P.E. 1969. Isolation and properties of secretory granules from rat islets of Langerhans. 3. Studies of the stability of the isolated b granules. J. Cell Biol. 41:167-176.

244. Howell, S.L., Kostianovsky, M., and Lacy, P.E. 1969. b granule formation in isolated islets of Langerhans: a study by electron microscopic radioautography. J. Cell Biol. 42:695-705.

245. Katsoyannis, P.G. 1966. Synthesis of insulin. Science 154:1509-1514.

246. De Meyts, P., van Obberghen, E., and Roth, J. 1978. Mapping of the residues responsible for the negative cooperativity of the receptor-binding region of insulin. Nature 273:504-509.

247. Kristensen, C., Kjeldsen, T., Wiberg, F.C., Schaffer, L., Hach, M., Havelund, S., Bass, J., Steiner, D.F., and Andersen, A.S. 1997. Alanine scanning mutagenesis of insulin. J. Biol. Chem. 272:12978-12983.

248. Liang, D.C., Chang, W.R., Zhang, J.P., and Wan, Z.L. 1992. The possible mechanism of binding interaction of insulin molecule with its receptor. Sci. China B 35:547-557.

249. Shoelson, S.E., Lee, J., Lynch, C.S., Backer, J.M., and Pilch, P.F. 1993. Bpa B25 insulins. Photoactivatable analogues that quantitatively cross-link, radiolabel, and activate the insulin receptor. J. Biol. Chem. 268:4085-4091.

250. Kurose, T., Pashmforoush, M., Yoshimasa, Y., Carroll, R., Schwartz, G.P., Burke, G.T., Katsoyannis, P.G., and Steiner, D.F. 1994. Cross-linking of a B25 azidophenylalanine insulin derivative to the carboxyl-terminal region of the α-subunit of the insulin receptor. Identification of a new insulin-binding domain in the insulin receptor. J. Biol. Chem. 269:29190-29197.

251. Xu, B., Hu, S.Q., Chu, Y.C., Wang, S., Wang, R.Y., Nakagawa, S.H., Katsoyannis, P.G., and Weiss, M.A. 2004. Diabetes-associated mutations in insulin identify invariant receptor contacts. Diabetes 53:1599-1602.

252. Xu, B., Huang, K., Chu, Y.C., Hu, S.Q., Nakagawa, S., Wang, S., Wang, R.Y., Whittaker, J., Katsoyannis, P.G., and Weiss, M.A. 2009. Decoding the cryptic active conformation of a protein by synthetic photoscanning: Insulin inserts a detachable arm between receptor domains. J. Biol. Chem. 284:14597-14608.

253. Tager, H.S. 1995. Diabetes: Clinical Science in Practice . Cambridge: Cambridge University Press. 492 pp.

254. Pullen, R.A., Lindsay, D.G., Wood, S.P., Tickle, I.J., Blundell, T.L., Wollmer, A., Krail, G., Brandenburg, D., Zahn, H., Gliemann, J., et al. 1976. Receptor-binding region of insulin. Nature 259:369-373.

255. Kwok, S.C., Steiner, D.F., Rubenstein, A.H., and Tager, H.S. 1983. Identification of a point mutation in the human insulin gene giving rise to a structurally abnormal insulin (insulin Chicago). Diabetes 32:872-875.

256. Shoelson, S., Haneda, M., Blix, P., Nanjo, A., Sanke, T., Inouye, K., Steiner, D., Rubenstein, A., and Tager, H. 1983. Three mutant insulins in man. Nature 302:540-543.

257. Shoelson, S., Fickova, M., Haneda, M., Nahum, A., Musso, G., Kaiser, E.T., Rubenstein, A., and Tager, H. 1983. Indentification of a mutant human insulin predicetd to contain a serine-for -phenylalanine substitition. Proc. Natl. Acad. Sci. U. S. A. 80:7390-7394.

258. Kobayashi, M., Ohgaku, S., Iwasaki, M., Maegawa, H., Shigeta, Y., and Inouye, K. 1982. Characterization of [Leu B24 ]- and [Leu B25 ]-insulin analogues. Receptor binding and biological activity. Biochem. J. 206:597-603.

259. Kobayashi, M., Ohgaku, S., Iwasaki, M., Maegawa, H., Shigeta, Y., and Inouye, K. 1982. Supernormal insulin: [D-Phe B24 ]-insulin with increased affinity for insulin receptors. Biochem. Biophys. Res. Commun. 107:329-336.

260. Nanjo, K., Sanke, T., Miyano, M., Okai, K., Sowa, R., Kondo, M., Nishimura, S., Iwo, K., Miyamura, K., Given, B.D., et al. 1986. Diabetes due to secretion of a structurally abnormal insulin (insulin Wakayama). Clinical and functional characteristics of [Leu A3 ] insulin. J. Clin. Invest. 77:514-519.

261. Hua, Q.X., Shoelson, S.E., Inouye, K., and Weiss, M.A. 1993. Paradoxical structure and function in a mutant human insulin associated with diabetes mellitus. Proc. Natl. Acad. Sci. U. S. A. 90:582-586.

262. Liu, M., Haataja, L., Wright, J., Wickramasinghe, D.N., Hua, Q.X., Phillips, N.B., Barbetti, F., Weiss, M.A., and Arvan, P. 2010. Mutant INS-gene induced diabetes of youth: proinsulin cysteine residues impose dominant-negative inhibition on nonmutant proinsulin transport. PLos-One 5:e13333.

263. Huang, K., Xu, B., Hu, S.Q., Chu, Y.C., Hua, Q.X., Qu, Y., Li, B., Wang, S., Wang, R.Y., Nakagawa, S.H., et al. 2004. How insulin binds: the B-chain α-helix contacts the L1 β-helix of the insulin receptor. J. Mol. Biol. 341:529-550.

264. Schäffer, L. 1994. A model for insulin binding to the insulin receptor. Eur. J. Biochem. 221:1127-1132.

265. De Meyts, P. 1994. The structural basis of insulin and insulin-like growth factor-I receptor binding and negative co-operativity, and its relevance to mitogenic versus metabolic signaling. Diabetologia 37:S135-S148.

266. De Meyts, P., Palsgaard, J., Sajid, W., Theede, A.M., and Aladdin, H. 2004. Structural biology of insulin and IGF-1 receptors. Novartis Found. Symp. 262:160-171.

267. De Meyts, P., and Shymko, R.M. 2000. Timing-dependent modulation of insulin mitogenic versus metabolic signalling. Novartis Found. Symp. 227:46-57.

268. Hansen, B.F., Kurtzhals, P., Jensen, A.B., Dejgaard, A., and Russell-Jones, D. 2011. Insulin X10 revisited: a super-mitogenic insulin analogue. Diabetologia 54:2226-2231.

269. Hemkens, L.G., Bender, R., Grouven, U., and Sawicki, P.T. 2009. Insulin glargine and cancer. Lancet 374:1743-1744.

270. Weinstein, D., Simon, M., Yehezkel, E., Laron, Z., and Werner, H. 2009. Insulin analogues display IGF-I mitogenic and anti-apoptotic activities in cultured cancer cells. Diabetes Metab. Res. Rev. 25:41-49.

271. Sciacca, L., Cassarino, M.F., Genua, M., Pandini, G., LeMoli, R., Squatrito, S., and Vigneri, R. 2010. Insulin analogues differently activate insulin receptor isoforms and post-receptor signaling. Diabetologia 53:1743-1753.

272. Sciacca, L., Le Moli, R., and Vigneri, R. 2012. Insulin analogs and cancer. Front. Endocrinol. (Lausanne) 3:21.

273. Cosmatos, A., Cheng, K., Okada, Y., and Katsoyannis, P.G. 1978. The chemical synthesis and biological evaluation of [1-L-alanine-A]-and [1-D-alanine-A]insulins. J. Biol. Chem. 253:6586-6590.

274. Geiger, R., Geisen, K., Summ, H.D., and Langer, D. 1975. (A1-D-alanine) insulin. Hoppe Seylers Z Physiol. Chem. 356:1635-1649.

275. Geiger, R., Geisen, K., Regitz, G., Summ, H.D., and Langner, D. 1980. Insulin analogues with substitution of A1-glycine by D-amino acids and omega-amino acids (author's transl). Hoppe Seylers Z Physiol. Chem. 361:563-570.

276. Geiger, R., Geisen, K., and Summ, H.D. 1982. Astausch von A1-Glycin in Rinderinsulin gegen L-und D-Tryptophan. Hoppe Seylers Z Physiol. Chem. 363:S1231-1239.

277. Nakagawa, S.H., and Tager, H.S. 1989. Perturbation of insulin-receptor interactions by intramolecular hormone cross-linking. Analysis of relative movement among residues A1, B1, and B29. J. Biol. Chem. 264:272-279.

278. Kitagawa, K., Ogawa, H., Burke, G.T., Chanley, J.D., and Katsoyannis, P.G. 1984. Critical role of the A2 amino acid residue in the biological activity of insulin: [2-glycine-A]- and [2-alanine-A]insulins. Biochemistry 23:1405-1413.

279. Nakagawa, S.H., and Tager, H.S. 1992. Importance of aliphatic side-chain structure at positions 2 and 3 of the insulin A chain in insulin-receptor interactions. Biochemistry 31:3204-3214.

280. Kitagawa, K., Ogawa, H., Burke, G.T., Chanley, J.D., and Katsoyannis, P.G. 1984. Interaction between the A2 and A19 amino acid residues is of critical importance for high biological activity in insulin: [19-leucine-A]insulin. Biochemistry 23:4444-4448.

281. Wieneke, H.J., Danho, W., Bullesbach, E.E., Gattner, H.G., and Zahn, H. 1983. The synthesis of [A 19-3-iodotyrosine] and [A 19-3,5-diiodotyrosine]insulin (porcine). Hoppe Seylers Z Physiol. Chem. 364:537-550.

282. Ohta, N., Burke, G.T., and Katsoyannis, P.G. 1988. Synthesis of an insulin analogue embodying a strongly fluorescent moiety, [19-tryptophan-A]insulin. J. Protein Chem. 7:55-65.

283. Du, X., and Tang, J.G. 1998. Hydroxyl group of insulin A19Tyr is essential for receptor binding: studies on (A19Phe)insulin. Biochem. Mol. Biol. Int. 45:255-260.

284. Sonne, O., Linde, S., Larsen, T.R., and Gliemann, J. 1983. Monoiodoinsulin labelled in tyrosine residue 16 or 26 of the B-chain or 19 of the A-chain. II. Characterization of the kinetic binding constants and determination of the biological potency. Hoppe Seylers Z Physiol. Chem. 364:101-110.

285. Hamlin, J.L., and Arquilla, E.R. 1974. Monoiodoinsulin. Preparation, purification, and characterization of a biologically active derivative substituted predominantly on tyrosine A14. J. Biol. Chem. 249:21-32.

286. Linde, S., Welinder, B.S., Hansen, B., and Sonne, O. 1986. Preparative reversed-phase high-performance liquid chromatography of iodinated insulin retaining full biological activity. J. Chromatogr. 369:327-339.

287. Chu, Y.C., Zong, L., Burke, G.T., and Katsoyannis, P.G. 1992. The A14 position of insulin tolerates considerable structural alterations with modest effects on the biological behavior of the hormone. J. Protein Chem. 11:571-577.

288. Slobin, L.I., and Carpenter, F.H. 1963. Action of carboxypeptidase- A on bovine insulin: preparation of desalanine-desasparagine-insulin. Biochemistry 2:16-22.

289. Chu, Y.C., Wang, R.Y., Burke, G.T., Chanley, J.D., and Katsoyannis, P.G. 1987. Possible involvement of the A20-A21 peptide bond in the biological activity of insulin. 1. [21-Desasparagine,20-cysteinamide-A]insulin and [21-desasparagine,20-cysteine isopropylamide-A]insulin. Biochemistry 26:6966-6971.

290. Chu, Y.C., Wang, R.Y., Burke, G.T., Chanley, J.D., and Katsoyannis, P.G. 1987. Possible involvement of the A20-A21 peptide bond in the expression of the biological activity of insulin. 3. [21- Des asparagine,20-cysteine ethylamide-A]insulin and [21- des asparagine,20-cysteine 2,2,2-trifluoroethylamide-A]insulin. Biochemistry 26:6975-6979.

291. Chu, Y.C., Burke, G.T., Chanley, J.D., and Katsoyannis, P.G. 1987. Possible involvement of the A20-A21 peptide bond in the expression of the biological activity of insulin. 2. [21-Asparagine diethylamide-A]insulin. Biochemistry 26:6972-6975.

292. Thomas, B., and Wollmer, A. 1989. Cobalt probing of structural alternatives for insulin in solution. Biol. Chem. Hoppe Seyler 370:1235-1244.

293. Kaarsholm, N.C., Ko, H.C., and Dunn, M.F. 1989. Comparison of solution structural flexibility and zinc binding domains for insulin, proinsulin, and miniproinsulin. Biochemistry 28:4427-4435.

294. Wollmer, A., Rannefeld, B., Stahl, J., and Melberg, S.G. 1989. Structural transition in the metal-free hexamer of protein-engineered [B13 Gln]insulin. Biol. Chem. Hoppe Seyler 370:1045-1053.

295. Kruger, P., Gilge, G., Cabuk, Y., and Wollmer, A. 1990. Cooperativity and intermediate states in the T----R-structural transformation of insulin. Biol. Chem. Hoppe Seyler 371:669-673.

296. Roy, M., Lee, R.W., Brange, J., and Dunn, M.F. 1990. 1 H NMR spectrum of the native human insulin monomer. Evidence for conformational differences between the monomer and aggregated forms. J. Biol. Chem. 265:5448-5452.

297. Mirmira, R.G., and Tager, H.S. 1989. Role of the phenylalanine B24 side chain in directing insulin interaction with its receptor: Importance of main chain conformation. J. Biol. Chem. 264:6349-6354.

298. Mirmira, R.G., Nakagawa, S.H., and Tager, H.S. 1991. Importance of the character and configuration of residues B24, B25, and B26 in insulin-receptor interactions. J. Biol. Chem. 266:1428-1436.

299. Mirmira, R.G., and Tager, H.S. 1991. Disposition of the phenylalanine B25 side chain during insulin-receptor and insulin-insulin interactions. Biochemistry 30:8222-8229.

300. Cao, Q.P., Geiger, R., Langner, D., and Geisen, K. 1986. Biological activity in vivo of insulin analogues modified in the N-terminal region of the B-chain. Biol. Chem. Hoppe Seyler 367:135-140.

301. Schwartz, G., and Katsoyannis, P.G. 1978. Synthesis of des(tetrapeptide B(1-4)) and des(pentapeptide B(1-5) human insulins. Two biologically active analogues. Biochemistry 17:4550-4556.

302. Geiger, R. 1977. In Molecular Endocrinology . I. MacIntyre, and M. Szelke, editors. Amsterdam: Holland Biomedical Press. 27-41.

303. Nakagawa, S.H., and Tager, H.S. 1991. Implications of invariant residue LeuB6 in insulin-receptor interactions. J. Biol. Chem. 266:11502-11509.

304. Hua, Q.X., Liu, M., Hu, S.Q., Jia, W., Arvan, P., and Weiss, M.A. 2006. A conserved histidine in insulin is required for the foldability of human proinsulin. Structure and function of an Ala B5 analog. J. Biol. Chem. 281:24889-24899.

305. Cosmatos, A., Ferderigos, N., and Katsoyannis, P.G. 1979. Chemical synthesis of [ des (tetrapeptide B27-30), Tyr(NH2)26-B] and [ des (pentapeptide B26-30), Phe(NH2)25-B] bovine insulins. Int. J. Protein Res. 14:457-471.

306. Fischer, W.H., Saunders, D., Brandenburg, D., Wollmer, A., and Zahn, H. 1985. A shortened insulin with full in vitro potency. Biol. Chem. Hoppe Seyler 366:521-525.

307. Zakova, L., Brynda, J., Au-Alvarez, O., Dodson, E.J., Dodson, G.G., Whittingham, J.L., and Brzozowski, A.M. 2004. Toward the insulin-IGF-I intermediate structures: functional and structural properties of the [TyrB25NMePheB26] insulin mutant. Biochemistry 43:16293-16300.

308. Žáková, L., Kazdová, L., Hančlová, I., Protivínská, E., Šanda, M., Buděšínský, M., and Jiráček, J. 2008. Insulin analogues with modifications at position B26. Divergence of binding affinity and biological activity. Biochemistry 47:5858-5868.

309. Dai, J.B., Lou, M.Z., You, J.M., and Liang, D.C. 1987. Refinement of the structure of despentapetide (B26-B30) insulin at 1.5 Å resolution. Sci. Sin. 30:55-65.

310. Casaretto, M., Spoden, M., Diaconescu, C., Gattner, H.G., Zahn, H., Brandenburg, D., and Wollmer, A. 1987. Shortened insulin with enhanced in vitro potency. Biol. Chem. Hoppe Seyler 368:709-716.

311. Nakagawa, S.H., and Tager, H.S. 1993. Importance of main-chain flexibility and the insulin fold in insulin-receptor interactions. Biochemistry 32:7237-7243.

312. Boelens, R., Ganadu, M.L., Verheyden, P., and Kaptein, R. 1990. Two-dimensional NMR studies on des-pentapeptide-insulin. Proton resonance assignments and secondary structure analysis. Eur. J. Biochem. 191:147-153.

313. Nakagawa, S.H., and Tager, H.S. 1986. Role of the phenylalanine B25 side chain in directing insulin interaction with its receptor. Steric and conformational effects. J. Biol. Chem. 261:7332-7341.

314. Xu, B., Hu, S.Q., Chu, Y.C., Huang, K., Nakagawa, S.H., Whittaker, J., Katsoyannis, P.G., and Weiss, M.A. 2004. Diabetes-associated mutations in insulin: consecutive residues in the B chain contact distinct domains of the insulin receptor. Biochemistry 43:8356-8372.

315. Shoelson, S.E., Lu, Z.X., Parlautan, L., Lynch, C.S., and Weiss, M.A. 1992. Mutations at the dimer, hexamer, and receptor-binding, surfaces of insulin independently affect insulin-insulin and insulin-receptor interactions. Biochemistry 31:1757-1767.

316. Pittman, I., Nakagawa, S.H., Tager, H.S., and Steiner, D.F. 1997. Maintenance of the B-chain β-turn in [Gly B24 ] insulin mutants: a steady-state fluorescence anisotropy study. Biochemistry 36:3430-3437.

317. De Meyts, P. 2004. Insulin and its receptor: structure, function and evolution. Bioessays 26:1351-1362.

318. Kiselyov, V.V., Versteyhe, S., Gaugin, L., and De Meyts, P. 2009. Harmonic oscillator model of the insulin and IGF1 receptors' allosteric binding and activation. Mol. Syst. Biol. 5:243.

319. Whittaker, J., Whittaker, L.J., Roberts, C.T., Jr., Phillips, N.B., Ismail-Beigi, F., Lawrence, M.C., and Weiss, M.A. 2012. α-Helical element at the hormone-binding surface of the insulin receptor functions as a signaling element to activate its tyrosine kinase. Proc. Natl. Acad. Sci. U. S. A. 109:11166-11171.

320. Ward, C.W., and Lawrence, M.C. 2009. Ligand-induced activation of the insulin receptor: a multi-step process involving structural changes in both the ligand and the receptor. Bioessays 31:422-434.

321. Henzen, C. 2012. Monogenic diabetes mellitus due to defects in insulin secretion. Swiss Med. Wkly. 142:w13690.

322. Chan, S.J., Seino, S., Gruppuso, P.A., Schwartz, R., and Steiner, D.F. 1987. A mutation in the B chain coding region is associated with impaired proinsulin conversion in a family with hyperproinsulinemia. Proc. Natl. Acad. Sci. U. S. A. 84:2194-2197.

323. Gruppuso, P.A., Gorden, P., Kahn, C.R., Cornblath, M., Zeller, W.P., and Schwartz, R. 1984. Familial hyperproinsulinemia due to a proposed defect in conversion of proinsulin to insulin. N. Engl. J. Med. 311:629-634.

324. Schwartz, G.P., Burke, G.T., and Katsoyannis, P.G. 1987. A superactive insulin: [B10-aspartic acid]insulin(human). Proc. Natl. Acad. Sci. U. S. A. 84:6408-6411.

325. Burgess, T.L., and Kelly, R.B. 1987. Constitutive and regulated secretion of proteins. Annu. Rev. Cell Biol. 3:243-293.

326. Carroll, R.J., Hammer, R.E., Chan, S.J., Swift, H.H., Rubenstein, A.H., and Steiner, D.F. 1988. A mutant human proinsulin is secreted from islets of Langerhans in increased amounts via an unregulated pathway. Proc. Natl. Acad. Sci. U. S. A. 85:8943-8947.

327. Robbins, D.C., Shoelson, S., Rubenstein, A., and Tager, H. 1984. Familial hyperproinsulinemia. Two cohorts secreting indistinguishable type II intermediates of proinsulin conversion. J. Clin. Invest. 73:714-719.

328. Gabbay, K.H., DeLuca, K., Fisher, J.N., Jr., Mako, M.E., and Rubenstein, A.H. 1976. Familial hyperproinsulinemia. An autosomal dominant defect. N. Engl. J. Med. 294:911-915.

329. Edghill, E.L., Flanagan, S.E., Patch, A.M., Boustred, C., Parrish, A., Shields, B., Shepherd, M.H., Hussain, K., Kapoor, R.R., Malecki, M., et al. 2008. Insulin mutation screening in 1044 patients with diabetes: mutations in the INS gene are a common cause of neonatal diabetes but a rare cause of diabetes diagnosed in childhood or adulthood. Diabetes 57:1034-1042.

330. Colombo, C., Porzio, O., Liu, M., Massa, O., Vasta, M., Salardi, S., Beccaria, L., Monciotti, C., Toni, S., Pedersen, O., et al. 2008. Seven mutations in the human insulin gene linked to permanent neonatal/infancy-onset diabetes mellitus. J. Clin. Invest. 118:2148-2156.

331. Molven, A., Ringdal, M., Nordbo, A.M., Raeder, H., Stoy, J., Lipkind, G.M., Steiner, D.F., Philipson, L.H., Bergmann, I., Aarskog, D., et al. 2008. Mutations in the insulin gene can cause MODY and autoantibody-negative type 1 diabetes. Diabetes 57:1131-1135.

332. Slingerland, A.S., and Hattersley, A.T. 2005. Mutations in the Kir6.2 subunit of the KATP channel and permanent neonatal diabetes: new insights and new treatment. Ann. Med. 37:186-195.

333. Babenko, A.P., Polak, M., Cave, H., Busiah, K., Czernichow, P., Scharfmann, R., Bryan, J., Aguilar-Bryan, L., Vaxillaire, M., and Froguel, P. 2006. Activating mutations in the ABCC8 gene in neonatal diabetes mellitus. N. Engl. J. Med. 355:456-466.

334. Park, S.Y., Ye, H., Steiner, D.F., and Bell, G.I. 2010. Mutant proinsulin proteins associated with neonatal diabetes are retained in the endoplasmic reticulum and not efficiently secreted. Biochem. Biophys. Res. Commun. 391:1449-1454.

335. Meur, G., Simon, A., Harun, N., Virally, M., Dechaume, A., Bonnefond, A., Fetita, S., Tarasov, A.I., Guillausseau, P.J., Boesgaard, T.W., et al. 2010. Insulin gene mutations resulting in early-onset diabetes: marked differences in clinical presentation, metabolic status, and pathogenic effect through endoplasmic reticulum retention. Diabetes 59:653-661.

336. Yoshioka, M., Kayo, T., Ikeda, T., and Koizumi, A. 1997. A novel locus, Mody4 , distal to D7Mit189 on chromosome 7 determines early-onset NIDDM in nonobese C57BL/6 (Akita) mutant mice. Diabetes 46:887-894.

337. Wang, J., Takeuchi, T., Tanaka, S., Kubo, S.K., Kayo, T., Lu, D., Takata, K., Koizumi, A., and Izumi, T. 1999. A mutation in the insulin 2 gene induces diabetes with severe pancreatic b -cell dysfunction in the Mody mouse. J. Clin. Invest. 103:27-37.

338. Oyadomari, S., Koizumi, A., Takeda, K., Gotoh, T., Akira, S., Araki, E., and Mori, M. 2002. Targeted disruption of the Chop gene delays endoplasmic reticulum stress-mediated diabetes. J. Clin. Invest. 109:525-532.

339. Izumi, T., Yokota-Hashimoto, H., Zhao, S., Wang, J., Halban, P.A., and Takeuchi, T. 2003. Dominant negative pathogenesis by mutant proinsulin in the Akita diabetic mouse. Diabetes 52:409-416.

340. Zuber, C., Fan, J.Y., Guhl, B., and Roth, J. 2004. Misfolded proinsulin accumulates in expanded pre-Golgi intermediates and endoplasmic reticulum subdomains in pancreatic β cells of Akita mice. FASEB J. 18:917-919.

341. Liu, M., Hodish, I., Rhodes, C.J., and Arvan, P. 2007. Proinsulin maturation, misfolding, and proteotoxicity. Proc. Natl. Acad. Sci. U. S. A. 104:15841-15846.

342. Yoshinaga, T., Nakatome, K., Nozaki, J., Naitoh, M., Hoseki, J., Kubota, H., Nagata, K., and Koizumi, A. 2005. Proinsulin lacking the A7-B7 disulfide bond, Ins2Akita, tends to aggregate due to the exposed hydrophobic surface. Biol. Chem. 386:1077-1085.

343. Hua, Q.X., Nakagawa, S.H., Jia, W., Hu, S.Q., Chu, Y.C., Katsoyannis, P.G., and Weiss, M.A. 2001. Hierarchical protein folding: asymmetric unfolding of an insulin analogue lacking the A7-B7 interchain disulfide bridge. Biochemistry 40:12299-12311.

344. Jia, X.Y., Guo, Z.Y., Wang, Y., Xu, Y., Duan, S.S., and Feng, Y.M. 2003. Peptide models of four possible insulin folding intermediates with two disulfides. Protein Sci. 12:2412-2419.

345. Herbach, N., Rahtkolb, B., Kemter, E., Pichl, L., Klaften, M., De Angelis, M.H., Halban, P., Wolf, E., Aigner, B., and Wanker, R. 2007. Dominant-negative effects of a novel mutated Ins2 allele causes early-onset diabetes and severe β-cell loss in Munich Ins2 C95S mutant mice. Diabetes 56:1268-1276.

346. Hodish, I., Liu, M., Rajpal, G., Larkin, D., Holz, R.W., Adams, A., Liu, L., and Arvan, P. 2010. Misfolded proinsulin affects bystander proinsulin in neonatal diabetes. J. Biol. Chem. 285:685-694.

347. Hodish, I., Absood, A., Liu, L., Liu, M., Haataja, L., Larkin, D., Al-Khafaji, A., Zaki, A., and Arvan, P. 2011. In vivo misfolding of proinsulin below the threshold of frank diabetes. Diabetes 60:2092-2101.

348. Haataja, L., Snapp, E., Wright, J., Liu, M., Hardy, A.B., Wheeler, M.B., Markwardt, M.L., Rizzo, M., and Arvan, P. 2013. Proinsulin intermolecular interactions during secretory trafficking in pancreatic β cells. J. Biol. Chem. 288:1896-1906.

349. Absood, A., Gandomani, B., Zaki, A., Nasta, V., Michail, A., Habib, P.M., and Hodish, I. 2013. Insulin therapy for pre-hyperglycemic β-cell endoplasmic reticulum crowding. PLoS One 8:e54351.

350. Shoelson, S.E., Polonsky, K.S., Zeidler, A., Rubenstein, A.H., and Tager, H.S. 1984. Human insullin B24 (Phe ® Ser), secretion and metabolic clearance of the abnormal insulin in man and in a dog model. J. Clin. Invest. 73:1351-1358.

351. Fonseca, S.G., Gromada, J., and Urano, F. 2011. Endoplasmic reticulum stress and pancreatic β-cell death. Trends Endocrinol. Metab. 22:266-274.

352. Thomas, S.E., Dalton, L., Malzer, E., and Marciniak, S.J. 2011. Unravelling the story of protein misfolding in diabetes mellitus. World J. Diabetes 2:114-118.

353. Ron, D. 2002. Proteotoxicity in the endoplasmic reticulum: lessons from the Akita diabetic mouse. J. Clin. Invest. 109:443-445.

354. Yuan, Q., Tang, W., Zhang, X., Hinson, J.A., Liu, C., Osei, K., and Wang, J. 2012. Proinsulin atypical maturation and disposal induces extensive defects in mouse Ins2 +/Akita β-cells. PLoS One 7:e35098.

355. Despa, F. 2010. Endoplasmic reticulum overcrowding as a mechanism of β-cell dysfunction in diabetes. Biophys. J. 98:1641-1648.

Hypogonadotropic Hypogonadism (HH) and Gonadotropin Therapy

TAKE HOME POINTS

1-The differential diagnosis of HH includes structural, functional and genetic abnormalities affecting the hypothalamic-pituitary-gonadal axis.

2-While our understanding of the genetic basis of HH has expanded considerably in the past decade, mutations have only been identified in approximately 40% of patients to date.

3-The traditional view that congenital HH is a simple monogenic disorder has been challenged by the demonstration of oligogenicity, which helps explain the phenotypic variability underlying this condition.

4-HH is a treatable cause of male infertility with larger testicular size and higher baseline inhibin B levels predicting a favorable response to gonadotropin or GnRH therapy, while negative prognostic indicators include cryptorchidism and possibly prior androgen use.

5-While lifelong androgen replacement is necessary in most men with HH, a brief discontinuation of hormonal therapy should be considered in all cases given that reversibility has been demonstrated in up to 10% of cases.

 

DEFINITIONS

Pulsatile secretion of gonadotropin-releasing hormone (GnRH) from the hypothalamus is required for both the initiation and maintenance of the reproductive axis in the human. Pulsatile GnRH stimulates biosynthesis of luteinizing hormone (LH) and follicle-stimulating hormone (FSH) that in turn initiate both intra-gonadal testosterone production and spermatogenesis as well as systemic testosterone secretion and virilization. Failure of this episodic GnRH secretion or action, or disruption of gonadotropin secretion, results in the clinical syndrome of hypogonadotropic hypogonadism (HH). Disorders causing HH are differentiated from primary testicular disease by the demonstration of low/normal gonadotropin levels in the setting of low testosterone concentrations and sperm counts (Fig. 1). Congenital abnormalities leading to HH are rare but well described and are usually the consequence of deficient GnRH secretion occurring either in isolation (normosmic congenital hypogonadotropic hypogonadism (nCHH)), or in association with anosmia (Kallmann syndrome (KS)). A growing number of loci have been associated with congenital GnRH deficiency and mutations in the GnRH receptor, as well as both LH-b and FSH-b subunits, have been reported. Acquired causes of HH are more common and can be due to any disorder that affects the hypothalamic-pituitary axis.

Fig. 1. Schematic of the hypothalamic-pituitary-gonadal axis in a normal adult male and in the setting of primary and secondary hypogonadism.

NORMAL GnRH SECRETION ACROSS DEVELOPMENT IN THE MALE

In the human, the pattern of GnRH-induced gonadotropin secretion is constantly changing during sexual development. Therefore, the establishment of a robust normative database is critical for understanding pathologic states like HH.

Fetal/Neonatal Life

GnRH neurons originate outside the central nervous system in the olfactory placode, migrate along the olfactory, terminalis, and vomeronasal nerves up the nasal septum, and through the cribriform plate to the forebrain, ultimately reaching their final destination in the arcuate nucleus of the hypothalamus (1). WhileGnRH neurons have been demonstrated in the fetal hypothalamus by 9 weeks gestation, it is not until 16 weeks that functional connections are established between these neurons and the portal system. From mid-gestation until 6 months of postnatal life, pulsatile secretion of GnRH stimulates gonadotropin biosynthesis and secretion that, in turn, initiates gonadal sex steroid production (2).

Childhood Period

During childhood, the hypothalamic-pituitary axis is not completely quiescent and is characterized by low amplitude GnRH secretion as mirrored by LH secretion using ultrasensitive LH assays (3).

Puberty                                              

The onset of puberty is marked by sleep-entrained reactivation of the reproductive axis characterized by a striking increase in the amplitude of LH pulses with a lesser change in frequency (4).  This nocturnal rise of LH secretion stimulates gonadal secretion of both sex steroids and inhibin B, which return to prepubertal levels during the daytime. As puberty progresses, secretion of gonadotropins occurs during both day and night, allowing sexual development to be completed. The precise neuroendocrine trigger to puberty is still unknown. However, it is likely to be a process that removes inhibition of GnRH release rather than one increasing GnRH synthesis as abundant GnRH mRNA is present in the hypothalamic neurons of primates at an equivalent developmental stage (5, 6).

Adulthood

During adulthood, gonadotropins are secreted in a pulsatile fashion. In the adult male, LH is secreted in pulses approximately every 2 hours  (Fig. 2A) (7).  However, considerable variability is observed in LH pulse patterns and there is a wide range of testosterone secretory patterns. Indeed, in 15% of normal men whose hypothalamic-pituitary-gonadal (HPG) axis was examined using frequent blood sampling, serum testosterone levels as low as 3.5 nmol/L were recorded (to convert to ng/dL, multiply by 28.6) following long inter-pulse intervals of LH secretion, although mean testosterone levels remained within the normal range (Fig. 3). This within-patient variation must be considered when interpreting single LH and testosterone measurements obtained during the evaluation of a male with suspected hypogonadism. This variability is particularly important in middle-aged and older men as up to 30% men that are found to have a low testosterone concentration, will have a normal level on repeat testing (8).

 

Fig. 2. Spectrum of GnRH-induced LH secretion in men with GnRH deficiency. LH pulsations are indicated by asterisks. A, Normal adult male pattern of GnRH secretion with high amplitude regular LH pulsations and normal serum testosterone (T), testicular volume (TV) and sperm count; B, Disordered amplitude pattern of GnRH secretion in an CHH male, characterized by low amplitude LH pulsations, low serum T, and azoospermia; C, Sleep-entrained or developmental arrest pattern of LH secretion in an CHH male characterized by relatively low amplitude LH pulsations clustered during the night-time hours analogous to the pattern which normally occurs at puberty. Note that the TV is higher than in the subject with the apulsatile pattern; D, Apulsatile pattern of GnRH secretion in an CHH male with complete absence of endogenous LH pulsations, low serum T, prepubertal TV and azoospermia.

Fig. 3. Frequent blood sampling (q 10 min for 24 h) of serum testosterone (T) and LH in a normal male. Note that a T level of 3.2 nmol/L (to convert to ng/dL, multiply by 28.6) which is in the hypogonadal range was recorded after a long interpulse interval of LH secretion.

CLINICAL PRESENTATION

The clinical features of GnRH and/or gonadotropin deficiency are typically first manifested at puberty, a time when there is normally a marked increase in GnRH secretion. The phenotypic presentation of HH varies with age of onset (congenital vs. acquired), severity (complete vs. partial), and duration (functional vs. permanent).

HH may also be diagnosed in the neonatal period. The typical clinical phenotype is that of a male infant with normal sexual differentiation, cryptorchidism and micropenis in whom gonadotropin and sex steroid levels are inappropriately low given the normal activation of the HPG axis during this period. Early in fetal life, the testosterone production required for full sexual differentiation is thought to be stimulated by maternal hCG alone. However, endogenous secretion of GnRH in the late fetal/early neonatal periods appears necessary for inguino-scrotal descent of the testes and full growth of the external genitalia (9, 10). Accordingly, cryptorchidism and microphallus have been reported in up to 50% of patients with IHH and KS in some small series (11, 12) and may represent surrogate markers of failure of activation of GnRH secretion during the neonatal window.

Most often, the diagnosis of HH is delayed until adolescence, when there is a failure to go through puberty. The clinical presentation includes lack of development of secondary sex characteristics, eunuchoidal body proportions (upper/lower body ratio <1 with an arm span 6 cm > standing height), a high-pitched voice, mild anemia, delayed bone age, and pre-pubertal testes (12). However, the syndrome is clinically heterogeneous in that some patients have evidence of partial spontaneous pubertal development reflected by larger gonadal size despite hypogonadal testosterone levels and inappropriately low gonadotropin levels. Gynecomastia is not a typical feature of GnRH deficiency given the hypogonadotropic state and is most commonly seen in patients treated with gonadotropins (13, 14). HH may also present after completion of puberty resulting in a disruption in reproductive function in adulthood characterized by decreased libido, erectile dysfunction and oligo- or azoospermia (15).

DIFFERENTIAL DIAGNOSIS

Hypogonadotropic hypogonadism may be broadly classified into congenital or acquired disorders (Table 1).

 

Table 1. Differential Diagnosis of Hypogonadotropic Hypogonadism (HH)

Congenital HH

  • Normosmic hypogonadotropic ypogonadism
  • Kallmann syndrome
  • Adult onset idiopathic HH
  • Fertile eunuch syndrome
  • Adrenal Hypoplasia Congenita

Genetic defects of the gonadotropin subunits
HH associated with other pituitary hormone deficiencies
HH associated with obesity

  • Prader-Willi syndrome
  • Laurence- Moon-Biedl syndrome
Acquired HH
Structural

  • Tumors
    • Craniopharyngiomas
    • Pituitary adenomas (e.g. prolactinoma, non functioning tumor)
    • Germinoma, glioma, meningioma
  • Infiltrative disorders
    • Sarcoidosis, hemochromatosis, histiocytosis X
  • Head trauma
  • Radiation therapy
  • Pituitary apoplexy

Functional

  • Exercise
  • Dieting
  • Anabolic steroids
  • Glucocorticoid therapy
  • Narcotics
  • Critical illness

Congenital Hypogonadotropic Hypogonadism

CHH is characterized by an isolated defect in GnRH secretion as evidenced by: i) complete or partial absence of GnRH-induced LH pulsations (Fig. 1) (16, 17); ii) normalization of pituitary-gonadal axis function in response to physiological regimens of exogenous GnRH replacement (16-18); iii) otherwise normal hormonal testing of the anterior pituitary including a normal ferritin level and; iv) normal imaging of the hypothalamic-pituitary region.

CHH was first reported in association with anosmia and termed Kallmann syndrome (19). However, it was subsequently appreciated that several patients with CHH lack evidence of an olfactory defect and thus have a normosmic form of CHH.  While the majority of CHH patients present with lack of pubertal development, there is considerable clinical heterogeneity. Depending on the degree of prior spontaneous pubertal development, testicular size in men with CHH may range from prepubertal to near-normal adult testes. In addition to anosmia, a variety of other anomalies have been reported to occur in CHH with an increased frequency in the KS subset including cleft lip and palate, synkinesia, sensorineural deafness, cerebellar ataxia, renal agenesis, digital bony abnormalities and dental agenesis (19-21). Some of these clinical features are highly associated with particular genetic causes of KS so that the clinical phenotype can be utilized to prioritize genetic testing (22).

Given that it is a rare disease, data on the incidence of CHH is limited with estimates varying from 1/10,000 to 1/86,000 (23, 24).  Isolated GnRH deficiency occurs more commonly in men than in women. Based on our review of 250 consecutive cases seen at the Massachusetts GeneralHospital, the male:female ratio is 4:1.

Variant or Partial Forms of GnRH Deficiency

i) Adult onset IHH

An acquired form of isolated GnRH deficiency termed adult-onset idiopathic HH was first reported in 1997 (15).  In this group of patients, puberty occurs normally and is followed years later by a decrease in libido, sexual function and fertility.  The biochemical profile of these patients is indistinguishable from subjects with congenital GnRH deficiency in that they have an apulsatile pattern of LH secretion associated with low serum testosterone levels. In addition, more than 90% of cases have normal restoration of the pituitary-gonadal axis when treated with physiologic GnRH replacement regimens supporting a hypothalamic defect as the origin of the disorder. Unlike patients with functional GnRH deficiency, no factors known to impair GnRH secretion transiently such as stress, exercise or weight loss have been identified in this population. In addition, longitudinal follow-up of these cases of adult-onset IHH for over a decade suggests that the neuroendocrine defect is permanent in that all remained frankly hypogonadal (25).

ii) Fertile Eunuch Syndrome

In 1950, McCullagh et al provided the first description of a patient with the fertile eunuch syndrome characterized by eunuchoidal proportions and lack of secondary sexual characteristics in the presence of normal size testes and preserved spermatogenesis (26). In this disorder, enfeebled endogenous GnRH secretion appears sufficient to achieve the intra-gonadal testosterone levels needed to support spermatogenesis and testicular growth, but is insufficient to induce virilization.  The clinical picture of the fertile eunuch is rather similar to that of mid-pubertal boys; indeed, frequent blood sampling in two men with the fertile eunuch syndrome demonstrated a nocturnal rise of LH and testosterone secretion synchronous with sleep, analogous to the pattern seen in mid-puberty (27). In contrast, we described a patient with the fertile eunuch syndrome displaying a detectable but apulsatile pattern of LH secretion, who was found to harbor a partially inactivating mutation of the GnRH receptor (GnRH-R) (28). The clinical presentation of the fertile eunuch syndrome also reveals some similarity to adult onset IHH in that both are characterized by GnRH deficiency in association with normal or near-normal testicular size.  However, “fertile eunuchs” are distinguished by the preservation of spermatogenesis and the achievement of fertility with testosterone or hCG therapy alone (29, 30). 

iii) Delayed puberty

Frequently, there is a history of delayed, but otherwise normal, puberty among the families of patients with CHH (21).  While the incidence of delayed puberty in the general population is less than 1% (31), rates of up to 12% have been observed in families with CHH (21). These data suggest that delay in initiating, but subsequent normal progression through, puberty may represent the mildest end of the phenotypic spectrum of CHH.

Acquired  Hypogonadotropic Hypogonadism
Functional

Functional forms of HH are characterized by a transient defect in GnRH secretion. This type of presentation occurs most commonly in female hypogonadotropic subjects with hypothalamic amenorrhea (HA). In susceptible individuals, HA may be precipitated by factors such as significant weight loss, exercise or stress (32-34). In addition, a genetic predisposition to HA has also been identified with the demonstration of rare variants in genes associated with CHH in some women with this disorder (35). Typically, GnRH secretion will resume after correcting the underlying abnormality and menstruation will be restored.  While in women the presence or absence of menses acts as a useful clinical marker of the functioning of the HPG axis, there is no comparable clinical marker in the male. Moderate to severe dietary restriction in otherwise healthy men has been shown to decrease testosterone levels by impairing secretion of GnRH (36, 37). In addition, some (38, 39) but not all (40) studies have shown that strenuous physical exercise may adversely affect testosterone concentrations. However to date, a clinical syndrome of functional GnRH deficiency in men that is analogous to HA has not been definitively established.

Drug-Induced GnRH Deficiency

Use of anabolic steroids may result in a functional form of HH manifested by decreased concentrations of both testosterone and dihydrotestosterone and a marked impairment of spermatogenesis (41, 42).  While the suppression of the HPG axis induced by anabolic steroids is reversible, rate of recovery following cessation of steroid use is variable and can range from 4-12 months (41, 43).  Chronic treatment with glucocorticoids may also lead to hypogonadism. In one study, 16 men with chronic pulmonary disease who received high-dose glucocorticoids for at least one month had a mean serum testosterone of 6.9 nmol/L, compared to 15.6 nmol/L in 11 men matched for age and disease (44). Given that serum LH levels did not increase in this study, these data suggest a predominantly central mechanism for glucocorticoid-induced hypogonadism. Chronic use of oral and intrathecal narcotic analgesics may also suppress LH secretion and result in reversible HH (45). A common side-effect of psychotropic medications such as phenothiazines or risperidone is hyperprolactinemia, which inhibits endogenous GnRH release resulting in HH (46).

Critical Illness

Any period of severe chronic (47), or acute illness such as surgery (48), myocardial infarction (49), burn injury (50), and renal disease (51) may result in low testosterone levels (52). Acute injury is accompanied by a prompt and direct suppression of Leydig cell function (53). When severe stress becomes prolonged, hypogonadotropism ensues largely due to attenuation of pulsatile LH release (53). Endogenous dopamine or opiates may be involved in the pathogenesis of HH induced by critical illness (52).

Structural

Structural lesions of the hypothalamus and pituitary can interfere with the normal pattern of GnRH and/or gonadotropin secretion. The majority of patients with HH secondary to such tumors have multiple pituitary hormone deficiencies in addition to that of gonadotropins (54, 55). However, a mass lesion in the pituitary or hypothalamus is more likely to disrupt the secretion of gonadotropins than that of ACTH or TSH. Thus, patients may present with hypogonadism in the absence of adrenal or thyroid hormone deficiency.

In children, craniopharyngioma is the most common tumor resulting in HH, and is often associated with growth retardation, visual field defects and diabetes insipidus. In adults, prolactinomas are the most frequent cause of HH and may do so by either interfering with GnRH secretion, or in the case of macroadenomas, by local destruction and compression of the gonadotropes.  Hyperprolactinemia results in altered dopaminergic function, which has been shown to reduce GnRH mRNA levels and decrease serum levels of LH, FSH and testosterone (56, 57). Although men with hyperprolactinemia may develop galactorrhea, it occurs much less frequently than in women, due to lack of prior stimulation by estrogen and progesterone.

Rarer causes of HH include infiltrative disorders of the hypothalamus or pituitary such as hemochromatosis, sarcoidosis, lymphocytic hypophysitis and histiocytosis, in which case the presence of systemic signs and symptoms frequently leads one to the diagnosis (58-59). Cranial irradiation for the treatment of CNS tumors or leukemia may also result in the gradual onset of hypothalamic-pituitary failure (60). The degree of impairment depends upon both the dose and type of radiation employed and typically reflects hypothalamic dysfunction since the hypothalamus is significantly more radiosensitive than the pituitary. In general, the younger the patient the greater the susceptibility to endocrine dysfunction following radiation therapy (60). Sudden and severe hemorrhage into the pituitary can also results in permanent impairment of pituitary function including hypogonadism (61).

PATHOPHYSIOLOGY

The combined use of both frequent blood sampling and genetic studies has contributed to our understanding of the pathophysiology of isolated GnRH deficiency. Due to its confinement within the hypophyseal-portal blood supply, direct sampling of GnRH is not feasible in the human. In addition, measurements of GnRH in the peripheral circulation do not accurately reflect its secretion due to its rapid half-life of 2-4 min (62, 63).  Consequently, inferential approaches must be used to study GnRH secretion in the human.

Traditionally, LH has been used as a surrogate marker of GnRH activity (64, 65).  More recently, the pulsatile component of free alpha subunit (FAS) secretion has been shown to be tightly correlated with that of LH (66)and to be driven by GnRH based on its eradication by GnRH receptor blockade (67). Given its half-life of 12-15 min, FAS is useful in tracking GnRH secretion at fast pulse frequencies (68).

Patterns of GnRH Secretion in Subjects with GnRH Deficiency

A range of abnormalities in the neuroendocrine pattern of GnRH secretion mirrors the clinical spectrum of CHH.  We examined pulsatile gonadotropin secretion in 50 men with isolated GnRH deficiency during 10-min blood sampling for up to 24 h (16, 17). The largest subset of patients (84%) had no detectable LH pulses (apulsatile pattern, Fig. 2D) and were found to have neither historical nor physical evidence of puberty, thus representing the most severe form of GnRH deficiency. A second pattern of GnRH secretion was observed in which LH pulses occurred at a normal frequency but were of diminished amplitude compared to those of normal men (decreased amplitude pattern, Fig. 2B).  This decreased pulse amplitude pattern is suggestive of either enfeebled hypothalamic release of GnRH or a state of GnRH resistance as may be seen with a partial defect at the level of the GnRH-R. A third group of subjects demonstrated predominantly nocturnal LH pulsations (developmental arrest pattern, Fig. 2C); these patients had some testicular growth and a history consistent with an arrest of puberty. Finally, pulsatile LH activity appeared normal by RIA in one patient; however, LH bioactivity was absent when tested in the dispersed rat Leydig cell assay (69). Thus in CHH men, the spectrum of abnormalities in GnRH secretion or action is likely to contribute to the clinical and biochemical heterogeneity of this disorder.

Genetics of CHH

Considerable genetic heterogeneity has also been found to underlie CHH. Most cases of CHH are sporadic (80%), suggesting that either the frequency of spontaneous mutations in this disorder is high or that the etiology of many cases is not genetic. In the last decade considerable advances have been made in unraveling the genetic basis for CHH and to date, mutations have been identified in approximately 40% of cases (70). Unique genetic mechanisms have been described for both KS and nCHH. However, some probands with KS have family members with hypogonadism but normal olfaction (Fig. 4) and the same mutations have been shown to be associated with both KS and nCHH suggesting that they should no longer be viewed as distinct diagnostic subsets (71).

Fig. 4. A family with GnRH deficiency and multiple affected members in two generations. The pedigree is compatible with autosomal dominant transmission with incomplete penetrance. Note the presence of GnRH deficiency both with and without anosmia in the same family.

Human GnRH deficiency has traditionally been viewed as a monogenic disorder that could be inherited as an autosomal dominant, autosomal recessive, or X-linked trait. However, the long appreciated phenotypic variability within and across families with single gene defects combined with the recent demonstration that patients with CHH may harbor a mutation in more than one gene (72, 73) suggest a more complex mode of inheritance. Indeed, a systematic study of nearly 400 patients with isolated GnRH deficiency who were screened for 8 known loci revealed 17% of patients harbor multiple mutations consistent with an oligogenic rather than a monogenic mode of inheritance (74).

Kallmann Syndrome Gene (KAL1)

Mutations in the KAL1 gene were the first characterized genetic defect reported to cause GnRH deficiency (75). Anosmin, the protein encoded by the KAL gene, shares homology with molecules involved in neural development and contains 4 contiguous fibronectin type III repeats found in neural cell adhesion molecules (76). Confirmation of the causative role of the KAL gene in KS was provided by a study of a KS fetus with a deletion from Xp22.31 to Xpter, i.e. including the entire gene. Histologic examination of the brain of this fetus demonstrated that migration of the GnRH and olfactory neurons was arrested just below the telencephalon at the cribriform plate (77). Therefore, mutations in KAL1 appear to cause premature termination of migration of both the olfactory and GnRH neurons to the brain resulting in anosmia and IHH. This profound defect of GnRH neuronal migration results in the complete failure of activation of the HPG axis in patients with X-linked KS (X-KS). Indeed, X-KS patients typically present with more severe forms of CHH characterized by a high incidence of microphallus, cryptorchidism, complete absence of puberty with small testes, and very low inhibin B levels (78, 79).  Further, this subset of men with CHH has the poorest response to therapy to induce spermatogenesis (80).

Mutations in KAL1 are distributed throughout the gene, although most point mutations cluster in the four fibronectin type III repeat domains (81-82). Most X-KS mutations cause alteration of splicing, frameshift or stop codons and result in synthesis of a truncated anosmin protein (83). Missense mutations have rarely been described (83, 84). There appears to be some variability in the phenotypic expression of mutations both within and between families (84, 85). No correlation has been demonstrated between phenotype and location of the mutation described (86).

Autosomal Genes

Despite the biology revealed by the X-linked cases, several lines of evidence suggest that autosomal genes account for the majority of familial cases. Of 36 familial cases with GnRH deficiency studied by Waldstreicher and colleagues, only 21% could be attributed to X-linkage (21).  When the analysis was extended to include surrogate markers of IHH (isolated congenital anosmia and delayed puberty), the X-linked pedigrees comprised only 11%, autosomal recessive 25% and autosomal dominant 64%. These data suggest that in familial cases, the X-linked form of GnRH deficiency is the least common.

GnRH1 and GNRHR Mutations      

Although an obvious candidate gene, mutations of GNRH1, the gene encoding the preprohormone that is processed to GnRH, remained elusive for many years and it was not until 2009 that mutations in this gene were identified by two independent groups (87, 88).  In the first report a homozygous GNRH1 frameshift mutation was identified in a brother and sister that completely deleted the GnRH decapeptide sequence resulting in severe GnRH deficiency manifested by complete absence of puberty (87). Exogenous GnRH administration was shown to restore normal LH pulsatility indicating normal pituitary responsiveness.

While mutations in GNRH1 are a rare cause of CHH, loss-of-function mutations in the GnRH receptor are much more common with 22 novel mutations having been reported to date (89). In this autosomal recessive form of GnRH deficiency, affected patients are homozygous for one mutation or compound heterozygous (with two mutations, each on one allele), and in rare sporadic nIHH cases, heterozygous mutations have been identified (74). Patients harboring GNRHR mutations present with a wide phenotypic spectrum ranging from the fertile eunuch syndrome (28) to partial CHH (90, 91) to the most complete form of GnRH deficiency characterized by cryptorchidism, microphallus, undetectable gonadotropins, and absence of pubertal development (92-93). Patients with hypogonadism caused by inactivating mutations of the GnRH receptor tend to have a poor response to pulsatile GnRH administration (94). However, successful ovulation and conception have been described in a patient with an inactivating receptor mutation receiving high doses of pulsatile GnRH, thus indicating that many GnRH receptor mutations create a relative rather than an absolute resistance to pulsatile GnRH administration (95).

Kisspeptin 1 receptor (KISS1R: formerly GPR54) and its ligand KISS1

Linkage analysis of a large consanguineous family with HH identified this locus for autosomal recessive CHH on chromosome 19p13 (96).  Shortly thereafter Seminara et al confirmed KISS1R as an important regulator of puberty in a different consanguineous family (97). Murine studies demonstrated a hypogonadal phenotype similar to that seen in the human that was responsive to exogenous GnRH (97). While mutations in KISS1R appear to be an uncommon cause of GnRH deficiency in the human accounting for <5% of cases, these discoveries opened a new area of exploration in relation to the genetics of puberty and GnRH biology upstream of the hypothalamus.  Recently, heterozygous variants in KISS1 have been noted in association with CHH (98).

Fibroblast Growth Factors

The study of two patients affected by different contiguous gene syndromes, both of which included KS, led to the discovery of fibroblast growth factor receptor-1 (FGFR1) as a gene associated with congenital GnRH deficiency (99).  Interestingly, mutations in FGFR1 can cause both KS and nCHH (100, 101) and have been identified in as many as 10% of CHH cases.   Patients harboring FGFR1 mutations exhibit a wide range of both reproductive and non-reproductive phenotypes including cleft palate, agenesis of the corpus callosum, dental agenesis, unilateral hearing loss, and skeletal anomalies (100-103).  The variable expressivity of phenotypes among patients and family members harboring identical FGFR1 mutations is consistent with a more complex, oligogenic model for GnRH deficiency (74, 103).

In 2008, a critical ligand for FGFR-1, FGF8, was found to be involved in GnRH neuron ontogeny (104).  Subjects harboring FGF8 mutations also display a broad spectrum of pubertal development ranging from absent, to partial, to complete puberty (in a male with adult onset HH).  Additionally, the non-reproductive associated phenotypes are wide ranging and include hearing loss and a variety of skeletal anomalies (high arched palate, cleft lip/palate, severe osteoporosis, camplodactyly, and hyperlaxity of the digits) (105).  Further, the variable expressivity observed in FGFR1 families is also evident in family members harboring the same FGF8 mutation (103, 105).

Recently, the key developmental role of FGF signaling in reproduction has been highlighted by the identification of mutations in genes encoding a broader range of modulators of the FGFR1 pathway including FGF17 and other members of the so-called FGF8 synexpression group (106).  Mutations in these genes are associated with complex modes of inheritance and they are thought to act primarily as contributors to an oligogenic genetic architecture underlying CHH.

Prokineticins (PmROKR2 and PROK2).

Similar to what has been seen with FGFR1, mutations in the prokineticin system (PROKR2 and PROK2) have been associated with both KS and nCHH (107-110).  It is unusual for homozygous mutations to occur and mutations in both ligand and receptor are typically heterozygous.  In a review of a large cohort of KS patients harboring mutations in PROK2/PROKR2, Sarfati and colleagues noted that those harboring biallelic mutations exhibit more severe and less variable reproductive phenotypes than those with monoallelic mutations (111).

Neurokinins (TAC3 and TACR3)

The strategy of studying consanguineous families with CHH, which led to the discovery of KISS1R, was also successful in identifying mutations in other autosomal genes (TAC3 and TACR3) with a hitherto unrecognized role in reproduction (112).  Patients harboring mutations in TAC3 and TACR3 have a particularly high incidence of microphallus suggesting that this pathway may be important in regulating the “mini-puberty” of infancy (113). A second striking phenotype associated with mutations in this pathway is that recovery of the hypothalamic-pituitary-gonadal axis has been observed in a significant number of patients (113).  Further, careful characterization of patients harboring TAC3/TAC3R mutations reveals these patients exhibit low frequency GnRH pulses and a relatively elevated FSH/LH ratio, a biochemical clue for selective gene screening (114).

Genetic Defects of the Gonadotropins

Genetic defects in the genes encoding LH-b subunit and FSH-b are rare and have been summarized in a recent review (115).  The initial case report of a male who failed to go through puberty identified a homozygous mutation in amino acid 54 of the LH-b subunit that rendered LH incapable of binding to its receptor (116).  The mutant hormone had no biologic activity but had normal immunoreactivity.  Testosterone levels were reduced in association with elevated LH and FSH levels. This clinical picture could easily be confused with primary hypogonadism had the patient not been shown to have a normal serum testosterone response to hCG administration.  The first reported male with an FSH-b gene mutation (Cys82Arg missense) appeared normally virilized and had normal testosterone levels, but small testes, azoospermia, a high LH and a low FSH level (117).  The second reported case of a man with an FSH-b gene mutation (Val61X) had very small testes (1-2 mL) and an unexpected absence of pubertal development; he was found to harbor an additional defect in Leydig cell function as evidenced by high LH and low testosterone levels (118).

Other rare genetic syndromes associated with hypogonadotropic hypogonadism

HH also occurs as one of a constellation of features in a number of rare syndromes including X-linked adrenal hypoplasia congenita (AHC) caused by defects in the DAX1 (NROB1) gene (Xp21) (119, 120), CHARGE syndrome (coloboma, heart defects, atresia of the choanae, retardation of growth and development, genital anomaly, ear abnormalities) resulting from CHD7 (8q12.1) defects (121, 122), leptin deficiency (123, 124), and prohormone convertase 1/3 deficiency (5q15-21) (125).

In addition, combined pituitary hormone deficiency has been linked with rare abnormalities in genes encoding transcription factors necessary for pituitary development.  Mutations in the gene, PROP1 (Prophet of Pit-1) appear to be the most common cause of both familial and sporadic congenital combined pituitary hormone deficiency (126).  Most patients with inactivating mutations in PROP1 exhibit low gonadotropin levels and fail to enter puberty as well as having additional deficiencies in GH, PRL, and TSH (126).  However, there is a report of two sibships with a homozygous mutation, Arg120Cys in PROP1 in which the affected children entered puberty spontaneously and then developed gonadotropin deficiency over time (127). The same sequence of events was observed in several children with mutations causing a complete loss of function raising questions as to whether this should more correctly be viewed as a model of acquired rather than congenital gonadotropin deficiency (128).

Recently the critical role of PROK2, PROKR2 and FGF8 in the developing pituitary has been noted in the genetic overlap between KS and developmental disorders including combined pituitary hormone deficiency, septo-optic dysplasia, and holoprosencephaly (129, 130). Mutations in the gene SOX2 have been identified in patients with HH in association with anopthalmia or microphhalmia (131, 132), while SOX10 and IL17RD are key genes to consider when deafness is present in association with CHH (106, 133).

EVALUATION

Hypogonadism is defined as a defect in one of the two major functions of the testes i.e. production of testosterone and spermatogenesis. The presence of hypogonadism can reflect disorders intrinsic to the testes (primary or hypergonadotropic hypogonadism) or disorders of the pituitary or hypothalamus (secondary or hypogonadotropic hypogonadism). These two entities can be distinguished by measuring serum LH and FSH concentrations (Fig. 5).  Primary hypogonadism is characterized by a low serum testosterone level and oligo- or azoospermia in the presence of elevated serum LH and FSH concentrations. In contrast, secondary hypogonadism is diagnosed in the setting of a low testosterone level and sperm count in association with low or inappropriately normal serum LH and FSH concentrations.

Figure 5.Algorithm for evaluation of hypogonadism in a male.

It is best to measure testosterone in a morning sample given the normal diurnal rhythm of testosterone secretion. Repeat measurements should be performed if the initial reading is low. In secondary hypogonadism, measuring the serum LH response to a single bolus of exogenous GnRH is not helpful in distinguishing pituitary from hypothalamic disease, because a subnormal response may occur in both settings.  Patients with HH due to congenital absence of hypothalamic GnRH secretion may have had no prior exposure to GnRH; in this setting, repeated administration of GnRH is needed to prime the gonadotropes to elicit a gonadotropin response. Characterization of the pulsatile pattern of LH secretion with frequent blood sampling is useful in refining the diagnosis in a research setting, but is not practical for clinical purposes.

A semen analysis should be obtained to assess sperm production and the count, motility, and morphology determined. Based on WHO criteria, the parameters for a normal semen analysis are a sperm density ³ 15 million sperm/mL of ejaculate with total motility ³ 40% and strictly defined normal morphology of ³ 4% (134). Two normal semen analyses are sufficient to indicate that the sperm density and motility are normal.

In the case of secondary hypogonadism, it is critical to assess the rest of the pituitary axis, including a prolactin level to ensure that the defect is isolated to the HPG axis.  Transferrin saturation should be measured to exclude hemochromatosis. Patients with HH typically undergo adrenarche at a normal age and should therefore have normal adult male levels of DHEAS.  Radiographic evaluation should include a bone age determination, MRI of the pituitary and hypothalamic area and, in adults, a DEXA scan to assess bone mineral density.  Renal ultrasound should be considered to assess for unilateral renal agenesis in patients with KS especially those known to have a KAL1 mutation.

The growing genetic heterogeneity of CHH poses a significant challenge to clinicians in the diagnostic work of patients. While in familial cases the mode of inheritance can be used to guide genetic testing, the majority of CHH cases are actually sporadic. However, a careful clinical evaluation can be helpful in prioritizing genetic testing.  In a recent analysis of 219 patients with KS the following clinical features were highly associated with specific gene defects: synkinesia (KAL1), dental agenesis (FGF8/FGFR1), digital bony abnormalities (FGF8/FGFR1), and hearing loss (CHD7) (135).  Weaker associations were observed between renal agenesis and KAL1, cleft lip/palate and FGF8/FGFR1, CHD7 and HS6ST1, and male reversal and FGF8/FGFR1 and HS6ST1 (135).

TREATMENT

Choosing the most appropriate treatment for men with HH should be based upon an informed discussion between the patient and his physician.  Androgen therapy, whether by exogenous testosterone replacement or induction of endogenous testosterone production by hCG is needed in all HH patients.  Androgens play a number of important physiologic roles in the human and are required not only for virilization and normal sexual function, but also for maintenance of both muscle and bone mass, as well as normal mood and cognition.

While testosterone is the primary treatment modality used to induce and maintain secondary sexual characteristics and sexual function in men with hypogonadism, treatment with testosterone does not restore fertility.  Therefore, in patients in whom fertility is the treatment goal, induction of gonadotropin secretion by pulsatile GnRH or exogenous gonadotropins is necessary. While hormone therapy is the mainstay of treatment for CHH, cryptorchidism, if present, should be treated surgically with orchidopexy, ideally before the age of 2 years to improve fertility outcomes and reduce the risk of future testicular malignancy (136).
The traditional teaching has been that patients with CHH require life long therapy but the recent demonstration of reversal of hypogonadism suggests that this many not always be the case (113, 137, 138).  In a study of a large cohort of men with CHH, 10% of cases demonstrated a sustained reversal after discontinuation of treatment. Reversal occurred across a spectrum of phenotypes including patients with and without anosmia and those with absent or partial spontaneous puberty (137). A key, but sometimes subtle, clinical finding that can predict reversal is spontaneous testicular enlargement on testosterone therapy (137, 138).  Reversal has been documented in patients with a variety of genetic defects including GNRHR, FGFR1 and CHD (137, 138) but has been shown to be particularly common in those with defects in the neurokinin pathway (113).  While the mechanism of reversal in CHH is not understood it is possible is that exposure to sex steroids may enhance the plasticity of the neuronal network responsible for GnRH secretion.  In light of these studies, it seems reasonable to recommend that a brief discontinuation of hormonal therapy be considered in patients with CHH to assess for reversibility. The trial period off treatment will depend on the formulation used but should be approximately 2-3 months for intramuscular testosterone and 4 weeks for those receiving transdermal preparations.

Testosterone Substitution

Testosterone is currently available in a variety of formulations for clinical use including injectable esters, transdermal patches, and gel preparations, each with its own unique pharmacokinetic profile.  This topic is discussed in detail in Chapter 2.

Induction of Spermatogenesis

In patients who desire fertility, the options to induce spermatogenesis include exogenous gonadotropins, which are freely available or pulsatile GnRH, use of which is restricted to specialist centers. Conventional therapy uses hCG as an LH substitute in conjunction with FSH in the form of either human menopausal gonadotropins, highly purified urinary FSH preparations or recombinant FSH formulations.  Both FSH and LH seem to be required.

The role of FSH in the initiation and maintenance of spermatogenesis in the human is controversial.  There is discordance between the phenotype of men with mutations in the FSH receptor gene (FSHR) who are oligospermic (139) and those with mutations of the FSH-b subunit who are azoospermic (117, 118).  If one assumes that the FSHR mutations were able to abolish FSH action completely, this report suggests that LH and testosterone alone are sufficient to induce spermatogenesis given that all five reported cases had sperm in the ejaculate and two had fathered children (139).  However, the demonstration that the two patients with a mutation in the FSH-ß subunit were azoospermic (117, 118) suggests that the degree of inactivation of the FSH-R may have been incomplete.  However, interpretation of this report is confounded by the fact that one of these patients also had hypogonadal testosterone levels so that the azoospermia in this case could not be attributed to lack of FSH alone (118).  The importance of FSH in initiating spermatogenesis is further called in to question by the demonstration that in a subset of patients with HH and larger testicular size, spermatogenesis can be stimulated with hCG alone (140).  However, this group largely comprises men with the fertile eunuch syndrome (29, 30), who may well have sufficient endogenous FSH secretion to sustain normal spermatogenesis with hCG alone.

Several other lines of evidence support a role for FSH in the maintenance of spermatogenesis. In contraceptive trials using testosterone esters, azoospermia is achieved only when serum FSH levels are rendered undetectable (141, 142).  Similarly, there is a case report of a patient with a history of a hypophysectomy for a pituitary adenoma who was found to be fertile when treated with testosterone alone due a concomitant activating mutation of the FSH-R (143). A small Finnish study of boys with very small testes (< 3mL) resulting from severe GnRH deficiency or panhypopituitarism following treatment of intracranial tumors showed that treatment with recombinant human FSH (r-hFSH) alone prior to induction of puberty with hCG and FSH increased testicular size leading the authors to postulate that priming with FSH may have beneficial effects on spermatogenesis later in life (144).  We recently expanded these findings in a randomized, open label study of the effects of r-hFSH on both hormonal and histologic outcomes in a cohort of men with CHH and prepubertal testes (145).  Seven men were randomized to 4 months of r-hFSH followed by 24 months of GnRH and six men received 24 months of pulsatile GnRH only. Treatment with r-hFSH resulted in a doubling of testicular size, increased inhibin B levels into the normal range, and induced Sertoli cell and spermatagonia proliferation. All men receiving r-hFH pretreatment developed sperm in the ejaculate versus 4/6 of the GnRH only group and trended towards higher maximal sperm counts. However, given the small sample size r-hFSH pretreatment was not shown to be statistically superior to GnRH alone (145).

In addition to FSH, adequate intra-testicular testosterone concentrations are also essential for normal spermatogenesis as illustrated by a study of hypogonadal men in which spermatogenesis could be induced by the combination of hCG and hMG but not by a regimen comprising purified FSH and testosterone (146). In summary, the evidence suggests that qualitatively and quantitatively normal spermatogenesis is best maintained in the presence of both FSH and LH-induced T secretion.

1.    Exogenous Gonadotropin Therapy

The typical gonadotropin regimen for induction of spermatogenesis in men comprises hCG in combination with FSH. Purified hCG is an effective substitute for LH given the structural homology between these 2 hormones which act through the same receptor on Leydig cells.  While a variety of FSH formulations are now available in different countries, there is little to choose between them in terms of therapeutic efficacy. Traditionally, FSH has been administered in the form of human menopausal gonadotropins (hMG) derived from the urine of postmenopausal women.  Although hMG has both FSH and LH activity, FSH activity predominates and LH activity is so low that combined administration with hCG is necessary to achieve fertility. Subsequently, highly purified urinary FSH preparations were developed, which have enhanced specific activity in comparison to hMG (10,000 IU/mg of protein vs 150 IU/mg of protein for hMG).  In the early 1990s, r-hFSH formulations were developed, which have greater purity and specific activity than any of the urinary preparations and no intrinsic LH activity (147, 148).  r-hFSH is produced in genetically engineered Chinese hamster ovary cells, in which the genes encoding the alpha and beta subunits have been introduced using recombinant DNA technology (147).  Pharmacokinetic studies of r-hFSH indicate a half-life of 48 ± 5 h and a dose-dependent increase in the serum level of FSH (147).

In our experience, the subcutaneous route of administration is as effective as the intramuscular route for both gonadotropins and significantly increases patient compliance. Therapy is typically initiated with hCG alone at a dose of 1,000 IU on alternate days and the dose titrated based on trough testosterone levels and testicular growth (149). Alternatively, recombinant human hCG can also be administered subcutaneously from a prefilled syringe. In some patients the dose of hCG can be decreased over time as testicular size increases. In the majority of patients with larger testes at baseline, spermatogenesis can be initiated with hCG alone most likely due to residual FSH secretion (140, 150).  Once there is a plateau in the response to hCG which typically occurs at around 6 months, therapy with FSH (in one of the three forms described above) should be added at a dose of 75 IU on alternate days.  If sperm output and testicular growth remain suboptimal the dose of FSH can be gradually increased to 150 U daily. Continuation of this combined regimen for 12-24 months induces testicular growth in almost all patients, spermatogenesis in approximately 80% and pregnancy rates in the range of 50% (151-153).  In an Australian study of 75 men with HH treated with gonadotropins the median time for sperm to appear in the ejaculate was 7.1 months and for conception was just over 28 months (152).  Similar data were reported in a compilation of clinical trial data from Asian, European, Australian and American patients (153). Factors predictive of better outcome include larger baseline testicular size and absence of cryptorchidism.  The Australian study reported that prior androgen use is also a negative prognostic indicator of response (152).  While the study investigators propose that gonadotropin therapy be considered to induce puberty based on their results, confirmation that such an approach is superior to the conventional practice of giving testosterone would require a large clinical trial the logistics of which would be challenging given the rarity of this patient population.  Gynecomastia is the most common side effect of gonadotropin therapy and is due to hCG stimulation of aromatase causing increased secretion of estradiol.  This undesirable side effect can be prevented by using the lowest dose of hCG capable of maintaining serum testosterone levels towards the lower end of the normal range.

In the majority of HH patients treated with gonadotropins sperm density remains below the normal range. However, failure to achieve a normal sperm density does not preclude fertility. Indeed, the median sperm concentrations reported at conception range from 5-8 million/mL (140, 152).  While spermatogenesis can be initiated even in patients with very small testes (18, 140), a longer duration of therapy is typically required and it may take up to 24 months for spermatogenesis to be induced.  Accordingly when discussing the issue of fertility with patients, we recommend that they start treatment at least 6 to 12 months prior to the time at which fertility is desired.  Once pregnancy is achieved, we advise continuing therapy until at least the second trimester.  If the couple plans to have another child in the near future, then hCG monotherapy should be continued. However, if a long interval is expected to elapse before the next pregnancy, it may be more convenient for the patient to resume testosterone therapy. Patients should also be offered the option of storing sperm for subsequent use in intrauterine insemination or intracytoplasmic sperm injection.  In patients in whom the combination of hCG and FSH is required to induce spermatogenesis initially, treatment with hCG alone may be sufficient for subsequent pregnancies due to larger testicular size.

In patients with panhypopituitarism who fail to respond to gonadotropin therapy, the addition of recombinant growth hormone (rGH) therapy should be considered. It is thought that a direct effect of growth hormone on Leydig cells may play a role in the delayed puberty encountered commonly in patients with isolated growth hormone deficiency (154, 155).  While small non-randomized studies suggest that rGH may enhance the testosterone response to hCG administration (154), larger, randomized studies are needed before a definitive decision about the benefit of rGH in inducing spermatogenesis in men with hypopituitarism can be reached.

2. Pulsatile GnRH Therapy

The alternative to gonadotropin therapy is pulsatile administration of GnRH, which may be administered by a programmable, portable mini-infusion pump (Fig. 6) (156).  While intravenous administration produces the most physiologic GnRH pulse contour and ensuing LH response (157), the subcutaneous route is clearly more practical for the long term treatment required to stimulate spermatogenesis. Based on our normative data (7), the frequency of GnRH administration that we employ is every 2 hours.  The dose of GnRH is titrated for each individual to ensure normalization of testosterone, LH and FSH and varies from 25 to 600 ng/kg per bolus. Patients on longterm therapy are monitored with serum testosterone and gonadotropin levels at monthly intervals. Once testicular volume reaches 6-8 mL, regular semen analyses are obtained. The majority of patients require treatment for 18-24 months to maximize testicular growth and achieve spermatogenesis, although the time taken to reach these endpoints tends to be shorter in those with a larger initial gonadal size. In our analysis of 76 men with CHH, pulsatile GnRH restored normal testosterone levels in 93% of cases and was successful in inducing spermatogenesis in 77% by 12 months and 82% by 24 months (158). As with gonadotropin therapy, testicular size is an important predictor of successful treatment while negative predictors include a history of cryptorchidism and a pre-treatment inhibin B level <60 pg/mL (158).

Fig. 6. LH and testosterone (T) at baseline and in response to GnRH therapy in a man with complete GnRH deficiency. Baseline LH secretion pattern is apulsatile with a hypogonadal T level. Exogenous pulsatile GnRH therapy administered every 2 h stimulates normal LH pulses which in turn induce a normal serum T level.

If pulsatile GnRH treatment fails, a mutation of GNRHR should be considered (94) and genetic testing arranged, if available.  A second cause of failure of pulsatile GnRH treatment is the appearance of anti-GnRH antibodies, which typically occur in the setting of erratic compliance and are associated with a progressive decrease in T and gonadotropins levels.

Both exogenous gonadotropins and pulsatile GnRH are very effective in stimulating spermatogenesis.  Most studies have no shown no advantage of either therapy in terms of testicular  growth, onset of  spermatogenesis, final  sperm counts or pregnancy rates (151, 159).

However, pulsatile GnRH therapy is not approved for induction of spermatogenesis by the Food and Drug Administration in the United States, and its use is thus confined to specialist centers.

Assisted Reproductive Technology

With the advances that have been made in the area of assisted reproductive technology (ART), couples who would previously have been offered donor insemination or adoption are now achieving pregnancies despite persistent severe impairments in sperm number and/or quality after hormonal therapy.  While male factor infertility was initially considered a contraindication to in vitro fertilization (IVF), IVF is now considered as an acceptable treatment option for the infertile male.  However, for men with severe male factor infertility, IVF is often unsuccessful and more sophisticated techniques to assist fertilization such as intracytoplasmic sperm injection (ICSI) are required.  The success rates of ICSI are high with fertilization rates of 50-60% and pregnancy rates of ~30% per cycle. If no sperm are present in the ejaculate, sperm retrieved either from epididymal aspiration or from testicular aspiration/biopsy can be used successfully for ICSI.  (See Chapter 7)

The timeframe for referring men who are undergoing therapy to stimulate spermatogenesis for ART has changed in the last decade as both access to ICSI and success rates have improved. In the study by Liu et al of pregnancy outcomes in men treated with gonadotropins, 7/58 pregnancies (12%) resulted from IVF (152).  Early referral for ICSI should certainly be considered in couples where there is concern that reduced ovarian reserve due to the age of the female partner may pose an additional barrier to natural conception.   However, for other couples lack of insurance coverage may make the procedure prohibitively expensive.  In addition, couples may have concerns about the ethics or safety of ART given that follow up of children conceived by ICSI suggests that there is a small increase in the risk of major malformations and chromosomal abnormalities particularly hypospadias and imprinting disorders (160-164). While the evidence is still inconclusive due to the inherent bias associated with observational studies (165), the potential for harm to the offspring underscores the importance of genetic counseling of couples contemplating assisted reproduction.

 

 

REFERENCES

 

1.         Sullivan KA, Silverman A-J. 1993 The ontogeny of gonadotropin-releasing hormone neurons in the chick. Neuroendocrinology. 58:597-608.

2.         Waldhauser F, Weissenbacher G, Frisch H, Pollak A. 1981 Pulsatile secretion of gonadotropins in early infancy. European Journal of Pediatrics. 137:71-74.

3.         Apter D, Cacciatore B, Alfthan H, Stenman UH. 1989 Serum luteinizing hormone concentrations increase 100-fold in females from 7 years to adulthood, as measured by time-resolved immunofluorometric assay. J Clin Endocrinol Metab. 68:53-57.

4.         Boyar RM, Rosenfeld RS, Kapen S, et al. 1974 Human puberty. Simultaneous augmented secretion of luteinizing hormone and testosterone during sleep. J Clin Invest. 54:609-18.

5.         Wiemann JN, Clifton DK, Steiner RA. 1989 Pubertal changes in gonadotropin-releasing hormone and proopiomelanocortin gene expression in the brain of the male rat. Endocrinology. 124:1760-1767.

6.         El Majdoubi M, Sahu A, Ramaswamy S, Plant TM. 2000 Neuropeptide Y: A hypothalamic brake restraining the onset of puberty in primates. Proc Natl Acad Sci USA. 97:6179-6184.

7.         Spratt DI, O'Dea LSL, Schoenfeld D, Butler J, Rao PN, Crowley Jr WF. 1988 Neuroendocrine-gonadal axis in men: frequent sampling of LH, FSH, and testosterone. American Journal of Physiology. 254:E658-E666.

8.         Brambilla DJ, O’Donnell AB, Matsumoto AM, McKinlay JB. 2007 Intraindividual variation in levels of serum testosterone and other reproductive and adrenal hormones in men. Clin Endocrinol (Oxf) 67:853-862

 

9.         Hutson JM. 1985 A biphasic model for the hormonal control of testicular descent. Lancet. 2:419-21.

10.       Berkowitz GS, Lapinski RH, Dolgin SE, Gazella JG, Bodian CA, Holzman IR. 1993 Prevalence and natural history of cryptorchidism. Pediatrics. 92:44-9.

11.       Bardin CW, Ross GT, Rifkind AB, Cargille CM, Lipsett MB. 1969 Studies of the pituitary-Leydig cell axis in young men with hypogonadotropic hypogonadism and hyposmia: comparison with normal men, prepubertal boys, and hypopituitary patients. J Clin Invest. 48:2046-2056.

12.       Van Dop C, Burstein S, Conte FA, Grumbach MM. 1987 Isolated gonadotropin deficiency in boys: Clinical characteristics and growth. Journal of Pediatrics. 111:684-692.

13.       Boyar RM, Finkelstein JW, Witkin M, Kapen S, Weitzman E, Hellman L. 1973 Studies of endocrine function in "isolated" GnRH deficiency. Journal of Clinical Endocrinology and Metabolism. 36:64-72.

14.       Lieblich JM, Rogol AD, White BJ, Rosen SW. 1982 Syndrome of anosmia with hypogonadotropic hypogonadism (Kallmann syndrome). American Journal of Medicine. 73:506-519.

15.       Nachtigall LB, Boepple PA, Pralong FP, Crowley Jr WF. 1997 Adult-onset idiopathic hypogonadotropic hypogonadism-a treatable form of male infertility. New England Journal of Medicine. 336:410-415.

16.       Santoro N, Filicori M, Crowley Jr WF. 1986 Hypogonadotropic disorders in men and women: diagnosis and therapy with pulsatile gonadotropin-releasing hormone. Endocrine Reviews. 7:11-23.

17.       Spratt DI, Carr DB, Merriam GR, Scully RE, Rao PN, Crowley Jr WF. 1987 The spectrum of abnormal patterns of gonadotropin-releasing hormone secretion in men with idiopathic hypogonadotropic hypogonadism: clinical and laboratory correlations. Journal of Clinical Endocrinology and Metabolism. 64:283-291.

18.       Whitcomb RW, Crowley Jr WF. 1990 Clinical review 4: diagnosis and treatment of isolated gonadotropin-releasing hormone deficiency in men. Journal of Clinical Endocrinology and Metabolism. 70:3-7.

19.       Kallmann FJ, Schoenfeld WA. 1944 The genetic aspects of primary eunuchoidism. American Journal of Mental Deficiency. 158:203-236.

20.       Schwankhaus JD, Currie J, Jaffe MJ, Rose SR, Sherins RJ. 1989 Neurologic findings in men with isolated hypogonadotropic hypogonadism. Neurology. 39:223-226.

21.       Waldstreicher J, Seminara SB, Jameson JL, et al. 1996 The genetic and clinical heterogeneity of gonadotropin-releasing hormone deficiency in the human. Journal of Clinical Endocrinology and Metabolism. 81:4388-4395.

22.       Costa-Barbosa FA, Balasubramanian R, Keefe KW, et al. 2013 Prioritizing genetic testing in patients with Kallmann syndrome using clinical phenotypes. Journal of Clinical Endocrinology and Metabolism. 98:E943-953.

23.       Filippi G. 1986 Klinefelter's syndrome in Sardinia. Clinical report of 265 hypogonadic males detected at the time of military check-up. Clinical Genetics. 30:276-284.

24.       Fromantin M, Gineste J, Didier A, Rouvier J. 1973 Les impuberismes et les hypogonadismes a l'incorporation. Probl Actuels Endocrinol Nutr. 16:179-199.

25.       Dwyer AA, Hayes FJ, Plummer L, Pitteloud N, Crowley WF Jr. 2010 The long-term clinical follow-up and natural history of men with adult-onset idiopathic hypogonadotropic hypogonadism. Journal of Clinical Endocrinology and Metabolism. 95:4235-4243.

26.       McCullagh EP, Beck JC, Schaffenburg CA. 1953 A syndrome of eunuchoidism with spermatogenesis, normal urinary FSH and low or normal ICSH: "Fertile Eunuchs". Journal of Clinical Endocrinology and Metabolism. 13:489-509.

27.       Boyar RM, Wu RH, Kapen S, Hellman L, Weitzman ED, Finkelstein JW. 1976 Clinical and laboratory heterogeneity in idiopathic hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 43:1268-75.

28.       Pitteloud N, Boepple PA, DeCruz S,  Valkenburgh SB, Crowley WF Jr, Hayes FJ. 2001 The fertile eunuch variant of idiopathic hypogonadotropic hypogonadism: spontaneous reversal associated with a homozygous mutation in the GnRH receptor. J Clin Endocrinol Metab.86: 2470-75.

29.       Rogol AD, Mittal KK, White BJ, McGinniss MH, Lieblich JM, Rosen SW. 1980 HLA-compatible paternity in two "fertile eunuchs" with congenital hypogonadotropic hypogonadism and anosmia (the Kallmann syndrome). Journal of Clinical Endocrinology and Metabolism. 51:275-279.

30.       Smals AGH, Kloppenborg PWC, van Haelst UJG, Lequin R, Benraad TJ. 1978 Fertile eunuch syndrome versus classic hypogonadotrophic hypogonadism. Acta Endocrinologica. 87:389-399.

31.       Styne DM. 1991 Puberty and its disorders in boys. Endocrine and Metabolism Clinics of North America. 20:43-69.

32.       Berga SL, Girton LG. 1989 The psychoneuroendocrinology of functional hypothalamic amenorrhea. Psychiatric Clinics of North America. 12:105-116.

33.       Laughlin GA, Dominguez CE, Yen SSC. 1998 Nutritional and endocrine-metabolic aberrations in women with functional hypothalamic amenorrhea. Journal of Clinical Endocrinology and Metabolism. 83:25-32.

34.       Warren MP. 1980 The effects of exercise on pubertal progression and reproductive function in girls. Journal of Clinical Endocrinology and Metabolism. 51:1150-1157.

35.       Caronia LM, Martin C, Welt CK, et al. 2011 A genetic basis for functional hypothalamic amenorrhea. New Engl J Med. 364:215-225.

36.       Klibanski A, Beitins IZ, Badger T, Little R, McArthur JW. 1981 Reproductive function during fasting in men. Journal of Clinical Endocrinology and Metabolism. 53:258-263.

37.       Veldhuis JD, Iranmanesh A, Evans WS, Lizarralde G, Thorner MO, Vance ML. 1993 Amplitude suppression of the pulsatile mode of immunoradiometric luteinizing hormone release in fasting-induced hypoandrogenemia in normal men. Journal of Clinical Endocrinology and Metabolism. 76:587-593.

38.       McColl EM, Wheeler GD, Gomes P, Bhambhani Y, Cumming DC. 1989 The effects of acute exercise on pulsatile LH release in high-mileage male runners. Clin Endocrinol (Oxf). 31:617-21.

39.       Wheeler GD, Wall SR, Belcastro AN, Cumming DC. 1984 Reduced serum testosterone and prolactin levels in male distance runners. Journal of the American Medical Association. 252:514-6.

40.       Snegovskaya V, Viru A. 1993 Steroid and pituitary hormone responses to rowing: relative significance of exercise intensity and duration and performance level. Eur J Appl Physiol Occup Physiol. 67:59-65.

41.       Alen M, Reinila M, Vihko R. 1985 Response of serum hormones to androgen administration in power athletes. Med Sci Sports Exerc. 17:354-9.

42.       Ruokonen A, Alen M, Bolton N, Vihko R. 1985 Response of serum testosterone and its precursor steroids, SHBG and CBG to anabolic steroid and testosterone self-administration in man. J Steroid Biochem. 23:33-8.

43.       Turek PJ, Williams RH, Gilbaugh JH 3rd, Lipshultz LI. 1995 The reversibility of anabolic steroid-induced azoospermia. J Urol 153:1628-1630.

44.       MacAdams MR, White RH, Chipps BE. 1986 Reduction of serum testosterone levels during chronic glucocorticoid therapy. Ann Intern Med. 104:648-51.

 

45.       Abs R, Verhelst J, Maeyaert J, et al. 2000 Endocrine consequences of long-term intrathecal administration of opioids. J Clin Endocrinol Metab. 85:2215-22.

46.       Molitch ME. 2005 Medication-induced hyperprolactinemia. Mayo Clinic Proc. 80:1050-1057.

47.       Van den Berghe G, de Zegher F, Lauwers P, Veldhuis JD. 1994 Luteinizing hormone secretion and hypoandrogenaemia in critically ill men: effect of dopamine. Clin Endocrinol (Oxf). 41:563-9.

48.       Wang C, Chan V, Yeung RT. 1978 Effect of surgical stress on pituitary-testicular function. Clin Endocrinol (Oxf). 9:255-66.

49.       Wang C, Chan V, Tse TF, Yeung RT. 1978 Effect of acute myocardial infarction on pituitary-testicular function. Clin Endocrinol (Oxf). 9:249-53.

50.       Lephart ED, Baxter CR, Parker CR, Jr. 1987 Effect of burn trauma on adrenal and testicular steroid hormone production. J Clin Endocrinol Metab. 64:842-8.

51.       Iglesias P, Carrero JJ, Díez JJ. 2012 Gonadal dysfunction in men with chronic kidney disease: clinical features, prognostic implications and therapeutic options. J Nephrol. 25:31-42.

52.       Van den Berghe G, de Zegher F, Bouillon R. 1998 Clinical review 95: Acute and prolonged critical illness as different neuroendocrine paradigms. J Clin Endocrinol Metab. 83:1827-34.

53.       Spratt DI, Bigos ST, Beitins I, Cox P, Longcope C, Orav J. 1992 Both hyper- and hypogonadotropic hypogonadism occur transiently in acute illness: bio- and immunoactive gonadotropins. J Clin Endocrinol Metab. 75:1562-70.

54.       Bates AS, Van't Hoff W, Jones PJ, Clayton RN. 1996 The effect of hypopituitarism on life expectancy. J Clin Endocrinol Metab. 81:1169-72.

55.       Arafah BM. 1986 Reversible hypopituitarism in patients with large nonfunctioning pituitary adenomas. J Clin Endocrinol Metab. 62:1173-9.

56.       Selmanoff M, Shu C, Petersen SL, Barraclough CA, Zoeller RT. 1991 Single cell levels of hypothalamic messenger ribonucleic acid encoding luteinizing hormone-releasing hormone in intact, castrated, and hyperprolactinemic male rats. Endocrinology. 128:459-66.

57.       Carter JN, Tyson JE, Tolis G, Van Vliet S, Faiman C, Friesen HG. 1978 Prolactin-secreting tumors and hypogonadism in 22 men. N Engl J Med. 299:847-52.

58.       Charbonnel B, Chupin M, Le Grand A, Guillon J. 1981 Pituitary function in idiopathic haemochromatosis: hormonal study in 36 male patients. Acta Endocrinol (Copenh). 98:178-83.

59.       Thodou E, Asa SL, Kontogeorgos G, Kovacs K, Horvath E, Ezzat S. 1995 Clinical case seminar: lymphocytic hypophysitis: clinicopathological findings. J Clin Endocrinol Metab. 80:2302-11.

60.       Appelman-Dijkstra NM, Kokshoorn NE, Dekkers OM, Neelis KJ, Biermasz NR, Romijn JA, Smit JW, Pereira AM. 2011 Pituitary dysfunction in adult patients after cranial radiotherapy: systematic review and meta-analysis. J Clin Endocrinol Metab.  96: 2330-40.

61.       Randeva HS, Schoebel J, Byrne J, Esiri M, Adams CB, Wass JA. 1999 Classical pituitary apoplexy: clinical features, management and outcome. Clin Endocrinol (Oxf). 51:181-8.

62.       Arimura A, Kastin AJ, Gonzales-Barcena D, Siller J, Weaver RE, Schally AV. 1974 Disappearance of LH-releasing hormone in man as determined by radioimmunoassay. 3:421-425.

63.       Pimstone B, Epstein S, Hamilton SM, LeRoith D, Hendricks S. 1977 Metabolic clearance and plasma half-disappearance time of exogenous gonadotropin-releasing hormone in normal subjects and in patients with liver disease and chronic renal failure. Journal of Clinical Endocrinology and Metabolism. 44:356-360.

64.       Clarke IJ, Cummins JT. 1982 The temporal relationship between gonadotropin-releasing hormone (GnRH) and luteinizing hormone (LH) secretion in ovariectomized ewes. Endocrinology. 111:1737-1739.

65.       Levine JN, Pau KYF, Ramirez VD, Jackson GL. 1982 Simultaneous measurement of luteinizing hormone releasing hormone and luteinizing hormone release in unanesthetized, ovariectomized sheep. Endocrinology. 111:1449-1455.

66.       Whitcomb RW, O'Dea LSL, Finkelstein JS, Heavern DM, Crowley Jr WF. 1990 Utility of free alpha-subunit as an alternative neuroendocrine marker of gonadotropin-releasing hormone (GnRH) stimulation of the gonadotroph in the human: evidence from normal and GnRH-deficient men. Journal of Clinical Endocrinology and Metabolism. 70:1654-1661.

67.       Hall JE, Whitcomb RW, Rivier JE, Vale WW, Crowley Jr WF. 1990 Differential regulation of luteinizing hormone, follicle-stimulating hormone, and free à-subunit from the gonadotrope by gonadotropin-releasing hormone (GnRH): Evidence from the use of two GnRH antagonists. Journal of Clinical Endocrinology and Metabolism. 70:328-335.

68.       Hayes FJ, McNicholl DJ, Schoenfeld D, Marsh EE, Hall JE. 1999 Free alpha-subunit is superior to luteinizing hormone as a marker of gonadotropin-releasing hormone despite desensitization at fast pulse frequencies. Journal of Clinical Endocrinology and Metabolism. 1999:1028-1036.

69.       Beitins IZ, Dufau ML, O'Loughlin K, Catt KJ, McArthur JW. 1977 Analysis of biological and immunological activities in the two pools of LH released during constant infusion of luteinizing hormone-releasing hormone (LHRH) in men. Journal of Clinical Endocrinology and Metabolism. 45:605-608.

70.       Balasubramanian R, Dwyer A, Seminara SB et al. 2010 Human GnRH deficiency: a unique disease model to unravel the ontogeny of GnRH neurons. Neuroendocrinology. 92:81-99.

71.       Lewkowitz-Shpuntoff HM, Hughes VA, Plummer L, et al. 2012 Olfactory phenotypic spectrum in idiopathic hypogonadotropic hypogonadism: pathophysiological and genetic implications. J Clin Endocrinol Metab. 97:E136-44

72.       Dode C, Teixera L, Fouveaut C, et al. 2006 Kallmann syndrome: mutations in the genes encoding prokineticin-2 and prokineticin receptor-2. PLoS Genet. October; 2(10): e175.

73.       Pitteloud N, Quinton R, Pearse S, et al 2007 Digenic mutations account for variable phenotypes in idiopathic hypogonadotropic hypogonadism. J Clin Invest. 117:457-463.

74.       Sykiotis G, Plummer L, Hughes VA, et al. 2010 Oligogenic basis of isolated gonadotropin-releasing hormone deficiency. Proc Natl Acad Sci USA. 107:15140-15144.

75.       Bick D, Franco B, Sherins RJ, et al. 1992 Brief report: intragenic deletion of the KALIG-1 gene in Kallmann's syndrome. New England Journal of Medicine. 326:1752-1755.

76.       Edelman GM, Crossin KL. 1991 Cell adhesion molecules: implications for a molecular histology. 60:155-190.

77.       Schwanzel-Fukuda M, Pfaff D. 1989 Origin of luteinizing hormone-releasing hormone neurons. Nature. 338:161-164.

78.       Pitteloud N, Hayes FJ, Boepple PA, DeCruz S, Seminara SB, MacLaughlin DT, Crowley WF Jr.  2002. The role of prior pubertal development, biochemical markers of testicular maturation, and genetics in elucidating the phenotypic heterogeneity of idiopathic hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 87:152-160

79.       Salenave S, Chanson P, Bry H, et al. Kallmann's syndrome: a comparison of the reproductive phenotypes in men carrying KAL1 and FGFR1/KAL2 mutations. J Clin Endocrinol Metab. 93:758-763.

80.       Pitteloud N, Hayes FJ, Dwyer A, et al 2002. Predictors of outcome of long-term GnRH therapy in men with idiopathic hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 87:4128-4136.

81.       Ballabio A, Bardoni B, Carrozzo R, et al. 1989 Continguous gene syndromes due to deletions in the distal short arm of the human X chromosome. Proceedings of the National Academy of Sciences of the United States of America. 86:10001-10005.

82.       Cohen-Salmon M, Tronche F, del Castillo I, Petit C. 1995 Characterization of the promoter of the human KAL gene, responsible for the X-chromosome-linked Kallmann syndrome. Gene. 164:235-242.

83.       Hardelin JP, Levilliers J, Blanchard S, et al. 1993 Heterogeneity in the mutations responsible for X chromosome- linked Kallmann syndrome. Human Molecular Genetics. 2:373-377.

84.       Georgopoulos NA, Pralong FP, Seidman CE, Seidman JG, Crowley Jr WF, Vallejo M. 1997 Genetic heterogeneity evidenced by low incidence of KAL-1 gene mutations in sporadic cases of gonadotropin-releasing hormone deficiency. Journal of Clinical Endocrinology and Metabolism. 82:213-217.

85.       Parenti G, Rizzolo MG, Ghezzi M, et al. 1995 Variable penetrance of hypogonadism in a sibship with Kallmann syndrome due to a deletion of the KAL gene. American Journal of Medical Genetics. 57:476-478.

86.       Quinton R, Duke VM, de Zoysa PA, et al. 1996 The neuroradiology of Kallmann's syndrome: A genotypic and phenotypic analysis. Journal of Clinical Endocrinology and Metabolism. 81:3010-3017.

87.       Bouligand J, Ghervan C, Tello JA, et al. 2009 Isolated familial hypogonadotropic hypogonadism and a GNRH1 mutation. N Engl J Med. 360:2742-2748.

88.       Chang Y-M, de Guillebon A, Lang-Muritano M, et al. 2009 GNRH1 mutations in patients with idiopathic hypogonadotropic hypogonadism. Proc Natl Acad Sci USA. 106: 11703–11708.

89.       Chevrier L, Guimiot F, de Roux N. 2011 GnRH receptor mutations in isolated gonadotropic deficiency. Mol Cell Endocrinol. 346:21-28.

90.       de Roux N, Young J, Misrahi M, et al. 1997 A family with hypogonadotropic hypogonadism and mutations in the gonadotropin-releasing hormone receptor. New England Journal of Medicine. 337:1597-1602.

91.       de Roux N, Young J, Brailly-Tabard S, Misrahi M, Milgrom E, Schaison G. 1999 The same molecular defects of the gonadotropin-releasing hormone receptor determine a variable degree of hypogonadism in affected kindred. Journal of Clinical Endocrinology and Metabolism. 84:567-572.

92.       Pralong FP, Gomez F, Castillo E, et al. 1999 Complete hypogonadotropic hypogonadism associated with a novel inactivating mutation of the gonadotropin-releasing hormone receptor. Journal of Clinical Endocrinology and Metabolism. 84:3811-3816.

93.       Layman LC, Cohen DP, Jin M, et al. 1998 Mutations in gonadotropin-releasing hormone receptor gene cause hypogonadotropic hypogonadism. Nature Genetics. 18:14-15.

94.       Caron P, Chauvin S, Christin-Maitre S, et al. 1999 Resistance of hypogonadic patients with mutated GnRH receptor genes to pulsatile GnRH administration. Journal of Clinical Endocrinology and Metabolism. 84:990-996.

95.       Seminara SB, Beranova M, Oliveira LMB, et al. 2000 Successful use of pulsatile gonadotropin-releasing hormone (GnRH) for ovulation induction and pregnancy in a patient with GnRH receptor mutations. J Clin Endocrinol Metab. 85:556-562.

96.       de Roux N, Genin E, Carel JC, Matsuda F, Chaussain JL, Milgrom E. 2003. Hypogonadotropic hypogonadism due to loss of function of the KiSS1-derived peptide receptor GPR54. Proc Natl Acad Sci USA. 100(19):10972-10976.

97.       Seminara SB, Messager S, Chatzidaki EE, JF, et al. 2003. The GPR54 gene as a regulator of puberty. N Engl J Med. 349:1614-1627.

98.       Chan YM, Broder-Fingert S, Paraschos S, et al. 2011 GnRH-deficient phenotypes in humans and mice with heterozygous variants in KISS1/Kiss1. J Clin Endocrinol Metab. 96:E1771-1781

99.       Dodé C, Levilliers J, Dupont JM, De Paepe A, Le Dû N, Soussi-Yanicostas N, Coimbra RS, Delmaghani S, Compain-Nouaille S, Baverel F, Pêcheux C, Le Tessier D, Cruaud C, Delpech M, Speleman F, Vermeulen S, Amalfitano A, Bachelot Y, Bouchard P, Cabrol S, Carel JC, Delemarre-van de Waal H, Goulet-Salmon B, Kottler ML, Richard O, Sanchez-Franco F, Saura R, Young J, Petit C, Hardelin JP. 2003. Loss-of-function mutations in FGFR1 cause autosomal dominant Kallmann syndrome. Nat Genet. 33(4):463-465.

100.     Pitteloud N, Acierno JS Jr, Meysing A, Eliseenkova AV, Ma J, Ibrahimi OA, Metzger DL, Hayes FJ, Dwyer AA, Hughes VA, Yialamas M, Hall JE, Grant E, Mohammadi M, Crowley WF Jr.  2006. Mutations in fibroblast growth factor receptor 1 cause both Kallmann syndrome and normosmic idiopathic hypogonadotropic hypogonadism. Proc Natl Acad Sci USA. 103(16):6281-6286.

101.     Raivio T, Sidis Y, Plummer L, Chen H, Ma J, Mukherjee A, Jacobson-Dickman E, Quinton R, Van Vliet G, Lavoie H, Hughes VA, Dwyer A, Hayes FJ, Xu S, Sparks S, Kaiser UB, Mohammadi M, Pitteloud N. 2009 Impaired fibroblast growth factor receptor 1 signaling as a cause of normosmic idiopathic hypogonadotropic hypogonadism.  J Clin Endocrinol Metab. 94:4380-4390.

102.     Pitteloud N, Meysing A, Quinton R, Acierno JS Jr, Dwyer AA, Plummer L, Fliers E, Boepple P, Hayes F, Seminara S, Hughes VA, Ma J, Bouloux P, Mohammadi M, Crowley WF Jr. 2006 Mutations in fibroblast growth factor receptor 1 cause Kallmann syndrome with a wide spectrum of reproductive phenotypes. Mol Cell Endocrinol.  254-255:60-69.

103.     Miraoui H, Dwyer A, Pitteloud N. 2011. Role of fibroblast growth factor (FGF) signaling in the neuroendocrine control of human reproduction. Mol Cell Endocrinol. 346:37-43.

104.     Falardeau J, Chung WC, Beenken A, Raivio T, Plummer L, Sidis Y, Jacobson-Dickman EE, Eliseenkova AV, Ma J, Dwyer A, Quinton R, Na S, Hall JE, Huot C, Alois N, Pearce SH, Cole LW, Hughes V, Mohammadi M, Tsai P, Pitteloud N. 2008. Decreased FGF8 signaling causes deficiency of gonadotropin-releasing hormone in humans and in mice. J Clin Invest. 118(8):2822-2831.

105.     Trarbach EB, Abreu AP, Silveira LF, Garmes HM, Baptista MT, Teles MG, Costa EM, Mohammadi M, Pitteloud N, Mendonca BB, Latronico AC.  2010. Nonsense mutations in FGF8 gene causing different degrees of human gonadotropin-releasing hormone deficiency. J Clin Endocrinol Metab. 95:3491-6.

106.     Miraoui H, Dwyer AA, Sykiotis GP, et al. 2013 Mutations in FGF17, IL17RD, DUSP6, SPRY4, and FLRT3 are identified in individuals with congenital hypogonadotropic hypogonadism. Am J Hum Genet. 92(5):725-743.

107.     Dode C, Teixeira L, Levilliers J, et al. 2006 Kallmann syndrome: mutations in the genes encoding prokineticin-2 and prokineticin receptor-2. PLoS Genet 2:1648–1652

108.     Pitteloud N, Zhang C, Pignatelli D, Li JD, Raivio T, Cole LW, Plummer L, Jacobson-Dickman EE, Mellon PL, Zhou QY, Crowley Jr WF 2007 Loss-of-function mutation in the prokineticin 2 gene causes Kallmann syndrome and normosmic idiopathic hypogonadotropic hypogonadism. Proc Natl Acad Sci USA 104:17447–17452

109.     Abreu AP, Trarbach EB, de Castro M, Frade Costa EM, Versiani B, Matias Baptista MT, Garmes HM, Mendonca BB, Latronico AC 2008 Loss-of-function mutations in the genes encoding prokineticin-2 or prokineticin receptor-2 cause autosomal recessive Kallmann syndrome. J Clin Endocrinol Metab 93:4113–4118

110.     Cole LW, Sidis Y, Zhang C, Quinton R, Plummer L, Pignatelli D, Hughes VA, Dwyer AA, Raivio T, Hayes FJ, Seminara SB, Huot C, Alos N, Speiser P, Takeshita A, Van Vliet G, Pearce S, Crowley Jr WF, Zhou QY, Pitteloud N 2008 Mutations in prokineticin 2 (PROK2) and PROK2 receptor 2 (PROKR2) in human gonadotrophin-releasing hormone deficiency: molecular genetics and clinical spectrum. J Clin Endocrinol Metab 93:3551–3559

111.     Sarfati J, Guiochon-Mantel A, Rondard P, et al. 2010. A comparative phenotypic study of Kallmann syndrome patients carrying monoallelic and biallelic mutations in the prokineticin 2 or prokineticin receptor 2 genes. J Clin Endocrinol Metab. 95:659-69.

112.     Topaloglu AK, Reimann F, Guclu M, et al. 2009. TAC3 and TACR3 mutations in familial hypogonadotropic hypogonadism reveal a key role for neurokinin B in the central control of reproduction. Nat Genet. 41(3):354-358.

113.    Gianetti E, Tusset C, Noel SD, et al. 2010. TAC3/TACR3 mutations reveal preferential activation of gonadotropin-releasing hormone release by neurokinin B in neonatal life followed by reversal in adulthood. J Clin Endocrinol Metab. 95(6):2857-2867.

114.    Francou B, Bouligand J, Voican A, et al. 2011.  Normosmic congenital hypogonadotropic hypogonadism due to TAC3/TACR3 mutations: characterization of neuroendocrine phenotypes and novel mutations. PLoS One. 6(10):e25614.

115.     Salvi R, Pralong FP. Molecular characterizyation and phenotypic expression of mutations in genes for gonadotropins and their receptors in humans. Front Horm Res. 2010;39:1-12.

116.     Weiss J, Axelrod L, Whitcomb RW, Harris PE, Crowley Jr WF, Jameson JL. 1992 Hypogonadism caused by a single amino acid substitution in the beta subunit of luteinizing hormone. N Engl J Med. 326:179-183.

117. Lindstedt G, Nystrom E, Matthews C, Ernest I, Janson PO, Chatterjee K. 1998 Follitropin (FSH) deficiency in an infertile male due to FSH beta gene mutation. A syndrome of normal puberty and virilization but underdeveloped testicles with azoospermia, low FSH but high lutropin and normal serum testosterone concentrations. Clin Chem Lab Med. 36:663-5.

118. Phillip M, Arbelle JE, Segev Y, Parvari R. 1998 Male hypogonadism due to a mutation in the gene for the beta subunit of follicle stimulating hormone. New Engl J Med. 338:1729-1732.

119. Muscatelli F, Strom TM, Walker AP, et al. 1994.  Mutations in DAX-1 gene give rise to both X-linked adrenal hypoplasia congenita and hypogonadotrophic hypogonadism. Nature. 372:672-676.

120. Jadhav U, Harris RM, Jameson JL. 2011. Hypogonadotropic hypogonadism in subjects with DAX1 mutations. Mol Cell Endocrinol. 346:65-73.

121. Kim HG, Kurth I, Lan F, Meliciani I, et al. 2008. Mutations in CHD7, encoding a chromatin remodeling protein, cause idiopathic hypogonadotropic hypogonadism and Kallmann syndrome. Am J Hum Genet. 83:511-519.

122. Jongmans MC, van Ravenswaaij-Arts CM, Pitteloud N, et al. 2008. CHD7 mutations in patients initially diagnosed with Kallmann syndrome: the clinical overlap with CHARGE syndrome. Clin Genet. 75(1):65-71.

123. Montague CT, Farooqi IS, Whitehead JP, et al. 1997 Congenital leptin deficiency is associated with severe early-onset obesity in humans. Nature. 387:903-908.

124. Farooqi IS, Wangensteen T, Collins S, et al. 2007.  Clinical and molecular genetic spectrum of congenital deficiency of the leptin receptor. N Engl J Med. 356(3):237-47.

125.     Farooqi IS, Volders K, Stanhope R, et al.  2007 Hyperphagia and early-onset obesity due to a novel homozygous missense mutation in prohormone convertase 1/3.  J Clin Endocrinol Metab. 92(9):3369-73.

126.     Parks JS, Brown MR, Hurley DL, Phelps CJ, Wajnrajch MP. 1999  Heritable disorders of pituitary development. J Clin Endocrinol Metab. 84:4362-4370.

127.     Fluck C, Deladoey J, Rutishauser K, et al. 1998 Phenotypic variability in familial combined pituitary hormone deficiency caused by a PROP1 gene mutation resulting an the substitution of ArgàCys at Codon120 (R120C). J Clin Endocrinol Metab 83 :3727-3734.

128.     Deladoey J, Fluck C, Buyukgebiz A, et al. 1999 ‘Hot spot’ in the PROP1 gene responsible for combined pituitary hormone deficiency. J Clin Endocrinol Metab. 84:1645-50

129.     McCabe MJ, Gaston-Massuet C, Gregory LC, et al. 2013 Variations in PROKR2, but not PROK2, are associated with hypopituitarism and septo-optic dysplasia. J Clin Endocrinol Metab. 98(3):E547-57.

130.     Raivio T, Avbelj M, McCabe MJ, et al. 2012 Genetic overlap in Kallmann syndrome, combined pituitary hormone deficiency, and septo-optic dysplasia. J Clin Endocrinol Metab. 97:E694-9.

131.     Sato N, Kamachi Y, Kondoh H, Shima Y, Morohashi K, Horikawa R, Ogata T. 2007. Hypogonadotropic hypogonadism in an adult female with a heterozygous hypomorphic mutation of SOX2. Eur J Endocrinol 156:167-171

132.     Stark Z, Storen R, Bennetts B, Savarirayan R, Jamieson RV. 2011. Isolated hypogonadotropic hypogonadism with SOX2 mutation and anopthalmia/microphtalmia in offspring. Eur J Hum Genet. 19(7):753-6.

133.     Pingault V, Bodereau V, Baral V, et al. 2013 Loss-of-Function mutations in SOX10 cause Kallmann Syndrome with deafness. Am J Hum Genet. 92(5):707-24.

134.     World Health Organization Laboratory Manual for the evaluation and processing of human semen. 5th Edition (2010) CambridgeUniversity Press, Cambridge.

135.     Costa-Barbosa FA, Balasubramanian R, Keefe KW, Shaw ND, Al-Tassan N, Plummer L, Dwyer AA, Buck CL, Choi JH, Seminara SB, Quinton R, Monies D, Meyer B, Hall JE, Pitteloud N, Crowley WF Jr. 2013 Prioritizing genetic testing in patients with Kallmann syndrome using clinical phenotypes. J Clin Endocrinol Metab. 98(5):E943-953.

136.     Canaves F, Mussa A, Manenti M, et al. 2009 Sperm count of young men surgically treated for cryptorchidism in the first and second year of life: fertility is better in children treated at a younger age. Eur J Paedr Surg. 19 :388-391.

137.     Raivio T, Falardeau J, Dwyer A, Quinton R, Hayes FJ, Hughes VA, Cole LW, Pearce SH, Lee H, Boepple P, Crowley WF Jr, Pitteloud N.  2007 Reversal of idiopathic hypogonadotropic hypogonadism. N Engl J Med 357(9):863-873.

138.     Laitinen EM, Tommiska J, Sane T, Vaaralahti K, Toppari J, Raivio T. 2012 Reversible congenital hypogonadotropic hypogonadism in patients with CHD7, FGFR1 or GNRHR mutations. PLoS One. 2012;7(6):e39450.

139.     Tapanainen JS, Aittomaki K, Min J, Vaskivuo T, Huhtaniemi IT. 1997 Men homozygous for an inactivating mutation of the follicle-stimulating hormone (FSH) receptor gene present variable suppression of spermatogenesis and fertility. Nature (London). 15:205-206.

140.     Burris AS, Rodbard HW, Winters SJ, Sherins RJ. 1988 Gonadotropin therapy in men with isolated hypogonadotropic hypogonadism: the response to human chorionic gonadotropin is predicted by initial testicular size. J Clin Endocrinol Metab. 66:1144-51.

141.     Behre HM, Baus S, Kliesch S, Keck C, Simoni M, Nieschlag E. 1995 Potential of testosterone buciclate for male contraception: endocrine differences between responders and nonresponders. J Clin Endocrinol Metab. 80:2394-2403.

142.     O'Donnell L, Narula A, Balourdos G, et al. 2001 Impairment of spermatogonial development and spermiation after testosterone-induced gonadotropin suppression in adult monkeys (Macaca fascicularis). J Clin Endocrinol Metab. 86:1814-1822.

143.     Gromoll J, Simoni M, Nieschlag E. 1996 An activating mutation of the follicle-stimulating hormone receptor autonomously sustains spermatogenesis in a hypophysectomized man. J Clin Endocrinol Metab. 81:1367-1370.

144.     Raivio T, Wikström AM, Dunkel L. 2007. Treatment of gonadotropin-deficient boys with recombinant human FSH: long-term observation and outcome. Eur J Endocrinol. 156(1):105-111.

145.     Dwyer AA, Sykiotis GP, Hayes FJ, Poepple PA, Lee H, Loughlin KR, Dym M, Sluss PM, Crowley WF JR, Pitteloud N.  2013 Trial of recombinant follicle stimulating hormone (rFSH) pre-treatment for GnRH-induced fertility in patients with congenital hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 98:E1790-1795.

146.     Schaison G, Young J, Pholsena M, Nahoul K, Couzinet B. 1993 Failure of combined follicle-stimulating hormone-testosterone administration to initiate and/or maintain spermatogenesis in men with hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 77:1545-1549.

147.     Recombinant Human FSH Product Development Group 1998 Recombinant follicle stimulating hormone: development of the first biotechnology product for the treatment of infertility. Hum Reprod Update. 4:862-81.

148.     Mannaerts B, Fauser B, Lahlou N, et al. 1996 Serum hormone concentrations during treatment with multiple rising doses of recombinant follicle stimulating hormone (Puregon) in men with hypogonadotropic hypogonadism. Fertil Steril. 65:406-410.

149.     King TFJ, Hayes FJ. 2012 Long-term outcome of idiopathic hypogonadotropic hypogonadism. Curr Opin Endocrinol Diabetes Obes. 19:204-210.

150.     Vicari E, Mongioì A, Calogero AE, Moncada ML, Sidoti G, Polosa P, D'Agata R. 1992. Therapy with human chorionic gonadotrophin alone induces spermatogenesis in men with isolated hypogonadotrophic hypogonadism--long-term follow-up. Int J Androl. 15:320-329.

151.     Büchter D, Behre HM, Kliesch S, Nieschlag E. 1998. Pulsatile GnRH or human chorionic gonadotropin/human menopausal gonadotropin as effective treatment for men with hypogonadotropic hypogonadism: a review of 42 cases. Eur J Endocrinol. 139:298-303.

152.     Liu PY, Baker HW, Jayadev V, Zacharin M, Conway AJ, Handelsman DJ. 2009 Induction of spermatogenesis and fertility during gonadotropin treatment of gonadotropin-deficient infertile men: predictors of fertility outcome. J Clin Endocrinol Metab. 94:801-808.

153.     Warne DW, Decosterd G, Okada H, Yano Y, Koide N, Howles CM. 2009 A combined analysis of data to identify predictive factors for spermatogenesis in men with hypogonadotropic hypogonadism treated with recombinant human follicle-stimulating hormone and human chorionic gonadotropin. Fertil Steril. 92:594-604.

154.     Kulin HE, Samojlik E, Santen R, Santner S. 1981 The effect of growth hormone on the Leydig cell response to chorionic gonadotrophin in boys with hypopituitarism. Clin Endocrinol (Oxf). 15:463-472.

155.     Chatelain PG, Sanchez P, Saez JM. 1991 Growth hormone and insulin-like growth factor I treatment increase testicular luteinizing hormone receptors and steroidogenic responsiveness of growth hormone deficient dwarf mice. Endocrinology. 128:1857-62.

156.     Hoffman AR, Crowley WF, Jr. 1982. Induction of puberty in men by long-term pulsatile administration of low-dose gonadotropin-releasing hormone. N Engl J Med. 307: 1237-1241.

157.     Spratt DI, Crowley Jr WF, Butler JP, Hoffman AR, Conn PM, Badger TM. 1985 Pituitary luteinizing hormone responses to intravenous and subcutaneous administration of gonadotropin releasing hormone in men. Journal of Clinical Endocrinology and Metabolism. 61:890-895.

158.     Pitteloud N, Hayes FJ, Dwyer A, Boepple PA, Lee H, Crowley WF Jr. 2002. Predictors of outcome of long-term GnRH therapy in men with idiopathic hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 87(9):4128-4136.

159.     Liu L, Banks SM, Barnes KM, Sherins RJ. 1988 Two-year comparison of testicular responses to pulsatile gonadotropin-releasing hormone and exogenous gonadotropins from the inception of therapy in men with isolated hypogonadotropic hypogonadism. J Clin Endocrinol Metab. 67:1140-1145.

160.     Bonduelle M, Wennerholm UB, Loft A, Tarlatzis BC, Peters C, Henriet S, Mau C, Victorin-Cederquist A, Van Steirteghem A, Balaska A, Emberson JR, Sutcliffe AG. A multi-centre cohort study of the physical health of 5-year-old children conceived after intracytoplasmic sperm injection, in vitro fertilization and natural conception. Hum Reprod. 2005;20(2):413.

161.     Belva F, Henriet S, Liebaers I, Van Steirteghem A, Celestin-Westreich S, Bonduelle M. Medical outcome of 8-year-old singleton ICSI children (born>or=32 weeks' gestation) and a spontaneously conceived comparison group. Hum Reprod. 2007;22(2):506.

162.    Whitelaw N, Bhattacharya S, Hoad G, Horgan GW, Hamilton M, Haggarty P. Epigenetic status in the offspring of spontaneous and assisted conception.Hum Reprod. 2014 Jul;29(7):1452-8

163.     Hansen M, Bower C, Milne E, de Klerk N, Kurinczuk JJ. Assisted reproductive technologies and the risk of birth defects—a systematic review. Hum Reprod 2005;20:328-338.

164.     Reefhuis J, Honein MA, Schieve LA, Correa A, Hobbs CA, Rasmussen SA, National Birth Defects Prevention Study.  2009. Assisted reproductive technology and major structural birth defects in the United States. Hum Reprod. 24(2):360-366.

165.     Bukulmez O. 2009 Does assisted reproductive technology cause birth defects? Curr Opin Obstet Gynecol. 2009 Jun;21(3):260-264.

Test Chaper

Test Audio Narration

Below is a sample narration from The Adventures of Huckleberry Finn. The audio file can be started when the "Play" button is clicked. Once the audio has started, the player indicates the position on the audio and has volume control.

Figure 4.Mechanisms of Pharmacologic Interference with Catecholamines and Metanephrines (164). Monoamine oxidase- (MAO), dihydroxyphenylglycol- (DHPG), DOPA- dihydroxyphenylalanine.

Figure 4.Mechanisms of Pharmacologic Interference with Catecholamines and Metanephrines (164). Monoamine oxidase- (MAO), dihydroxyphenylglycol- (DHPG), DOPA- dihydroxyphenylalanine.