Archives

Evaluation and Treatment of Polycystic Ovary Syndrome

Abstract

Polycystic ovary syndrome (PCOS) is the most common endocrinopathy among adult women in the developed world and is characterized by anovulation, androgen excess (primarily ovarian, but also adrenal in origin) and the appearance of polycystic ovaries on ultrasound.   Diagnostic criteria are expert-based and debated as they do not incorporate known metabolic abnormalities related to aberrant insulin action, such as glucose intolerance, diabetes, and dyslipidemia, that affect many women with the syndrome. Symptoms that are most troublesome to patients include hirsutism, obesity, infertility and menstrual disorders. Long-term sequelae of the syndrome, such as an increased risk for cardiovascular events based on risk factor profiling, are unclear from epidemiologic studies. The etiology of the syndrome is likely heterogeneous and genetic studies have been consistent with a complex genetic disease,.   Interestingly, the candidate genes identified in multiple genome wide association studies that fit best into existing ideas of the pathophysiology are gonadotropin and gonadotropin receptor genes.   Treatment tends to be symptom based, and the search for a single treatment that addresses both reproductive and metabolic abnormalities continues. Some of the most common treatments used for chronic management of PCOS include hormonal contraceptives, progestins   and metformin. Treatment of infertility focuses on ovulation induction therapies which may involve drugs such as letrozole or clomiphene or gonadotropin therapy.   Treatment of hirsutism often involves the combination of hormonal contraceptives and the adjuvant use of anti-androgens. Weight loss in obese women with PCOS may be beneficial for both the treatment of infertility and long term management.  For  complete coverage of this and related areas of ENDOCRINOLOGY, please see our FREE on-line web-book, www.endotext.org.

 

 

Introduction

Polycystic ovary syndrome (PCOS) is an ovarian disorder characterized by hyperandrogenism, ovulatory dysfunction, and polycystic ovaries. It may be the most common female endocrinopathy in the developed world.   However, it most likely represents a heterogeneous disorder and one whose pathophysiology and etiology are debated.     PCOS affects young women with oligo-ovulation (which can lead to oligomenorrhea), infertility, acne and hirsutism. It also has notable metabolic sequelae, including an elevated risk of diabetes and cardiovascular risk factors, and long term treatment should also consider these factors. These multiple stigmata have led to a multi-pronged treatment approach, with most therapies targeting individual symptoms.   The search for the single unifying theory to this disorder will hopefully yield the single best treatment, but this quest remains one of the Holy Grails of reproductive endocrinology. This chapter will discuss the diagnosis, clinical evaluation, pathophysiology, and treatment of women with PCOS.

 

Diagnostic Criteria

There is no universally accepted definition of PCOS and expert generated diagnostic criteria have proliferated in recent years (Figure 1). They share a common focus on PCOS as an ovarian disorder. The definition of PCOS has largely been dependent on the technology used to ascertain the condition. Thus the earliest definition of PCOS, or the Stein Leventhal Syndrome, was based on the triad of enlarged ovaries, hirsutism, and oligomenorrhea (1).     As assays became available, first urinary and then serum, researchers noted gonadotropin abnormalities with elevated LH levels, and then as androgen assays evolved, elevation in androgen levels.   However, these multiple tools to assess women with androgen excess and oligomenorrhea led to multiple diagnostic criteria (Figure 1) and often each investigative group had their own unique set, making the comparison of clinical studies often difficult if not impossible.

Figure 1: Recommended diagnostic schemes for PCOS by varying expert groups. All recommend excluding possible other etiologies of these signs/symptoms (See Differential Diagnosis) and more than one of the signs or symptoms must be present to make a diagnosis. Red box - not required for diagnosis; black box - mandatory criteria; white box - possible diagnostic criteria but not necessarily required to be present. Hyperandrogenism may be either the presence of hirsutism or biochemical hyperandrogenemia.

Figure 1: Recommended diagnostic schemes for PCOS by varying expert groups. All recommend excluding possible other etiologies of these signs/symptoms (See Differential Diagnosis) and more than one of the signs or symptoms must be present to make a diagnosis. Red box - not required for diagnosis; black box - mandatory criteria; white box - possible diagnostic criteria but not necessarily required to be present. Hyperandrogenism may be either the presence of hirsutism or biochemical hyperandrogenemia.

It was not until the early 1990s at an NIH-sponsored conference on PCOS that formal diagnostic criteria were proposed and afterwards were largely utilized (2).   These criteria, often referred to colloquially as “the NIH criteria” were published in the conference proceedings and received large scale acceptance in the research and clinical communities. These criteria defined PCOS as unexplained hyperandrogenic anovulation. They required the presence of oligomenorrhea AND hyperandrogenism, either clinical or biochemical along with the exclusion of phenocopies.   The enduring portion of these criteria accepted by all other criteria was the exclusion of phenocopies such that PCOS remains a diagnosis of exclusion.

 

The improved technology and utilization of ultrasound in women’s health led to the ultrasound definition of polycystic ovaries, defined primarily on the morphology and the number of small antral follicles (3) (Figure 2).    The failure to recognize the polycystic ovary in the NIH definition of polycystic ovary syndrome led to the convening of an expert consensus conference to reconsider the NIH diagnostic criteria in Rotterdam in the Netherlands. The subsequent “Rotterdam criteria” incorporated the ultrasound determined size and morphology of the ovary into the diagnostic criteria (4,5).   Ultrasound criteria for the diagnosis of polycystic ovaries were also decided by expert consensus (6), though the cutoff for antral follicles was recently raised again by expert opinion due to improvements in resolution allowing increased follicle detection (7) (Table 1). Because of the limited availability of ultrasound and trained ultrasonographers in many practices (family practice, medical or pediatric endocrinology) as well as in low resource settings, there has been interest in using Anti-Mullerian Hormone (AMH) levels to diagnose polycystic ovaries in lieu of ultrasound.(8) AMH is produced by the granulosa cells of small antral follicles and correlates well with their count.(9) Currently, however, there is no accepted cutoff and there are similar concerns about the effects of age and hormonal contraception on this parameter as antral follicle counts. The Rotterdam criteria have been criticized for including more mild phenotypes, for example, the combination of polycystic ovaries with oligomenorrhea. These additional phenotypes may complicate the generalizability of clinical trials to treat PCOS, and may also elevate the prevalence of PCOS in the general population.

Figure 2: Transvaginal ultrasound of a polycystic ovary. Note the increased number of antral follicles ringing the outside of the ovary and the increased central stroma.

Figure 2: Transvaginal ultrasound of a polycystic ovary. Note the increased number of antral follicles ringing the outside of the ovary and the increased central stroma.

Table 1: Expert consensus recommendations for the ultrasound diagnosis of polycystic ovaries. The ultrasound exam assumes that if there is a follicle > 10 mm the scan should be repeated during a period of ovarian quiescence in order to calculate the ovarian volume.

2003 ASRM/ESHRE Consensus 2014 AE-PCOS Consensus
Follicles 12 or more follicles measuring 2-9 mm in diameter 25 or more follicles < 10 mm in diameter per ovary
Volume > 10 Cm3 > 10 Cm3

 

 

The Androgen Excess Society criteria subsequently attempted to establish hyperandrogenism as a sine qua non diagnostic factor in combination with other stigmata of the syndrome (10). The focus on hyperandrogenism was to eliminate milder phenotypes and based on evidence that hyperandrogenism tends to track with both reproductive (i.e., acne, hirsutism, and androgenic alopecia) and metabolic (i.e., insulin resistance, dyslipidemia, and elevated cardiovascular risk) stigmata of the syndrome.

 

Although the diagnostic criteria are still debated today, recent consensus statements such as the Endocrine Society’s clinical practice guidelines(11) and the NIH Evidence-Based Methodology Executive Summary recommended maintaining the Rotterdam Criteria for PCOS.(12) The latter group did suggest that re-naming the disorder would better focus interest on the diverse implications of the syndrome,(12)   and some authors have advocated re-naming the syndrome based on its metabolic underpinnings.(13)

 

There are, however, unifying trends to all diagnostic criteria. Hyperandrogenism in all schemas can be established on the basis of clinical findings (e.g., hirsutism or acne) and/or serum hormone measurement (most commonly serum testosterone levels). All diagnostic schemes recommend that secondary causes (such as adult-onset congenital adrenal hyperplasia, hyperprolactinemia, and androgen-secreting neoplasms) should first be excluded (discussed below under differential diagnosis).   All diagnostic schemes also require more than one sign or symptom. Polycystic ovaries alone, for example, are a nonspecific finding and also are frequently noted in women with no endocrine or metabolic abnormalities (14), especially among normal healthy younger women.(15) Insulin resistance has been noted consistently among many women with PCOS, especially in those with hyperandrogenism, but it is not included in any of the diagnostic criteria.

 

We can conclude that there is a thread of continuity between the varying diagnostic criteria. All agree that it is an ovarian disorder and diagnostic criteria revolve around ovarian determined stigmata, such as hyperandrogenism, oligo-ovulation, and polycystic-appearing ovaries.   The utility of the varying diagnostic criteria is still being debated by experts, but will ultimately be answered by well-designed clinical studies.   There are no diagnostic criteria that are accepted for diagnosing PCOS in pre-pubertal or peri-pubertal girls nor in menopausal women and the relative androgen excess and oligo/amenorrhea that characterize these states likely overlap too much to separate out groups of affected from unaffected women.

 

Incidence of PCOS

The incidence of PCOS varies according to the diagnostic criteria. Polycystic ovaries on ultrasound are noted in up to 25%-30% of reproductive aged women (14,16).   Thus, the vast majority of women with polycystic ovaries do not have the syndrome. Women with unexplained hyperandrogenic chronic anovulation (i.e., NIH criteria) make up approximately 7% of reproductive aged women (17).   There is debate as to whether minorities are disproportionately affected with PCOS (18). Other studies, for example, have shown higher rates of insulin resistance and type 2 diabetes in minorities, including Latino, Native American, and African-American populations. However, the evidence for this in women with PCOS is less certain. For example, in the best study of an unselected population in the U.S., i.e., women applying for jobs at an academic medical center, there were no significant differences in the prevalence of PCOS or stigmata of PCOS, such as hirsutism or elevated circulating androgen levels between white and black women (17).     The broader Rotterdam criteria increase the prevalence of PCOS by 50% over the NIH criteria (19), and the prevalence according to the AES criteria is somewhere in between.

 

PCOS is increasingly associated with obesity, and the obesity epidemic worldwide has been linked to an increased prevalence of PCOS (20). There are still marked differences in the prevalence of obesity and morbid obesity among women with PCOS according to country of origin as noted in Figure 3. Obesity, and severe obesity appear to be less common in the European PCOS population (21) and in Asia.(22,23) It appears from the published literature that the U.S. tends to have the highest prevalence of severe obesity in its population and its PCOS population.   Large multi-center trials of women with PCOS and infertility routinely report a mean BMI of 35 among study participants(24,25). While it is debated whether obesity per se can cause PCOS, there are mixed data supporting an increased population based prevalence of PCOS with increasing obesity.(26,27)   Interestingly a European Genome Wide Association Study (GWAS) identified increasing BMI as a risk factor through Mendelian randomization analysis. (28)

Figure 3: Distribution (counts on y axis) of BMI in women with PCOS from a large cohort of women diagnosed with PCOS in the United Kingdom ( N = 1741)(21) compared to that from a cohort in the United States ( N = 398) (Legro, unpublished data). Compare a mode of BMI of 20 for the UK women with a BMI of 35 for the U.S. women.

Figure 3: Distribution (counts on y axis) of BMI in women with PCOS from a large cohort of women diagnosed with PCOS in the United Kingdom ( N = 1741)(21) compared to that from a cohort in the United States ( N = 398) (Legro, unpublished data). Compare a mode of BMI of 20 for the UK women with a BMI of 35 for the U.S. women.

Etiology and Pathophysiology

The genetic contribution to PCOS has been closely studied in recent years with multiple Genome Wide Association Studies (GWAS) reporting results in European(28,29) and Han Chinese Cohorts (22,23). However, these GWAS’s have only identified candidate gene regions that explain a small proportion of the heritability of PCOS (i.e., less than 10%). There is currently no recommended genetic screening test for PCOS. While genes likely contribute to the PCOS phenotype, the GWAS findings support that PCOS is a complex genetic trait. Interestingly the GWASs from all groups have identified regions of the genome associated with gonadotropin production or action (i.e., gonadotropin receptors), supporting a primary hypothalamic-pituitary dysfunction in the etiology of the syndrome.

 

However, many of the significant GWAS genes or regions do not have a clear functional relationship to the clinical presentation of PCOS.     An example is the DENND1A gene, which encodes a protein named connecdenn 1, which has a clathrin-binding domain and is thought to facilitate endocytosis and receptor mediated turnover, including of gonadotropin and insulin receptors.   A variant of this gene over expressed in human thecal cells created excess androgen production and could be knocked down to restore a normoandrogenic phenotype.(30) There have not been studies to date on the effects of this variant on insulin action.

 

No specific environmental substance has been identified as causing PCOS, although certain medications such as valproate have been shown in vitro (31) or in clinical series in women with epilepsy to induce hyperandrogenism (32). Obesity, however, likely increases its prevalence (as noted above and discussed in pathophysiology below). There has been much interest and suspicion that environmental disrupting chemicals (EDCs) may also contribute to PCOS. However, the data are sparse, although there have been reports of an association between PCOS and elevated levels of Bisphenol A (BPA). However, such association studies are similar to the early genetic association studies examining single alleles with a disorder in which there was a high rate of false positive associations, not replicated in larger studies or studies of multiple variants. When multiple EDCs are measured, the associations become more difficult to interpret.(33)

 

There are three common theories for the etiology of PCOS: one that it is due to hypothalamic-pituitary dysfunction, the second that it is due to ovarian (and adrenal) hyperandrogenism, and the third that it is primarily a disorder of peripheral insulin resistance.   We will address each theory in turn.

 

Primary Disordered Gonadotropin Secretion. The first biochemical abnormality that was identified in women with PCOS was disordered gonadotropin secretion, with a preponderance of LH relative to FSH.   As the two-cell theory of the ovary evolved, i.e., that thecal cells can only produce androgens under stimulation of LH whereas granulosa cells can only aromatize androgens from the theca cells into estrogens under the influence of FSH, this preponderance of LH was thought to be the primary etiology of the syndrome.   Excess LH led to excess thecal cell development and androgen production, but in the face of inadequate FSH stimulation of granulosa cell development and aromatase production, these androgens were not converted to estrogen, leading to multiple abnormalities. The GWAS studies which have identified the FSH and LH/hCG receptor genes as potential contributors to the PCOS phenotype support this etiologic claim.

 

This theory explained the morphology of the ovary, hirsutism, and anovulation. Androgen excess led to ovarian follicular arrest in the preantral stage, as estrogen is critical to the development and selection of a dominant follicle. The ovary thus contained multiple small preantral follicles due to this ongoing process and increased central stroma due to excessive thecal and stromal hyperplasia from the disordered gonadotropin exposure. Secondarily this resulted in spillover of excess androgens into the circulating pool, resulting in inappropriate feedback to the hypothalamic-pituitary axis and a vicious feedback loop where excess LH leads to excess ovarian androgen production which in turn leads to further LH.   Finally the excess circulating androgen led to stimulation of the pilosebaceous unit, increasing sebum production, inducing terminal hair differentiation, and in rare instances in the scalp leading to pilosebaceous unit atresia and androgenic alopecia.

 

Studies of gonadotropin secretion in women with PCOS have established that women have augmented release of LH in response to a GnRH challenge with appropriate increases in FSH secretion (34).     This has led to the use of a GnRH challenge test to diagnose PCOS by some investigative groups (35); however, this requires blood tests up to 24h after the challenge and is unwieldy in a clinical setting. Similarly, random serum samples of LH tend to have poor sensitivity and specificity for diagnosing PCOS (36). This is because of the variability of serum levels due to the pulsatile secretion of the hormones and is also due to modifying factors such as concurrent medications and conditions, most importantly obesity.

 

Obesity tends to blunt baseline LH levels and GnRH stimulated levels in women with PCOS (37), although their responses remains elevated when compared to appropriate age and weight matched control women. The ontogeny of disordered gonadotropin secretion may lie in the hyperandrogenemia of puberty, as the GnRH pulse generator shows an insensitivity to progesterone feedback in hyperandrogenic obese adolescent girls, thus perpetuating the state of disordered gonadotropin secretion.(38,39)

 

Primary Ovarian and Adrenal Hyperandrogenism. Because most diagnostic criteria support the notion that PCOS is an ovarian disorder, it becomes therefore the prime target for the cause of the syndrome.   Ovarian steroidogenesis is perturbed in the syndrome with increased circulating androgen levels frequently noted in women with stigmata of PCOS. Further intrafollicular androgen levels tend to be elevated in antral follicles, supporting a lack of adequate granulosa aromatase activity (40).   As noted above, a primary defect in ovarian steroidogenesis could lead through the same feedback loop noted above to disordered gonadotropin secretion and stigmata of peripheral hyperandrogenism. including acne, hirsutism, and alopecia. Thecal cells from women with PCOS put into long term culture exhibit defects in steroidogenesis. including hyperproduction of androgens, implying this is a permanent and possibly genetic defect in the cells (41).   Family studies also support a high prevalence of hyperandrogenemia and hyperandrogenism in first degree relatives of women with PCOS (42-44), further supporting a familial contribution to these stigmata.     Finally, 20-30% of women with PCOS have evidence of adrenal hyperandrogenism, primarily based on elevated levels of DHEAS, an androgen marker of adrenal function (45), suggesting that the defect in steroidogenesis is primary and affects both androgen secreting glands, i.e., the ovary and the adrenal.   Further there is familial clustering of elevated DHEAS levels in PCOS families in both female and male relatives, again supporting a heritable component to this trait (42,44,46).

 

However, to date, no specific genetic abnormality in the GWAS studies has been noted in steroidogenic enzymes or factors to explain the hyperandrogenism (47) (22,23,28,29). Further it is simplistic to imply that this defect is permanent. First, at least in terms of phenotype and androgen levels, it does not manifest till menarche and appears to resolve with menopause, implying this is not a constitutional phenotypic characteristic. Second, hyperandrogenism can be ameliorated by treatment with suppressive hormonal therapies or conversely with ovulation induction. Polycystic ovaries are a recognized risk factor for ovarian hyperstimulation characterized by multiple and excessive follicular development, elevated circulating levels, and after exposure to human chorionic gonadotropin, massive ovarian enlargement, vascular permeability, and accumulation of abdominal ascites.   This response appears consistent more with baseline inhibition of certain aspects of steroidogenesis combined with exaggerated ovarian response to a given challenge than a primary defect in steroidogenesis per se.

 

Other ovarian factors than disordered steroidogenesis may contribute to PCOS. For example there appears to be an increased density of small preantral follicles in polycystic ovaries (48). This could result from increased numbers of germ cells in the fetal ovary, from decreased loss of oocytes with age, or from decreased rate of loss of oocytes during late gestation, childhood, and puberty. Indeed, there is evidence in vitro to support increased survival and diminished atresia of PCOS follicles (49).

 

Primary Disorder of Insulin Resistance. Women with PCOS show multiple abnormalities in insulin action.   Dynamic studies of insulin action, including hyperinsulinemic euglycemic clamps and frequently sampled intravenous glucose tolerance tests, have shown that women with PCOS are more insulin resistant than weight-matched control women, a defect primarily present in skeletal muscle (50,51).   Early in the ontogeny of the syndrome, as in the ontogeny of type 2 diabetes, this is characterized by increased pancreatic beta cell production of insulin to control ambient glucose levels.   Thus many women with PCOS have fasting and meal-challenged hyperinsulinemia (52). However, this compensatory response by the pancreatic beta cell is often inadequate for the degree of peripheral insulin resistance leading initially to postpandrial hyperglycemia in these women and eventually to fasting hyperglycemia (51,53).   Further the beta cell response appears to be dysschronous, implying a further beta cell defect in these women, and is responsive to treatment with insulin sensitizing agents such as thiazolidinediones (54).

 

Hyperinsulinemia and/or disordered insulin action may perturb the reproductive axis in multiple ways. First insulin may act at the hypothalamic-pituitary axis to stimulate gonadotropin production. Infusions of insulin tend to have little effect on gonadotropin production in human studies, and insulin is not required for glucose transport into the nervous system.   In animal cell culture models, insulin has been found to enhance pituitary gonadotropin secretion (55). However, selective knock out of the insulin receptor in mouse models (the NIRKO mouse) exhibits increased food intake and fat mass, and an exaggerated response to GnRH stimulation (though their basal state in contrast to women with PCOS tends to be hypogonadotropic hypogonadism) (56).   Thus, the evidence for a central action of insulin may be the weakest link in the insulin resistance PCOS hypothesis in humans, perhaps because it is the most difficult to investigate.

 

Hyperinsulinemia is linked to ovarian and adrenal hyperandrogenism in a number of disorders of inherited insulin resistance with compensatory hyperinsulinemia including leprechaunism, the Rabson Mendenhall syndrome, and the lipodystrophies (57). These syndromes are characterized by selective tissue atrophy due to inability to utilize the primary anabolic hormone, insulin, and by excess gonadal androgen production. A less severe insulin resistance syndrome, the HAIR-AN syndrome, was defined on the basis of hyperinsulinemia, hyperandrogenemia, and the presence of acanthosis nigricans (a hyperproliferative skin condition found in skin folds due to insulin excess) and is more common (58).

 

This link between hyperinsulinism and hypergonadism is thought to reflect the ability of insulin in certain conditions to stimulate gonadal and adrenal androgen production. Hyperandrogenism has been further linked to insulin resistance and stigmata of the insulin resistance syndrome in women with PCOS and in family studies of those with PCOS which have found increasing prevalence of the metabolic syndrome in family members with increasing androgen levels (59).   Androgens also induce insulin resistance, best illustrated by the example of female-to-male transsexuals who have increased insulin resistance after supplementation with androgens (60). In vitro cultures of thecal cells from women with PCOS have been found to overproduce androgens in response to insulin supplementation (61). Further, as discussed below under therapeutic options, the use of insulin sensitizing agents, including both metformin and troglitazone have been associated with both lowering of circulating insulin levels and levels of both adrenal and ovarian androgens.

 

Finally, increased levels of insulin are associated with the peripheral availability of sex steroids through an impact on circulating sex hormone-binding globulin (SHBG). SHBG has been found to be partially regulated by circulating insulin levels with an inverse relationship (62). Decreasing levels of SHBG mean increasing levels of free and bioavailable androgens, especially since the preferred substrate of SHBG is androgens (as opposed to estrogen or progestin).  Increased free androgens mean increased androgen action in the periphery, which can affect the pilosebaceous unit and the hypothalamic-pituitary axis. Some have recommended that low circulating SHBG levels may be a good marker for women with PCOS as hyperandrogemia can also suppress SHBG.(63) Thus insulin resistance can contribute to hyperandrogenism in many ways (64).

Obesity per se is associated with insulin resistance and compensatory hyperinsulinemia. Obese women may be ovulatory but have longer follicular phases and thus longer menstrual cycles which could cause them to be misdiagnosed as oligo-ovulatory.(65) Similarly, as noted above, obesity may suppress circulating SHBG levels, leading to higher levels of free or bioavailable testosterone and leading again to the potential misdiagnosis of PCOS.(66)

 

Clinical Presentation

Women with PCOS commonly present with menstrual disorders (from amenorrhea to dysfunctional uterine bleeding) and infertility, as they have since the syndrome was first described. The compilation of presenting symptoms by Goldzieher et al. from the 1960’s is still relevant today (Figure 4) (67), although obesity, as noted previously, is now much more prevalent in the U.S. population.   Both due to the emphasis on menstrual history and the complaint of androgen excess (rare in children), PCOS classically presents at or after menarche.   The phenotype in pubertal and pre-pubertal girls is debated; there is some evidence to suggest that premature pubarche places girls at increased risk for developing PCOS as they go through puberty. Premature pubarche presents in girls with hyperinsulinemia and elevated DHEAS levels. However, this can only account for a small fraction of women with PCOS, as the prevalence of premature pubarche is a small fraction of PCOS. A national registry of all children in Denmark estimated the prevalence of premature pubarche in the Danish population at 22 to 23 cases per 10,000 girls, i.e., 0.0002% (68). At the other end of the reproductive spectrum, both menstrual irregularity (69) and hyperandrogenemia (70) appear to normalize as women with PCOS approach the perimenopause and menopause. Whether these completely normalize is unknown; for instance, mothers of women with PCOS have elevated testosterone levels compared to controls, suggesting that mild elevations may be familial and persist (43).

Figure 4: A classic reference indicating the prevalence of various presenting clinical symptoms and complaints among a large cohort of women with PCOS ( N = 1089) culled from 187 previously published papers (67). The frequency is still relevant to today’s population of women with PCOS.

Figure 4: A classic reference indicating the prevalence of various presenting clinical symptoms and complaints among a large cohort of women with PCOS ( N = 1089) culled from 187 previously published papers (67). The frequency is still relevant to today’s population of women with PCOS.

Skin disorders, especially those due to peripheral androgen excess, such as hirsutism and acne, and to a lesser degree androgenic alopecia, are common in women with PCOS and frequently the presenting complaint.  Obesity is frequently characterized by a centripetal distribution. This can be diagnosed by an elevated waist circumference (> 88 cm).   A history of weight gain may sometimes precede the onset of oligomenorrhea and hirsutism, leading to the suspicion that this is an acquired form of PCOS secondary to obesity.   All women with PCOS should have a BMI determined at baseline and at regular visits.   Other complaints which must be elicited are screens for mood disorders and depression, as many women with PCOS suffer from low self-esteem due to obesity, disfiguring hirsutism, and infertility.

 

Clinical Sequelae of PCOS:

Although the endocrine and reproductive features of the disorder may improve with age, the associated metabolic abnormalities, particularly glucose intolerance, may actually worsen with age. We shall now discuss common sequelae of PCOS including infertility due to ovulatory dysfunction, abnormalities of the pilosebaceous unit, certain gynecological cancers, type 2 and gestational DM, and cardiovascular disease (CVD).

 

Infertility due to Chronic Anovulation: Women with PCOS are not generally sterile, but subfertile due to the infrequency and unpredictability of their ovulations. Some women with PCOS might tend to conceive later in life as ovulatory function improves (71), although many women now seek treatment earlier in their reproductive lives.   As a rule, women with PCOS represent one of the most difficult groups in whom to induce ovulation both successfully and safely. Many women with PCOS are unresponsive or resistant to ovulation induction with clomiphene citrate. They may have an inappropriate or exaggerated response to the administration of human menopausal gonadotropins (menotropins) and are at increased risk for ovarian hyperstimulation syndrome (OHSS). OHSS is a syndrome of massive enlargement of the ovaries, development of rapid and symptomatic ascites, intravascular contraction, hypercoagulability, and systemic organ dysfunction. It can be life threatening and is best prevented. Increasing obesity may blunt the risk for developing the syndrome (72). These complications occur generally following treatment with menotropins, although ovarian hyperstimulation has even been reported in women with PCOS conceiving a singleton pregnancy spontaneously, or after clomiphene or pulsatile GnRH use (73).

 

In addition to anovulation, endometrial pathology such as hyperplasia may lead to implantation failure in women with PCOS.(74) Induced menstrual bleeding prior to ovulation induction may be associated with lower rates of subsequent pregnancy.(75)

 

Skin Disorders: Skin disorders in women with PCOS revolve primarily around abnormalities of the pilosebaceous unit. The development of hirsutism, acne or androgenic alopecia in PCOS has been attributed to the increased systemic and local production of androgens (see above) that activate abnormal development of the pilosebaceous unit. While insulin is essential for hair follicle growth in vitro, it is unclear whether the hyperinsulinemia of PCOS directly stimulates fine vellus hair to transform into thick, dark terminal hair with the development of hirsutism (76). Generally the ontogeny of abnormalities in the pilosebaceous unit tends to proceed from acne in the peri-pubertal period to hirsutism as a young adult to androgenic alopecia in the mature adult.   Androgenic alopecia is luckily rare among women with PCOS and its etiology also is complex (77).   Apparently similar factors which stimulate terminal midline hair in lower body regions also lead to follicular atresia in the scalp. For the most part, treatments for hirsutism are also relevant for androgenic alopecia (i.e., androgen suppression and androgen antagonism).   However, local vasodilators administered in a crème or lotion, i.e., minoxidil, have been shown to be effective for both male and female androgenic alopecia, whereas they have no known benefit on hirsutism.

 

Other skin disorders that are common include acanthosis nigricans and an increased frequency of skin tags. Acanthosis nigricans is a dermatologic condition marked by velvety, mossy, verrucous, hyperpigmented skin. It has been noted on the back of the neck, in the axillae, underneath the breasts, and even on the vulva (Figure 5). The presence of acanthosis nigricans appears to be more a sign of insulin resistance than a distinct disease unto itself.

Figure 5: Acanthosis nigricans on the nape of the neck in a woman with PCOS.

Figure 5: Acanthosis nigricans on the nape of the neck in a woman with PCOS.

Gynecological Cancers: Many gynecological cancers have been reported to be more common in women with PCOS including ovarian, breast, and endometrial carcinomas. However, the best case of an association between PCOS and cancer can be made for endometrial cancer, as many risk factors for this cancer are present in the PCOS patient, and the epidemiological evidence of an increased incidence in this group of women is growing stronger, with an approximate three-fold increased risk (78,79). In an analysis of 176 patients with endometrial cancer, hirsutism, increased body mass index (BMI) and hypertension were significantly more common in all patients, and nulliparity and infertility significantly were more common among younger patients compared to controls (80,81).

 

Sleep Apnea   Multiple groups have documented an increased risk for sleep apnea and other sleep disorders, such as sleep disordered breathing in women with PCOS (82,83).   This is notable as sleep apnea is relatively uncommon in women, especially premenopausal women (Figure 6). Increased risk for these disorders in women with PCOS has been associated with both hyperandrogenism and insulin resistance PCOS (82,83). Poor sleep may contribute to a vicious metabolic cycle of worsening insulin resistance and glucose tolerance in women with PCOS.(84) It is perhaps too early to recommend universal screening in obese women with PCOS, but it should be considered in women undergoing bariatric surgery, as it is a predictor of morbidity and mortality in patients undergoing bypass surgery (85).   Women with sleep disorders often complain of daytime sleepiness and fatigue after sleeping and may snore. Interestingly the traditional treatment for sleep apnea, i.e., continuous positive airway pressure (CPAP), has been found to improve insulin sensitivity, decrease sympathetic output, and reduce diastolic blood pressure in women with PCOS and sleep apnea.(86)

Figure 6: Prevalence of sleep apnea and other sleep disorders in a cohort of women with PCOS and an unselected control group of women. Women with PCOS had an OR of sleep apnea of 29 (95% CI 5-294) compared to this control group (82).

Figure 6: Prevalence of sleep apnea and other sleep disorders in a cohort of women with PCOS and an unselected control group of women. Women with PCOS had an OR of sleep apnea of 29 (95% CI 5-294) compared to this control group (82).

Non-alchoholic fatty liver disease (NAFLD). This disorder is fatty infiltration of the liver not due to alcohol abuse that is related to insulin resistance. Affected patients may have no symptoms or have mild, nonspecific symptoms such as fatigue or malaise. It is usually accompanied by elevated serum liver function tests, most commonly transaminases. Liver ultrasound may show steatosis, but liver biopsy remains the gold standard for diagnosis and shows evidence of inflammation and fibrosis. It may respond to weight loss and insulin sensitizing therapy. The prevalence of the disorder among women with PCOS is debated.   A recent meta-analysis reported a nearly four-fold higher incidence of NAFLD among women with PCOS compared to controls.(87)   While some reports have noted an increased prevalence, a recent multi-center trial that screened over 1,000 women with PCOS found that only a small fraction (~5%) had elevated liver transaminases (24). This prevalence is comparable to that found in the U.S. population in the NHANES survey.   Routine screening is probably unnecessary at this time and is not recommended by practice guidelines.(11)

 

Type 2 Diabetes Mellitus. The inherent insulin resistance present in many with PCOS, aggravated by the high prevalence of obesity in these individuals, places these women at increased risk for impaired glucose tolerance and type 2 DM. About 30% to 40% of obese reproductive-aged PCOS women have been found to have impaired glucose tolerance (IGT), and about 10% have frank type 2 DM based on a 2-hour glucose level > 200mg/dL (72)(88,89). Of note is that only a small fraction of women with PCOS and with either IGT or type 2 DM display fasting hyperglycemia consistent with diabetes as defined by the American Diabetes Association criteria (fasting glucose ³ 126 mg/dL) (Figure 7). In other PCOS populations with lower rates of obesity, the prevalence of impaired glucose tolerance is also lower, although higher than in control groups (90). The risk factors associated with glucose intolerance in women with PCOS—age, high body mass index (BMI), high waist–hip ratios, and family history of diabetes—are identical to those in other populations. The conversion rate to glucose intolerance varies depending on the population studied (91,92). However, because glucose tolerance tends to worsen with age, periodic rescreening every 3-5 years is recommended in patients with normal glucose tolerance. However, the level of insulin resistance found in women with PCOS based on dynamic measures of insulin action is comparable to that found in other populations (i.e., children of parents with diabetes) associated with a marked increased risk of developing type 2 DM. Recently there has been interest in substituting screening for dysglycemia with measurement of glycohemoglobin (HgbA1c) level rather than oral glucose challenge. However, HgbA1c will tend to miss most of the women with impaired glucose tolerance (close to three quarters(93)), and this is not recommended as it will miss most of the patients who may benefit from intervention to arrest or slow the progression to diabetes. Routine oral glucose tolerance screening has been recommended in populations such as the Chinese with lower risk factor profiles than the U.S. population.(94)

Figure 7: Distribution of glucose tolerance (NGT= normal glucose tolerance or 2h glucose < 140 mg/dL, IGT = impaired glucose tolerance or 2h glucose 140-199 mg/dL, Type 2 DM = 2h glucose ≥ 200 mg/dL) by fasting glucose level in a large cohort (N = 254) women with PCOS. The vertical lines at 110 mg/dL and 126 mg/dL on the fasting glucose x axis indicate the thresholds for impaired fasting glucose and type 2 diabetes by fasting levels (89).

Figure 7: Distribution of glucose tolerance (NGT= normal glucose tolerance or 2h glucose < 140 mg/dL, IGT = impaired glucose tolerance or 2h glucose 140-199 mg/dL, Type 2 DM = 2h glucose ≥ 200 mg/dL) by fasting glucose level in a large cohort (N = 254) women with PCOS. The vertical lines at 110 mg/dL and 126 mg/dL on the fasting glucose x axis indicate the thresholds for impaired fasting glucose and type 2 diabetes by fasting levels (89).

Cardiovascular Disease. Many of the studies suggesting an increased incidence of CVD are inferential based on risk factor models, with little evidence of increased or premature onset of CVD events such as stroke or myocardial infarction (95). There is a lack of prospective studies showing increased risk of cardiovascular events in women with PCOS. However, several cohort studies, including the Nurse’s Health Study, have suggested an increased risk of CVD disease or events in the presence of increasing oligomenorrhea. In this study, there was no determination of hyperandrogenism, so many of the cases may have had another menstrual disorder (96). This is a key confounder as women with primary ovarian insufficiency and prolonged estrogen deficiency likely have higher premature CVD morbidity and all-cause mortality. In other older populations, a history of irregular menses and/or hyperandrogenism has been associated with increased CV events, though it must again be acknowledged there are no accepted criteria for diagnosing PCOS in the menopause.(97) Among postmenopausal women evaluated for suspected ischemia in the Women's Ischemia Syndrome Evaluation (WISE) study, clinical features of PCOS defined by a premenopausal history of irregular menses and current biochemical evidence of hyperandrogenemia were initially associated with more angiographic CAD and worsening CV event-free survival(98); however, this article was retracted and a subsequent re-analysis by the same group showed not only similar CVD morbidity but also overall equal mortality with long term follow up compared to controls.(99) Probably the best evidence for an increased onset of premature cardiovascular events comes from ICD-10 billing-based identification of PCOS in younger women with PCOS; this code has been associated with increased hospitalization rates (3-4 fold higher) for ischemic heart disease and cerebrovascular disease.(100)

Figure 8: Longer-term mortality from the Women's Ischemia Syndrome Evaluation (WISE) study: by PCOS total N = 295, including 25 (8%) have clinical features of PCOS as defined, where 7 (28%) of the women with clinical features of PCOS died compared to 73 (27%) of the 270 without clinical features PCOS died (99).

Figure 8: Longer-term mortality from the Women's Ischemia Syndrome Evaluation (WISE) study: by PCOS total N = 295, including 25 (8%) have clinical features of PCOS as defined, where 7 (28%) of the women with clinical features of PCOS died compared to 73 (27%) of the 270 without clinical features PCOS died (99).

The data are less robust in cohorts of women with better-characterized PCOS. Studies examining subclinical atherosclerosis in premenopausal women with PCOS have detected an increased prevalence compared to controls (ranging in women with PCOS from less than 10% with increased carotid intimal medial thickness (101) to 40% with coronary artery calcification (102,103)).   Another newer marker of cardiovascular disease, cholesterol efflux, has also been noted to be elevated in women with PCOS.(104)

 

Many women with PCOS appear to form a subset of the metabolic syndrome first described by Reaven (i.e., Syndrome X or insulin resistance syndrome) consisting of insulin resistance, hypertension, dyslipidemia, glucose intolerance, and CVD (105). In fact, many women with PCOS have significant dyslipidemia, with lower HDL and higher triglyceride and LDL levels than age, sex, and weight-matched controls (106,107). The elevation in LDL levels is somewhat atypical for the insulin resistance syndrome. Women with PCOS, at least in later life, also appear to have a higher risk of developing hypertension (108,109). Metabolic syndrome appears very common among women with PCOS and in a report from the baseline cohort recruited to one large multi-center trial (including subjects with type 2 diabetes), the prevalence was 33.4% (59).   The most common finding was a waist circumference greater than 88 cm in 80% followed by an abnormal high-density lipoprotein cholesterol of less than 50 mg/dl (Figure 9).  Conversely what most protected against the metabolic syndrome was a normal waist circumference.

Figure 9: Prevalence of components of the metabolic syndrome among a large cohort of women with PCOS. HDL = high-density lipoprotein cholesterol less than 50 mg/dl; TTG= triglycerides greater than or equal to 150 mg/dl; HTN = blood pressure greater than or equal to 130/85 mm Hg; IFG = fasting glucose concentrations greater than or equal to 110 mg/dl (impaired fasting glucose).

Figure 9: Prevalence of components of the metabolic syndrome among a large cohort of women with PCOS. HDL = high-density lipoprotein cholesterol less than 50 mg/dl; TTG= triglycerides greater than or equal to 150 mg/dl; HTN = blood pressure greater than or equal to 130/85 mm Hg; IFG = fasting glucose concentrations greater than or equal to 110 mg/dl (impaired fasting glucose).

Mood Disorders. Women with PCOS appear to be at increased risk for diminished quality of life and mood disorders (110). Specifically they suffer from increased rates of anxiety(111) and depression(112) compared to other women. The magnitude is significant as noted in one study in which women with PCOS were at an increased risk for depressive disorders (new cases) compared with controls (21% vs. 3%; odds ratio 5.11 [95% confidence interval (CI) 1.26-20.69]; P<0.03) (113) A validated quality of life (QoL) questionnaire has been developed for women with PCOS (PCOSQ) (114). Recently a large controlled study of over 100 women (N = 1359) found a high prevalence of low quality of life in women with PCOS (110). Women with PCOS had lower quality of life on all seven factors of the modified PCOSQ (emotional disturbance, weight, infertility, acne, menstrual symptoms, menstrual predictability and hirsutism). Weight was the largest contributor to poor quality of life for women on and off medication for their PCOS. Clinical studies to date have incorporated QoL measures into the trial design. A recent substudy of a clinical trial examining OCP and weight loss and the combination of the two in women found both weight loss and OCP use result in significant improvements in quality of life, depressive symptoms, and anxiety disorders with possible added benefit to the combined therapies.(115)

Differential Diagnosis of PCOS

The differential diagnosis of PCOS includes other causes of androgen excess (Table 1), and PCOS remains a diagnosis of exclusion. Because the work up for many of these disorders is expensive and tests have varying degrees of sensitivity and specificity, some clinical acumen must be applied in the selection of tests. Generally, every woman with signs and symptoms of PCOS should be screened for thyroid dysfunction, prolactin excess, and non-classical congenital adrenal hyperplasia. These diagnoses occur relatively more commonly among women with menstrual disorders, and there are good screening tests to diagnose them. Both hyper- and hypothyroidism have been associated with menstrual disturbances, although their link with hyperandrogenism is less proven. Mild elevations in prolactin are common in women with PCOS (116). A prolactin level can identify prolactinomas that secrete large amounts of prolactin which may stimulate ovarian androgen production, but this is an extremely rare cause of hyperandrogenic chronic anovulation. Evaluating serum levels of thyroid-stimulating hormone is also useful, given the protean manifestations and frequency of thyroid disease in women with menstrual disorders.

 

Non-classical congenital adrenal hyperplasia, often referred to as late-onset congenital adrenal hyperplasia, can present in adult women with anovulation and hirsutism and is due almost exclusively to genetic defects in the steroidogenic enzyme, 21 hydroxylase (CYP21). In Europe and the U.S., congenital adrenal hyperplasia occurs with the highest frequency among Ashkenazi Jews, followed by Hispanics, Yugoslavs, Native American Inuit in Alaska, and Italians (117). Increasingly mandatory postnatal genetic screening is diagnosing this in U.S. born infants. To screen for non-classical congenital adrenal hyperplasia due to CYP21 mutations, a fasting level of 17-hydroxyprogesterone should be obtained in the morning. A value less than 2 ng/mL is considered normal. If the sample is obtained in the morning and during the follicular phase, some investigators have proposed cutoffs as high as 4 ng/mL (118). Specificity decreases if the sample is obtained in the luteal phase due to increased progesterone production. High levels of 17-hydroxyprogesterone should prompt an adrenocorticotropic hormone (ACTH) stimulation test to confirm the diagnosis.

 

As Cushing syndrome is extremely rare (1 in 1,000,000) and screening tests are not 100% sensitive or specific (119), routine screening for Cushing syndrome in all women with hyperandrogenic chronic anovulation is not indicated. Those who have coexisting signs of Cushing syndrome, including a moon facies, buffalo hump, abdominal striae, centripetal fat distribution, or hypertension, should be screened. Proximal myopathies and easy bruising, not typically present in women with PCOS, may also help identify patients with Cushing syndrome.

 

Androgen-secreting tumors of the ovary or adrenal gland are invariably accompanied by elevated circulating androgen levels. However, there is no absolute level that is pathognomonic for a tumor, just as there is no minimum androgen level that excludes a tumor. In the past, testosterone levels above 2 ng/mL and dehydroepiandrosterone sulfate (DHEAS) levels greater than 700 µg/dL were regarded as suspicious for a tumor of, respectively, ovarian and adrenal etiology, but these cutoff levels have poor sensitivity and specificity (120).

 

Evaluation of women with PCOS

History and physical exam is important in evaluation of women with PCOS (Table 2). The history should focus on the onset (peri-pubertal vs acquired later in life) of oligomenorrhea, the onset and duration of the various signs of androgen excess, and concomitant medications, including the use of exogenous androgens. While many medications are associated with hypertrichosis, a generalized increase in body hair, few are associated with increased midline androgen-dependent terminal hair growth. A family history of diabetes and cardiovascular disease (especially first-degree relatives with premature onset of cardiovascular disease [male < 55 years and female < 65 years]) is important. Additionally, multiple studies have shown that PCOS clusters in families, such that a sister or mother with PCOS likely increases risk for the disorder or stigmata of the disorder in other family members. Lifestyle factors such as smoking, alcohol consumption, diet, and exercise are particularly important in these women.   An astonishingly high number of women with PCOS are either current or past smokers. In one large multi-center trial, 17% were current smokers during the trial and 22% had a history of smoking (24). Further a history of recent smoking cessation may not be reliable when checked against urinary cotinine levels (a metabolite of nicotine).(121) Obviously for both fertility and prevention of cardiovascular disease, cessation should be a primary target of the treatment plan.

 

Table 2: Disorders to Consider in the Differential Diagnosis of PCOS

 

Androgen secreting tumor

Exogenous androgens

Cushing syndrome

Nonclassical congenital adrenal hyperplasia

Acromegaly

Genetic defects in insulin action (Leprechaunism, Rabson Mendenhall syndrome, Lipodystrophy)

HAIR-AN syndrome

Primary hypothalamic amenorrhea

Primary ovarian failure

Thyroid disease

Prolactin disorders

 

 

The physical examination should include evaluation of balding, acne, clitoromegaly, and body hair distribution, as well as pelvic examination to look for ovarian enlargement. The presence and severity of acne should be noted. Signs of insulin resistance such as hypertension, obesity, centripetal fat distribution, and the presence of acanthosis nigricans should be recorded. Other pathologic conditions associated with acanthosis nigricans should be considered, such as insulinoma and malignant disease, especially adenocarcinoma of the stomach.

 

The laboratory examination of patients should include tests at initial presentation to exclude other diagnoses as well as to evaluate circulating androgen (Table 3). The best measurement of circulating androgens to document unexplained androgen excess is a subject of debate, and recent expert consensus panels have recommended standardized testosterone assays and normative values for women and children (122). While mass spectrometry is increasingly becoming the gold standard for measurement of all sex steroids(123), studies have shown that even mass spectrometry has poor precision towards the lower levels seen in normal women.(124) Thus, there remains controversy about how to measure androgens in women.

 

Table 3: Focused History and Physical Exam components for Evaluation for PCOS

 

History

  • Onset and Duration of Oligo-ovulation
  • History of weight gain
  • Family history for PCOS, Diabetes, CVD, Endometrial Cancer, etc
  • Infertility (also screen for male and tubal factors)
  • Smoking and substance abuse

 

 

Physical

  • Blood pressure
  • BMI (weight in kg divided by height in m2)

25–30 = overweight, > 30 = obese

  • Waist circumference to determine body fat distribution

Value > 35 in = abnormal

  • Presence of stigmata of hyperandrogenism/insulin resistance

Acne, hirsutism, androgenic alopecia, skin tags, acanthosis nigricans

 

 

 

Both the adrenal glands and ovaries contribute to the circulating androgen pool in women. The adrenal gland preferentially secretes weak androgens such as dehydroepiandrosterone (DHEA) or DHEAS (up to 90% of adrenal origin). These hormones, in addition to androstenedione, may serve as prohormones for more potent androgens such as testosterone and dihydrotestosterone. The ovary is the preferential source of testosterone, and it is estimated that 75% of circulating testosterone originates from the ovary (mainly through peripheral conversion of prohormones by liver, fat, and skin, but also through direct secretion). Androstenedione, largely of ovarian origin, is the only circulating androgen that is higher in premenopausal women than in men, yet its androgenic potency is only 10% of testosterone. Dihydrotestosterone is the most potent androgen, although it circulates in negligible quantities and results primarily from the intracellular 5a-reduction of testosterone.

 

Many studies attempting to identify the best circulating androgen for differentiating women with PCOS from control women have usually identified testosterone, androstenedione or both. (66,125) Each clinician should be familiar with the analytical performance and the normal ranges of local laboratories, as there is no standardized testosterone assay (and no accepted testosterone standard) in the U.S. and the sensitivity and reliability in the female ranges are often poor (122). Evaluation of DHEAS levels may be useful in cases of rapid virilization (as a marker of adrenal origin), but its utility in assessing common hirsutism is questionable.

 

The Rotterdam criteria have led to increasing use of ultrasound in the initial diagnosis and evaluation of women with PCOS (Table 4). In addition to ovarian size, ultrasound can exclude leiomyomas and most mullerian anomalies and the determine the thickness of the endometrium. Some studies have found very high asymptomatic rates of endometrial hyperplasia among amenorrheic women with PCOS (126). However. routine screening with ultrasound of the endometrium or routine endometrial biopsy is not recommended in the absence of abnormal uterine bleeding.

 

Table 4:     Suggested Laboratory and Radiologic Examination of women with PCOS

 

 

Laboratory

  • Documentation of biochemical hyperandrogenemia

Total testosterone and SHBGor bioavailable/free testosterone

  • Exclusion of other causes of hyperandrogenism

Thyroid-stimulating hormone levels (thyroid dysfunction)

Prolactin (hyperprolactinemia)

17-hydroxyprogesterone (nonclassical congenital adrenal hyperplasia due to 21 hydroxylase deficiency)

Random normal level < 4 ng/mL or morning fasting level < 2 ng/mL

Consider screening for Cushing syndrome and other rare disorders such as acromegaly

  • Evaluation for metabolic abnormalities

2-hour oral glucose tolerance test (fasting glucose < 110 mg/dL = normal, 110–125 mg/dL = impaired, >126 mg/dL = type 2 diabetes) followed by 75-g oral glucose ingestion and then 2-hour glucose level (< 140 mg/dL = normal glucose tolerance, 140–199 mg/dL = impaired glucose tolerance, >200 mg/dL = type 2 diabetes)

  • Fasting lipid and lipoprotein level (total cholesterol, HDL < 50 mg/dL abnormal, triglycerides > 150 mg/dL abnormal

 

Ultrasound Examination

  • Determination of polycystic ovaries
  • Identify endometrial abnormalities

 

Optional Tests to Consider

  • Gonadotropin determinations to determine cause of amenorrhea
  • Fasting insulin levels in younger women, those with severe stigmata of insulin resistance and hyperandrogenism
  • 24-hour urine test for urinary free cortisol with late onset of PCOS symptoms or stigmata of Cushing syndrome

 

 

The metabolic evaluation of women with PCOS has become a standard part of the evaluation. Exclusion of diabetes and identification of glucose intolerance can be obtained with a standard 75g oral glucose tolerance test. At the same time a fasting lipid profile can be obtained. The routine use of insulin levels in the diagnosis and management of women with PCOS is probably not indicated, as they are poor markers of insulin resistance if there is beta cell dysfunction and they have not been found to predict response to therapy. The identification of the metabolic syndrome is a better clinical marker of insulin resistance.

 

Approach to TREATMENT OF WOMEN WITH PCOS

Treatment of women with PCOS tends to be symptom based, as there are few therapies which address the multitude of complaints with which women with PCOS present.   Arguably there are currently only two therapies that address the most common complaints, (i.e., infertility, hirsutism, menstrual disorders, and obesity) and these are either weight loss (as a result of lifestyle modification, medical or surgical therapy to reduce weight, or metformin therapy) (Table 5).   It is often difficult to treat all complaints at once, with the greatest difficulty in treating both anovulatory infertility and hirsutism concurrently.(127) Some therapies can also be counterproductive and thus contraindicated in this situation, for instance the use of oral contraceptives because they block ovulation or the use of anti-androgens because they are potentially teratogenic in a male fetus.   Because of these conundrums in clinical care, treatment tends to fall into two categories: either the treatment of anovulatory infertility or long-term maintenance treatment for PCOS-related symptoms (i.e., hirsutism, menstrual disorders, obesity, etc.)

 

Table 5:   Commonly used or proposed treatments for PCOS or stigmata of PCOS. Many of these are used off label.

 

Overview of Treatment of Anovulatory Infertility. One important consideration before treating subjects with anovulatory infertility is to screen the couple for other infertility factors. One large multi-center trial found that 10% of male partners of women with PCOS had co-existing severe oligospermia and close to 5% of women had bilateral occlusion of the fallopian tubes or some uterine factor (128). Obviously, the presence of these factors would significantly alter therapy, and their high prevalence justifies pre-treatment screening. There is no evidence-based schema to guide the initial and subsequent choices of approaches to ovulation induction in women with PCOS. The ASRM/ESHRE sponsored conference recommended that before any intervention is initiated, preconceptual counseling should emphasize the importance of lifestyle, especially weight reduction and exercise in overweight women, smoking cessation, and reducing alcohol consumption (129,130).

 

The recommended ASRM/ESHRE first-line treatment for ovulation induction remains the anti-estrogen clomiphene citrate (CC), and this view has been upheld by other groups including the World Health Organization(131). However there is now increasing evidence that letrozole, an aromatase inhibitor, is more efficacious and equally safe to mother and fetus(132). Recommended second-line intervention, should be CC or the combination of metformin and CC if first line therapy fails to result in pregnancy. Third-line therapy is either exogenous gonadotropins or laparoscopic ovarian surgery (129,130). The caregiver must carefully assess the reproductive toxicity of all medications used in women with PCOS, because several may increase ovulatory frequency and result in unexpected and unintended pregnancy and possible fetal exposure. Recently the FDA eliminated the categorization of teratogenicity of medications into categories (i.e., Category A, B, C, D and X) and instead ruled that package inserts should provide specific data about teratogenic risks in animals and humans or acknowledge the lack of such data. Eventually all package inserts will be modified to reflect true risk as opposed to theoretical risk based on mechanism of action of the drug.

 

Overview of Long Term Maintenance of PCOS. There is no known cure for PCOS; rather therapy revolves around suppression of symptoms.   Therapy tends to focus on the primary chief complaint. However often the triad of hirsutism, oligomenorrhea, and obesity forms the key presenting symptoms. In such cases, it may make sense to choose a primary metabolic parameter upon which to base initial treatment.   Glucose intolerance is the strongest risk factor for diabetes and is also an independent risk factor for cardiovascular events in women(133) and is one potential factor to use in selecting initial treatment.   A possible first-line strategy is found in Figure 11, which allows selection of the therapies that improve the triad of PCOS symptoms.     Additional targeted therapies for hirsutism and/or oligomenorrhea could be added depending on response to the initial therapy. Obviously, contraception should be considered if the patient is trying to avoid pregnancy.

Figure 10: Suggested first-line treatment plan for infertile women with PCOS.

Figure 10: Suggested first-line treatment plan for infertile women with PCOS.

Figure 11: Suggested first line treatment plan for women with PCOS not seeking pregnancy.

Figure 11: Suggested first line treatment plan for women with PCOS not seeking pregnancy.

Review of Efficacy of Individual Therapies on PCOS

 

Aromatase Inhibitors   Aromatase inhibitors, specifically letrozole, may be a first-line therapy for ovulation induction in women with PCOS. While the mechanism of action of aromatase inhibitors is likely similar to clomiphene in that the target is the hypothalamic pituitary axis and the normalization of gonadotropin secretion, the site of action may be multifocal, i.e., in the hypothalamus, in peripheral fat tissues, and perhaps even in the ovary. The proposed benefits of letrozole include oral administration, a shorter half-life than clomiphene, more favorable effects on the endometrium, potentially higher implantation rates, and lower multiple pregnancy rates due to monofollicular ovulation.(134) A large multicenter study conducted by the Reproductive Medicine Network of letrozole versus clomiphene upheld many of these hypotheses and showed a 44% improvement in the live birth rate with letrozole over clomiphene.(25) The greatest benefit was in the moderately obese group, though a tertile analysis trended towards benefit of letrozole over clomiphene in all weight classes (Figure 12).   Subsequent studies have replicated these results and the meta-analysis suggests a 50-60% live birth benefit with letrozole over clomiphene.(132) Although the trend in the Reproductive Medicine Network study was towards a lower multiple pregnancy rate with letrozole versus clomiphene (3.9% vs 6.8%), even larger studies will be necessary to confirm this trend.(135) Letrozole offers a higher per cycle and cumulative ovulation rate than clomiphene. Only 10% of women failed to ovulate at least once in the Reproductive Medicine Network study compared to close to 25% of women on clomiphene. There is also better fecundity per ovulated patient, suggesting a better quality of ovulation. When the results of a midluteal ovulation check (by both ultrasound and serum assays) are compared to baseline in the follicular phase, women with PCOS on letrozole have higher progesterone levels and lower estradiol levels, thus with a more physiologic hormonal milieu than with clomiphene. Women also have a relatively thinner endometrium on letrozole (against expectation) and lower antral follicle counts and AMH levels relative to clomiphene. Again, this suggests normalization of endometrial response and ovarian morphology with letrozole.

Figure 12: Tertile Analysis by BMI group of women with PCOS of live birth rate over time randomized to clomiphene or letrozole.

Figure 12: Tertile Analysis by BMI group of women with PCOS of live birth rate over time randomized to clomiphene or letrozole.

Safety has been closely studied in randomized trials of letrozole for ovulation induction in women with PCOS. Relative to clomiphene, letrozole is associated with significantly more fatigue and dizziness, but fewer episodes of hot flashes.(25) There is no increased incidence of serious adverse events with letrozole, and no clear pattern or relationship to the drug. Paramount to the use of letrozole is the concern about teratogenicity with letrozole relative to clomiphene. In two large studies using letrozole conducted by the Reproductive Medicine Network, the anomaly rates were comparable between clomiphene and letrozole.(25,136) In both studies they were under 5% with both drugs and within expected rates, especially when acknowledging that subfertile women have higher rates of fetal anomalies than women who conceive without assistance. Further there was no pattern to the reported anomalies, suggesting that a specific organ or organ system was altered by letrozole exposure. Case series have also supported the relative fetal safety of letrozole compared to clomiphene.(137-139) Finally such studies must consider that the underlying rate of congenital anomalies is higher among women with a history of subfertility or who have conceived through fertility treatments.(140) Although letrozole is not recommended as a fertility treatment agent in certain countries due to black box warnings, the source of this concern is unwarranted without supporting published data.

 

There are still many unanswered questions about letrozole including its use as an adjuvant agent with other medications used to treat PCOS, whether it is effective as a second-line solo treatment after clomiphene, and whether prolonged dosing would increase pregnancy rates. Of note is that anastrozole failed both as a high dose one-time administration to women with PCOS(141) and also as a lower dose multi-day therapy compared to clomiphene,(142) so all aromatase inhibitors are not alike.

 

Clomiphene Citrate. Clomiphene citrate (CC) has traditionally been the first-line treatment agent for anovulatory women, including those with PCOS, and several multi-center randomized controlled trials have upheld the use of clomiphene as first-line treatment. In fact, this may the area of study of PCOS with the largest and best designed studies. Clomiphene is a triphenylethylene derived nonsteroidal agent that is theorized to function at the level of the hypothalamus as an anti-estrogen to improve gonadotropin secretion. CC use is associated with hot flashes, mood changes, and rarely changes in vision thought due to pituitary swelling (thought to be a serious event and reason for discontinuing the drug). From a public health perspective, more concerning is the relatively high rate of multiple pregnancy after conception with clomiphene of 7.8%, although the majority are twins (143). However, there is nevertheless a high order (triplets or more) multiple pregnancy rate of 0.9% (143). Comparison of the multiple pregnancy rate after conception with clomiphene suggests that the multiple pregnancy rate may be slightly higher in women with unexplained infertility(136) than in women with PCOS (25), although the lower rate in women with PCOS may also be related to higher obesity rates. Six-month life birth rates range from 20-30%(25) and are higher over longer periods of observation (144). Half of all women who are going to conceive using clomiphene will do so at the 50-mg starting dose, and another 20% will do so at the 100-mg/d dose (145). Most pregnancies will occur within the first six ovulatory cycles, although constant monthly pregnancy rates were noted, suggesting there may be continued benefit to longer use (146).   Prognostic clinical factors for live birth with clomiphene include decreased BMI, less hirsutism, younger age, and shorter duration of attempted conception(147,148).

Alternative clomiphene regimens have been developed, including prolonging the period of administration (149), pretreating with oral contraceptives (150) adding dexamethasone (151), and adding metformin (152). Dexamethasone as adjunctive therapy with clomiphene citrate has been shown to increase ovulation and pregnancy rates in clomiphene-resistant women with PCOS (153). Finally some groups have recommended using similar compounds to clomiphene, such as tamoxifen, in lieu of clomiphene (154).

 

Clomiphene is usually started at 50 mg a day for 5 days and increased by 50 mg a day in subsequent cycles if the patient remains anovulatory up to a maximum daily dose of 150 mg/d. As noted above, induced withdrawal bleeding in the face of continued anovulation may lower subsequent ovulation and pregnancy rates and certainly requires more time and resources. This has led to the adoption of a “stair step” approach where the dose is escalated based on ultrasound and serum determination of follicular development and/or ovulation.(155) The primary discomfiting side effect with clomiphene is hot flashes, likely due to its anti-estrogenic effects in the hypothalamus. Rare side effects that require immediate attention and discontinuation of medication are a sudden change in vision or loss in vision.   There is currently thought to be no added risk of congenital anomalies to women who conceive on the medication as opposed to other therapies.

 

Gonadotropins. Gonadotropins are frequently used to induce ovulation in women with PCOS for whom clomiphene treatment has failed. Low-dose therapy with gonadotropins offers a higher rate of ovulation, monofollicular development, with a significantly lower risk of ovarian hyperstimulation syndrome (156). When given in controlled situations with strict cancellation policies for excessive follicular development, gonadotropins lead to higher pregnancy rates than does clomiphene with similar multiple pregnancy rates.(157) This has led some to recommend this as a first-line therapy, although the expense and higher complication rates in untrained hands remain major treatment considerations. A low-dose regimen is the ASRM/ESHRE (158) and WHO consensus recommendation (131) when using gonadotropins in women with PCOS.

 

In Vitro Fertilization. Women with PCOS who undergo IVF generally have a good prognosis for pregnancy and live birth compared to many other common indications for IVF. Data from the U.S. SART database suggest that women with PCOS have an increased chance for live birth compared to women with tubal disease.(159) Women with PCOS have a higher number of oocytes retrieved than women with tubal factor and live-birth rates were also increased in women with PCOS (34.8% vs. 29.1%; OR, 1.30; 95% CI, 1.24-1.35).(Figure 13)   A similar rate of decline in clinical pregnancy and live-birth rates was noted in both groups with age (20-44 years) and live-birth rates were not significantly different for each year after age 40 in the two groups. Thus, women with PCOS appear to enjoy the greatest benefit in increased live-birth rates over tubal factor between the ages of 30 and 40 years.

Figure 13: Adjusted analysis of outcomes in tubal factor infertility versus PCOS of live-birth rate by continuous age based on the SART age group data.

Figure 13: Adjusted analysis of outcomes in tubal factor infertility versus PCOS of live-birth rate by continuous age based on the SART age group data.

Figure 14: Kaplan Meier Curves of cumulative pregnancy rates in the six-month double-blind randomized trial of clomiphene, metformin, or the combination of both in treatment of anovulatory infertility in PCOS (Pregnancy in Polycystic Ovary Syndrome Study-PPCOS).

Figure 14: Kaplan Meier Curves of cumulative pregnancy rates in the six-month double-blind randomized trial of clomiphene, metformin, or the combination of both in treatment of anovulatory infertility in PCOS (Pregnancy in Polycystic Ovary Syndrome Study-PPCOS).

Women with PCOS are at increased risk for ovarian hyperstimulation syndrome as noted above. Recently a large multi-center trial from China examined the risk-benefit ratio of elective freezing of all embryos followed by frozen embryo transfer versus fresh embryo transfer in women with PCOS.(160) Frozen-embryo transfer resulted in a higher frequency of live birth after the first transfer than did fresh-embryo transfer (49.3% vs. 42.0%), for a rate ratio of 1.17 (95% confidence interval [CI], 1.05 to 1.31; P=0.004). This appeared to be largely mediated through reduced pregnancy loss after frozen-embryo transfer (22.0% vs. 32.7% in the fresh group), for a rate ratio of 0.67 (95% CI, 0.54 to 0.83; P<0.001), and of the ovarian hyperstimulation syndrome (1.3% vs. 7.1%), for a rate ratio of 0.19 (95% CI, 0.10 to 0.37; P<0.001), but a higher frequency of preeclampsia (4.4% vs. 1.4%), for a rate ratio of 3.12 (95% CI, 1.26 to 7.73; P=0.009). There were five neonatal deaths in the frozen-embryo group and none in the fresh-embryo group (P=0.06). These data suggest a mixed risk-benefit ratio of elective frozen embryo transfer in women with PCOS.

 

Ovarian Surgery. The value of laparoscopic ovarian drilling with laser or diathermy as a primary treatment for subfertile women with anovulation and PCOS is undetermined (161), and it is primarily recommended as second-line infertility therapy. Neither drilling by laser or diathermy has any obvious advantage, and there is insufficient evidence to suggest a difference in ovulation or pregnancy rates when drilling is compared with gonadotropin therapy as a secondary treatment (161). Multiple pregnancy rates are reduced in those women who conceive following laparoscopic drilling. In some cases, the fertility benefits of ovarian drilling may be temporary and adjuvant therapy after drilling with clomiphene may be necessary (162). Ovarian drilling does not appear to improve metabolic abnormalities in women with PCOS (163). Long term follow up of a Dutch cohort who underwent either laparoscopic drilling or gonadotropin therapy showed higher fecundity rates with laparoscopic drilling.(164) and greater cost effectiveness of the surgery.(165)

 

Ovarian drilling may also be used to restore menstrual cyclicity in women not seeking pregnancy, and there is evidence in some series of long term improvement in menses as a result of surgery (166). However, these series are uncontrolled and as noted above hyperandrogenism and oligomenorrhea tend to improve with age in women with PCOS regardless of any treatment.

 

Metformin. The use of metformin as first-line solo infertility therapy has not been supported by randomized trials, although there are emerging data about its utility as an adjuvant agent. In the largest trial to date, clomiphene was roughly three times more effective at achieving live birth compared to metformin alone (Figure 12)(146).   Meta-analysis of metformin studies in women with PCOS has not found a benefit in terms of live birth, but clinical pregnancy rates were improved for metformin versus placebo (pooled OR 2.31, 95% CI 1.52 to 3.51, 8 trials, 707 women).(167) Metformin combined with clomiphene may be the best combination in obese women with PCOS as noted in several randomized trials. (146,168)

 

Metformin has no known human teratogenic risk or embryonic lethality in humans and appears safe in pregnancy.  There is no solid evidence that metformin use early in pregnancy prevents pregnancy loss (169), and the randomized trials which stopped drug with pregnancy have shown similar miscarriage rates with metformin as with clomiphene (146,170). Similarly, the use of metformin throughout pregnancy in women with PCOS has not been associated with clear benefit beyond blunting gestational weight gain. (171) Surprisingly in this large multi-center trial, there was no prevention of gestational diabetes. In other populations, metformin has been found to have similar effects as insulin for the treatment of gestational diabetes yet is better tolerated by patients (172) and does not result in change in birth weights when given to obese women who are pregnant(173).

 

Metformin may be most useful in the long-term maintenance of PCOS. Metformin does lower serum androgen, increases ovulations, and improves menstrual frequency (174). While menstrual frequency is improved by roughly a third to a half from baseline, metformin does not always restore regular menstrual cycles in women with PCOS. There may also be favorable effects in preventing the progression to diabetes. The Diabetes Prevention Program demonstrated that metformin can prevent the development of diabetes in high-risk populations (e.g., those with impaired glucose tolerance) (175), and this result has been replicated for a number of anti-diabetic drugs in individuals at high risk. Metformin tends to be the drug of choice to treat glucose intolerance and elevated diabetes risk in women with PCOS because of its favorable safety profile and the familiarity a wide number of caregivers from varying specialties have with the medication. However, there are no adequately powered long term studies of metformin in women with PCOS to document diabetes prevention.   Among women with PCOS who use metformin, glucose tolerance improves or stays steady over time (176). Metformin also may be associated with weight loss in women with PCOS, although the results in other populations are inconsistent (146,177). Metformin is often used in conjunction with lifestyle therapy to treat PCOS. Recent studies suggest that there is limited benefit to the addition of metformin above lifestyle therapy alone in PCOS (178-182).

 

Metformin carries a small risk of lactic acidosis, most commonly among women with poorly controlled diabetes and impaired renal function. Gastrointestinal symptoms (diarrhea, nausea, vomiting, abdominal bloating, flatulence, and anorexia) are the most common adverse reactions and may be ameliorated by starting at a small dose and gradually increasing the dose or by using the extended-release version. The dose is usually 1500-2000 mg/day given in divided doses. The effects of metformin and other anti-diabetic drugs on preventing endometrial hyperplasia/neoplasia in women with PCOS are largely unknown.

 

Thiazolidinediones, Smaller trials have shown some benefit to this class of drugs for the treatment of infertility, usually in conjunction with clomiphene (183,184). However, the concern about hepatotoxicity, cardiovascular risk, weight gain, and reproductive toxicity in animal studies have limited the use of these drugs in women with PCOS. One of the thiazolidinediones, troglitazone, was removed from the market due to hepatotoxicity, and there has been increasing scrutiny of rosiglitazone because of increased cardiovascular events and of pioglitazone because of breast cancer. Nonetheless, improving insulin sensitivity with these drugs is associated with a decrease in circulating androgen levels, improved ovulation rate, and improved glucose tolerance (54,185-187).   However, given the restrictions on their use in patients with type 2 diabetes, the risk-benefit ratio appears very unfavorable for women with PCOS.

 

GLP-1 agonists. GLP-1 (Glucagon-Like Peptide) is an incretin secreted by the L cells of the intestine which increases pancreatic beta cell insulin production and insulin sensitivity. It also has CNS effects which lead to decreased appetite through a variety of mechanisms.   GLP-1 agonists have been approved both for the treatment of type 2 diabetes and in higher doses (liraglutide) for the treatment of obesity in the U.S.   These drugs thus offer a favorable dual treatment strategy for women with PCOS. Studies, however, have been limited, likely by two factors, the requirement for parenteral injection and the relative expense of the drugs compared to oral alternatives.   Small observational studies do support that women with PCOS tend to lose weight and experience improvements in metabolic parameters related to insulin resistance. (188,189) Side effects of concern with this class of drugs include pancreatitis and an increased risk of thyroid cancer (medullary) and concerns about CNS interactions in patients with psychiatric disorders.

 

Combination Oral Contraceptives. Oral contraceptives have been the mainstay of long-term management of PCOS among gynecologists even though there are few well designed trials in women with PCOS. They offer benefit through a variety of mechanisms, including suppression of pituitary LH secretion, suppression of ovarian androgen secretion, increased circulating SHBG levels (and thus decreased peripheral androgen exposure) as well as potential antagonism of steroidogenic enzymes or steroid receptors (most commonly the androgen receptor). Estrogen may be the most potent stimulator of SHBG production. Individual OC preparations may have different doses and drug combinations and thus have varying risk–benefit ratios. For instance, various progestins have been shown to have different effects on circulating SHBG levels (190), but whether that translates into any clinical differences among preparations is uncertain. The “best” oral contraceptive for women with PCOS is unknown based on data, only on marketing hype. Oral contraceptives also are associated with a significant reduction in risk for endometrial cancer with a reduction of risk by 56% after four years of use and 67% after eight years in users compared to non-users (191), but the magnitude of the effect in women with PCOS is not known.

 

Because women with PCOS may have multiple risk factors for adverse effects and serious adverse events on oral contraceptives, they must be screened carefully for risk factors for these events including smoking history, presence of obesity and hypertension, and history of clotting diathesis to mention some of the important factors (Table 6). Studies based on insurance claims have suggested women with PCOS may have a higher risk of thromboembolic events on or off of OCP.(192) In the larger U.S. population, oral contraceptive use has not been associated with an increased risk of developing type 2 diabetes (193). There is no convincing evidence that use of oral contraceptives contributes to the risk of diabetes in women with PCOS, although there are often adverse effects on insulin sensitivity that may be dose dependent (194,195). Our own study showed a short-term (16 week) 25% deterioration in glucose tolerance on a low dose OCP compared to baseline in obese women with PCOS.(196) However, longer follow up studies are needed. A low dose oral contraceptive pill is therefore recommended.  Concurrent lifestyle modification in obese women with PCOS may ameliorate the adverse metabolic effects of OCP.(196)

 

Table 6: Absolute and Relative Contraindications to Oral Contraceptive Use. Women with PCOS should be screened for these (common abnormalities in this group of women are underlined) and risk benefit ratios carefully discussed with them before initiating therapy.

Absolute contraindications

< 6 weeks postpartum if breastfeeding

Smoker over the age of 35 (≥ 15 cigarettes per day)

Hypertension (systolic ≥ 160mm Hg or diastolic ≥ 100mm Hg)

Current or past history of venous thromboembolism (VTE)

Ischemic heart disease

History of cerebrovascular accident

Complicated valvular heart disease

Migraine headache with focal neurological symptoms

Breast cancer (current)

Diabetes with retinopathy/nephropathy/neuropathy

Severe cirrhosis

Liver tumour (adenoma or hepatoma)

Relative Contraindications

Smoker over the age of 35 (< 15 cigarettes per day)

Adequately controlled hypertension

Hypertension (systolic 140–159mm Hg,diastolic 90–99mm Hg)

Migraine headache over the age of 35

Currently symptomatic gallbladder disease

Mild cirrhosis

History of combined oral contraceptive related cholestasis

Users of medications that may interfere with combined oral contraceptive metabolism

 

 

 

Oral contraceptives may also be associated with a significant elevation in circulating triglycerides as well as in HDL levels, although these increases do not appear to progress over time (197).   There is no evidence to suggest that women with PCOS experience more cardiovascular events than the general population when they use oral contraceptives, although risk factors for adverse events such as hypertension, obesity, clotting history, and smoking must be considered. The effect of progestins alone on metabolic risk factors varies and is not well understood.

 

No oral contraceptive has been approved by the FDA for the treatment of hirsutism although many have been approved for treatment of acne. A number of observational or nonrandomized studies have noted improvement in hirsutism in women with PCOS who use oral contraceptives (198). Few studies have compared outcomes of different types of oral contraceptives, and no one type of pill has been shown definitively to be superior in treating hirsutism in women with PCOS. The largest randomized study out of India suggested a greater benefit at 12 months in treating hirsutism with a pill containing cyproterone acetate, a pill not available in the U.S., compared to ones containing desogestrel or drospirenone. (199) There was no difference among groups at 6 months. The take home message from this and other studies is that the improvement in hirsutism is slow and steady, and longer time frames are required to document improvement in hirsutism. A number of studies have found additive benefit when oral contraceptives are combined with other treatment modalities, most commonly spironolactone, to treat hirsutism.(200) If a woman is taking an oral contraceptive that contains drospirenone (brand name Yasmin and Yaz), a progestin with anti-mineralocorticoid properties, it may be necessary to reduce her dose of spironolactone if used as additional therapy, and evaluate her levels of potassium. There have been several epidemiologic studies that have linked newer progestins, including drospirenone with an increased risk of thromboembolic events. However, these studies have been criticized for potential prescribing or detection bias.

There is a theoretical benefit to treating hyperandrogenism with extended cycle formulations, as these are less likely to result in rebound ovarian function and more likely to lead to more consistently suppressed ovarian steroid levels, including androgens (201).   However, there have been few studies to uphold this in practice, although these are increasingly utilized for other reasons, such as decreased vaginal bleeding and greater patient satisfaction.

 

Progestins. Both depot and intermittent oral medroxyprogesterone acetate(MP) (10 mg for 10 days) have been shown to suppress pituitary gonadotropins and circulating androgens in women with PCOS (202). Depot MPA has been associated with weight gain, mood changes and breakthrough bleeding, but provides effective contraception if needed.   No studies have addressed the long-term use of these compounds to treat hirsutism. The regimen of cyclic oral progestin therapy that most effectively prevents endometrial cancer in women with PCOS is unknown. There is also a paucity of data to address the varying risk-benefit ratios of varying classes of progestins. Progestin-only oral contraceptives are an alternative for endometrial protection, but they are associated with a high incidence of breakthrough bleeding.

 

Cyproterone acetate is a progestin not available commercially in the U.S. with anti-androgenic properties. It is frequently combined in an oral contraceptive in other countries and is popular in the treatment of PCOS. A newer progestin from the same class, drospirenone, has been marketed in the U.S. as especially effective for the treatment of female hyperandrogenism, although the data suggesting this is superior to other formulations is not based on head-to-head randomized trials (203).

 

Intrauterine Devices. There is increasing literature supporting the benefit of IUDs, especially a progestin (levonorgestrel) containing IUD, in treating a variety of endometrial disorders, including menorrhagia (204,205), simple endometrial hyperplasia (206) and complex hyperplasia(207). A recent meta-analysis, based on a small number of high quality studies, concluded that a levonorgestel containing IUD may be more effective than oral progestins in treating simple endometrial hyperplasia.(208)   Levonorgestrel containing IUDs have also been used to treat hyperplasia with atypia and even some local cases of endometrial adenocarcinoma in women desiring to preserve their uteruses for fertility.(209)

 

Uterine Surgery. In patients with intractable uterine bleeding who have completed their child-bearing, consideration may be given to either endometrial ablation or more definitive surgical therapy via hysterectomy. The long-term risk of endometrial cancer developing in isolated pockets of endometrium after ablation remains a theoretical concern without clear data in this group of women at high risk for endometrial cancer.

 

Statins. Another area where there is limited support in the literature for a cardiovascular and endocrine benefit in women with PCOS is with the use of statin(210). They have been shown to improve hyperandrogenemia, lipid levels, and reduce inflammation in women with PCOS (211,212). They have been studied in conjunction with both OCP and metformin with additive benefits noted.(213,214) However their long term effects in preventing cardiovascular disease in young women with PCOS is unknown, although a small but clinically significant preventive effect on CVD events was noted in a young population without dyslipidemia but with elevated C reactive protein levels.(215) There are theoretical concerns about teratogenicity with the use of this drug in reproductive aged women based primarily on its mechanism as a cholesterol synthesis inhibitor. The use of these drugs is still experimental in women with PCOS, and the comparative effects of varying statins in women with PCOS is unknown.

 

Lifestyle Modification. The gold standard for improving insulin sensitivity in obese PCOS women should be weight loss, diet, and exercise. It is recommended as the first-line of treatment in obese women who present with infertility as discussed above.   Obesity has become epidemic in our society and contributes substantially to reproductive and metabolic abnormalities in PCOS. Unfortunately, there are no effective treatments that result in permanent weight loss, and it is estimated that 90-95% of patients who experience a weight decrease will relapse. In markedly obese individuals, the only treatment that results in sustained and significant weight reduction is bariatric surgery (216). However, only a fraction of eligible patients ever elect or are qualified to receive bariatric surgery. In the U.S., it is estimated that only 1% of eligible patients receive bariatric surgery for the treatment of obesity.

 

There is no miracle diet in women with PCOS despite claims to the contrary. Hypocaloric diets result in appropriate weight loss in women with PCOS (arguing against any special defect towards weight retention). There is no clear evidence that any particular dietary composition benefits weight loss or reproductive or metabolic changes in women with PCOS (217,218), although “subtle differences” between diets were noted in a recent systematic review.(219) A two-year study in the general population found comparable weight loss among three types of diets of varying macronutrient composition and found comparable weight loss and similar improvement in lipid and insulin levels (220).   Thus, the consensus recommendations for women with PCOS is to utilize any type of hypocaloric diet that they can tolerate and maintain (129,130).

 

There have been unfortunately few studies on the effect of exercise alone on symptoms in PCOS women (221), although it is reasonable to assume that exercise would have the same beneficial effects in PCOS women as in women with type 2 DM. These benefits relate more to improved glycemic control and less to weight loss. In fact, exercise alone or in addition to caloric restriction adds only modestly to weight loss, in the range of 2-4% over a sustained period.(222) Exercise is thought to play a key role in weight maintenance after weight loss from caloric restriction.(223) However the exercise program must be tailored to the degree of obesity and the patient’s baseline fitness. Women with PCOS and morbid obesity may be poor candidates for weight bearing aerobic exercise due to musculoskeletal overload. Additionally, there may be medical contraindications to certain form of exercise.

 

Bariatric Surgery. Bariatric surgery is increasingly used in morbidly obese patients as a first-line obesity therapy.   The current National Institutes of Health recommendations are to utilize bariatric surgery in patients with a BMI greater than 40 or with a BMI greater than 35 and serious medical co-morbidities (224).   PCOS has been listed as a co-morbidity justifying bariatric surgery by some experts. Some women with PCOS appear to experience a dramatic improvement in symptoms after surgery (225,226). However, these studies are primarily case series and need further validation in prospective studies. Randomized studies have documented that bariatric surgery is superior to medical treatment in controlling type II diabetes induced hyperglycemia, as well as providing a lower body weight and improved quality of life up to three years after surgery.(227)   Weight loss may result in resumed ovulation and pregnancy during the period of rapid weight loss (first 6-12 months after surgery), which has led to concern about the effects of malnutrition on the fetus and general recommendations to refrain from pregnancy for 12-24 months. Data support that women who conceive after bariatric surgery are at increased risk for small-for-gestational-age babies and shorter pregnancies.(228)

 

The ideal bariatric procedure for PCOS is unknown. Previously, it was thought that gastric banding was ideal, where the gastric band could be adjusted to accommodate larger caloric loads in case of pregnancy.  However, there were long-term complications from band erosion. Roux-en-Y Gastric Bypass (RYGB) was until recently the most commonly performed procedure and results in significantly more weight loss than gastric banding (“lap banding”). However, this procedure is now being overtaken by Vertical Sleeve Gastrectomy (VSG), which has lower operative and long-term morbidity due to the lack of bowel re-anastomosis that characterizes RYGB. VSG offers long term weight loss slightly less than RYGB.(229)

 

Pharmacologic treatment of obesity. Because weight loss generally improves stigmata of PCOS, there have been a number of studies using medications for the treatment of obesity as primary treatments of PCOS. They appear in limited and small trials to offer some benefit (230-234). The current medications approved for the treatment of obesity in the U.S. and their risk/benefit ratio are summarized in Table 7. It is important to note that one medication, sibutramine (a sympathomimetic central appetite suppressant) was removed recently from the U.S. market by the FDA because of concerns of increased CVD events with its use. Similarly, rimonabant (a cannabinoid CB1 receptor antagonist central appetite suppressant) which was approved in Europe (but not in the U.S.) was removed from their market because of concerns about suicidal ideation and suicides on the drug. Currently there are insufficient data about the benefits of these drugs on signs and symptoms of PCOS to recommend them as a treatment for endocrine-related symptoms of PCOS, but they all have proven weight loss efficacy.

 

Table 7: Drugs Approved for the Treatment of Obesity in the U.S. (All are contraindicated during pregnancy or with known hypersensitivity to the drug).

 

Generic Name(s) Mechanism of action Relative Weight Loss compared to other drugs(239) Contraindications and Cautions Warnings about Rare Side Effects Common Side Effects
Orlistat Gastric Lipase inhibitor-inhibits fat absorption Less 1. Reduced gallbladder function

3. Use with caution with pancreatic or liver disease

Some patients may develop increased levels of urinary oxalate and kidney stones 1.steatorrhea

2. diarrhea

3. flatulence

4. increased stooling

 

Phentermine Central appetite suppressant, sympathomimetic amine Better

 

Intended as short term agent (< 6 mos),

1. History of cardiovascular disease

2. During or within 2 weeks following the administration of monoamine oxidase inhibitors

3. Hyperthyroidism.

4. Glaucoma.

5. Agitated states.

6. History of drug abuse

1. May impair the ability of the patient to engage in potentially hazardous activities such as operating machinery or driving a motor vehicle

2. Abuse or addiction

1.feeling restless; 2.headache, 3.dizziness, 4. tremors; 5.poor sleep,

6. dry mouth

Lorcarserin Central appetite suppressant, aserotonin 2C receptor agonist Less caution if

1. renal failure

2. CHF, bradycardia or heart block

3. diabetes mellitus

4, priapism

or penile deformities

5. depression

1. valvular heart disease (symptoms: shortness of breath, edemas)

2. mental illness (depression, suicidal mood)

3. serotonin   neuroleptic malignant syndrome (symptoms: excitement, nausea, sweating, tachycardia)

1.hypoglycemia

2. mental issues

3. bradycardia

4.headache

5. dizziness

6.drowsiness

7.fatigue,

8. nausea

9.dry mouth

10.constipation

11. painful erections

Liraglutide Central appetite suppressant, long-acting glucagon-like peptide-1 receptor agonist Better 1. Personal or family history of medullary thyroid carcinoma (MTC) or in patients with Multiple Endocrine Neoplasia syndrome type 2 (MEN 2).

2. Avoid in patients with history or pancreatitis

Caution

1. May

1. Possible thyroid tumors including cancer

2. Pancreatitis

1. nausea/ vomiting

2.hypoglycemia

3. diarrhea

4.constipation

5.headache

6. fatigue

7. dizziness

8. increased lipase

Phentermine/ Topiramate Phentermine Central appetite suppressant, sympathomimetic amine

 

Topiramate is an anticonvulsant that has weight loss side effects

Best 1. History of cardiovascular disease

2. During or within 2 weeks following the administration of monoamine oxidase inhibitors

3. Hyperthyroidism.

4. Glaucoma.

5. Agitated states.

6. History of drug abuse

1. Suicidal behavior and ideation

2. Acute myopia and secondary angle glaucoma

3. Metabolic acidosis

4. Elevation in creatinine

5. CNS depression with concomitant CNS depressants including alcohol

6. Potential seizures with abrupt withdrawal of drug

7. Patients with renal impairment

8. Kidney stones

9. Oligohidrosis and hyperthermia

10.Hypokalemia

1.mild dizziness 2.anxiety

3. fatigue or irritable

4. constipation

5. memory problems

6. poor sleep

7. numbness of tingly feeling

8. altered sense of taste

9.dry mouth

Naltrexone/ Buproprion naltrexone, an opioid antagonist.

 

bupropion, a relatively weak inhibitor of the neuronal reuptake of dopamine and norepinephrine

Better 1. History of seizures

2. History of an eating disorder

3. Taking opioid pain medicines, 4. Taking medicines to stop opioid addiction,

5. taking an MAOI within 2 weeks.

6.Abrupt termination of alcohol, benzodiazepines, barbiturates, or antiepileptic drugs

1. Suicidal thoughts and behaviors

2. Neuropsychiatric reactions

1.nausea

2.headache

3.vomiting

4.constipation

5.diarrhea

6.dizziness

7. Poor

8. dry mouth

 

 

 

 

Spironolactone. Spironolactone is primarily used to treat hirsutism and acne and appears effective, even though the evidence is limited (235). Spironolactone is a diuretic and aldosterone antagonist and also binds to the androgen receptor as an antagonist. It has other mechanisms of action, including inhibition of ovarian and adrenal steroidogenesis, competition for androgen receptors in hair follicles, and direct inhibition of 5a-reductase activity. The usual dose is 25–100 mg twice a day, and the dose is titrated to balance efficacy while avoiding side effects such as orthostatic hypotension. A full clinical effect may take 6 months or more. About 20% of women using spironolactone will experience increased menstrual frequency (236). Because it can cause and exacerbate hyperkalemia, spironolactone should be used cautiously in women with renal impairment. Because of its mechanism of action as an androgen receptor antagonist, it is contraindicated in women seeking or at risk for pregnancy due to potential teratogenic effects on the formation of male external genitalia. Rarely, however has exposure resulted in ambiguous genitalia in male infants. Concurrent use of an OCP with spironolactone will eliminate the risk of an unplanned pregnancy in compliant patients.

 

Flutamide. Flutamide, an androgen-receptor agonist, is another nonsteroidal anti-androgen that has been shown to be effective against hirsutism in smaller trials The most common side effect is dry skin, but its use has been associated with hepatitis in rare cases. The common dosage is 250 mg/d. The risk of teratogenicity with this compound is significant, and contraception should be used. Flutamide has also been combined with lifestyle and metformin therapy for treatment of PCOS and may have additive effects (237).

5a-reductase inhibitors. Finasteride inhibits two forms of the enzyme 5a-reductase, type 2 and 3 (type 1, predominantly found in the skin and scalp, and type 3, predominantly found in the prostate and reproductive tissues and type 3, widely expressed in adults). It is available as a 5-mg tablet for the treatment of prostate cancer and as a 1-mg tablet for the treatment of male alopecia. Finasteride is better tolerated than other anti-androgens, with minimal hepatic and renal toxicity; however, it has well-documented risk for teratogenicity in male fetuses, and adequate contraception should be used. Overall, randomized trials have found that spironolactone, flutamide and finasteride have similar efficacy in improving hirsutism (235). There are other newer and more comprehensive 5a-reductase inhibitors, including dutasteride that inhibits all isoforms of 5a-reductase, which have not been thoroughly studied for hirsutism or PCOS.

 

Ornithine decarboxylase inhibitors. These have been developed for the treatment of female hirsutism. Ornithine decarboxylase is necessary for the production of polyamines and is also a sensitive and specific marker of androgen action in the prostate. Inhibition of this enzyme limits cell division and function in the pilosebaceous unit. Recently a potent inhibitor of this enzyme, eflornithine, has been found to be effective as a facial crème for the treatment of unwanted facial hair (238) (Brand name Vaniqa). It is available as a 13.9% crème of eflornithine hydrochloride and is applied to affected areas twice daily. In clinical trials, 32% of patients had marked improvement after 24 weeks compared to 8% of placebo-treated women, and the benefit was first noted at eight weeks. It is a pregnancy category C drug. It appears to be well tolerated, with only about 2% of patients developing skin irritation or other adverse reactions.

 

Mechanical and cosmetic means of hair reduction and destruction. Mechanical hair removal (shaving, plucking, waxing, depilatory creams, electrolysis, and laser vaporization) can assist in controlling hirsutism, and often constitute the front-line of treatment used by women.

Electrolysis (i.e., electroepilation) results in long-term hair destruction, albeit slowly. The main objective of laser therapy for hair removal is to selectively cause thermal damage of the hair follicle without destroying adjacent tissues, a process termed selective photothermolysis. In general, laser hair removal is most successful in patients with lighter skin who have dark colored hairs, although therapies are being developed for those with darker skin. However, repeated therapies are necessary, and complete and permanent hair removal is rarely achieved.   After laser-assisted hair removal, most patients experience erythema and edema lasting no more than 48 hours. Blistering or crusting may occur in some patients, as well as some changes in skin pigmentation.

 

Conclusion

PCOS is a heterogeneous disorder with varying diagnostic criteria. The core criteria are hyperandrogenism, either clinical (i.e., hirsutism) or biochemical (i.e., elevated free testosterone or free androgen index), oligomenorrhea reflective of oligo-ovulation, and polycystic ovaries.     The Rotterdam criteria are increasingly accepted as the core diagnostic criteria. Women with PCOS tend to be insulin resistant, obese, and at risk for diabetes and an adverse cardiovascular risk profile.   Treatment tends to be symptom based, with focused treatments for infertility, obesity, hirsutism, etc.   Few therapies address all signs and symptoms of the syndrome. It is hoped that a deeper understanding of the genetics and pathophysiology of the syndrome will lead to more specific therapies.

 

REFERENCES

  1. Stein IF, Leventhal ML. Amenorrhea associated with polycystic ovaries. American journal of obstetrics and gynecology 1935; 29:181-191
  2. Zawadski JK, Dunaif A. Diagnostic criteria for polycystic ovary syndrome; towards a rational approach. In: Dunaif A, Givens JR, Haseltine FP, Merriam GR, eds. Polycystic Ovary Syndrome. Boston: Blackwell Scientific; 1992:377-384.
  3. Adams J, Polson DW, Franks S. Prevalence of polycystic ovaries in women with anovulation and idiopathic hirsutism. British Medical Journal 1986; 293:355-359
  4. Revised 2003 consensus on diagnostic criteria and long-term health risks related to polycystic ovary syndrome. Fertility and Sterility 2004; 81:19-25
  5. group TREAsPcw. Revised 2003 consensus on diagnostic criteria and long-term health risks related to Polycystic Ovary Syndrome (PCOS). Hum Reprod 2004; 19:41-47
  6. Balen AH, Laven JS, Tan SL, Dewailly D. Ultrasound assessment of the polycystic ovary: international consensus definitions. Human reproduction update 2003; 9:505-514
  7. Dewailly D, Lujan ME, Carmina E, Cedars MI, Laven J, Norman RJ, Escobar-Morreale HF. Definition and significance of polycystic ovarian morphology: a task force report from the Androgen Excess and Polycystic Ovary Syndrome Society. Human reproduction update 2014; 20:334-352
  8. Pigny P, Jonard S, Robert Y, Dewailly D. Serum anti-Mullerian hormone as a surrogate for antral follicle count for definition of the polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2006; 91:941-945
  9. Fanchin R, Schonauer LM, Righini C, Guibourdenche J, Frydman R, Taieb J. Serum anti-Mullerian hormone is more strongly related to ovarian follicular status than serum inhibin B, estradiol, FSH and LH on day 3. Hum Reprod 2003; 18:323-327
  10. Azziz R, Carmina E, Dewailly D, Diamanti-Kandarakis E, Escobar-Morreale HF, Futterweit W, Janssen OE, Legro RS, Norman RJ, Taylor AE, Witchel SF. Criteria for defining polycystic ovary syndrome as a predominantly hyperandrogenic syndrome: an androgen excess society guideline. The Journal of clinical endocrinology and metabolism 2006; 91:4237-4245
  11. Legro RS, Arslanian SA, Ehrmann DA, Hoeger KM, Murad MH, Pasquali R, Welt CK, Endocrine S. Diagnosis and treatment of polycystic ovary syndrome: an Endocrine Society clinical practice guideline. The Journal of clinical endocrinology and metabolism 2013; 98:4565-4592
  12. Executive Summary of National Institutes of Health Evidence-based Methodology Workshop on Polycystic Ovary Syndrome. 2012; http://prevention.nih.gov/workshops/2012/pcos/docs/PCOS_Final_Statement.pdf. Accessed September 13, 2013.
  13. Dunaif A, Fauser BC. Renaming PCOS--a two-state solution. The Journal of clinical endocrinology and metabolism 2013; 98:4325-4328
  14. Polson DW, Adams J, Wadsworth J, Franks S. Polycystic ovaries--a common finding in normal women. Lancet 1988; 1:870-872
  15. Johnstone EB, Rosen MP, Neril R, Trevithick D, Sternfeld B, Murphy R, Addauan-Andersen C, McConnell D, Pera RR, Cedars MI. The polycystic ovary post-rotterdam: a common, age-dependent finding in ovulatory women without metabolic significance. The Journal of clinical endocrinology and metabolism 2010; 95:4965-4972
  16. Clayton RN, Ogden V, Hodgkinson J, Worswick L, Rodin DA, Dyer S. How common are polycystic ovaries in normal women and what is their significance for the fertility of the population? [see comments]. Clin Endocrinol 1992; 37:127-134
  17. Azziz R, Woods KS, Reyna R, Key TJ, Knochenhauer ES, Yildiz BO. The prevalence and features of the polycystic ovary syndrome in an unselected population. The Journal of clinical endocrinology and metabolism 2004; 89:2745-2749
  18. Coney P, Ladson G, Sweet S, Legro RS. Does polycystic ovary syndrome increase the disparity in metabolic syndrome and cardiovascular-related health for African-American women? Seminars in reproductive medicine 2008; 26:35-38
  19. Broekmans FJ, Knauff EA, Valkenburg O, Laven JS, Eijkemans MJ, Fauser BC. PCOS according to the Rotterdam consensus criteria: Change in prevalence among WHO-II anovulation and association with metabolic factors. BJOG : an international journal of obstetrics and gynaecology 2006; 113:1210-1217
  20. Ehrmann DA. Polycystic ovary syndrome. The New England journal of medicine 2005; 352:1223-1236
  21. Balen AH, Conway GS, Kaltsas G, Techatrasak K, Manning PJ, West C. Polycystic ovary syndrome: the spectrum of the disorder in 1741 patients. Hum Reprod 1995; 10:2107-2111
  22. Shi Y, Zhao H, Shi Y, Cao Y, Yang D, Li Z, Zhang B, Liang X, Li T, Chen J, Shen J, Zhao J, You L, Gao X, Zhu D, Zhao X, Yan Y, Qin Y, Li W, Yan J, Wang Q, Zhao J, Geng L, Ma J, Zhao Y, He G, Zhang A, Zou S, Yang A, Liu J, Li W, Li B, Wan C, Qin Y, Shi J, Yang J, Jiang H, Xu JE, Qi X, Sun Y, Zhang Y, Hao C, Ju X, Zhao D, Ren CE, Li X, Zhang W, Zhang Y, Zhang J, Wu D, Zhang C, He L, Chen ZJ. Genome-wide association study identifies eight new risk loci for polycystic ovary syndrome. Nature genetics 2012; 44:1020-1025
  23. Chen ZJ, Zhao H, He L, Shi Y, Qin Y, Shi Y, Li Z, You L, Zhao J, Liu J, Liang X, Zhao X, Zhao J, Sun Y, Zhang B, Jiang H, Zhao D, Bian Y, Gao X, Geng L, Li Y, Zhu D, Sun X, Xu JE, Hao C, Ren CE, Zhang Y, Chen S, Zhang W, Yang A, Yan J, Li Y, Ma J, Zhao Y. Genome-wide association study identifies susceptibility loci for polycystic ovary syndrome on chromosome 2p16.3, 2p21 and 9q33.3. Nature genetics 2011; 43:55-59
  24. Legro RS, Myers ER, Barnhart HX, Carson SA, Diamond MP, Carr BR, Schlaff WD, Coutifaris C, McGovern PG, Cataldo NA, Steinkampf MP, Nestler JE, Gosman G, Guidice LC, Leppert PC. The Pregnancy in Polycystic Ovary Syndrome study: baseline characteristics of the randomized cohort including racial effects. Fertility and sterility 2006; 86:914-933
  25. Legro RS, Brzyski RG, Diamond MP, Coutifaris C, Schlaff WD, Casson P, Christman GM, Huang H, Yan Q, Alvero R, Haisenleder DJ, Barnhart KT, Bates GW, Usadi R, Lucidi S, Baker V, Trussell JC, Krawetz SA, Snyder P, Ohl D, Santoro N, Eisenberg E, Zhang H, Network NRM. Letrozole versus clomiphene for infertility in the polycystic ovary syndrome. The New England journal of medicine 2014; 371:119-129
  26. Yildiz BO, Knochenhauer ES, Azziz R. Impact of obesity on the risk for polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2008; 93:162-168
  27. Teede HJ, Joham AE, Paul E, Moran LJ, Loxton D, Jolley D, Lombard C. Longitudinal weight gain in women identified with polycystic ovary syndrome: results of an observational study in young women. Obesity (Silver Spring) 2013; 21:1526-1532
  28. Day FR, Hinds DA, Tung JY, Stolk L, Styrkarsdottir U, Saxena R, Bjonnes A, Broer L, Dunger DB, Halldorsson BV, Lawlor DA, Laval G, Mathieson I, McCardle WL, Louwers Y, Meun C, Ring S, Scott RA, Sulem P, Uitterlinden AG, Wareham NJ, Thorsteinsdottir U, Welt C, Stefansson K, Laven JS, Ong KK, Perry JR. Causal mechanisms and balancing selection inferred from genetic associations with polycystic ovary syndrome. Nat Commun 2015; 6:8464
  29. Hayes MG, Urbanek M, Ehrmann DA, Armstrong LL, Lee JY, Sisk R, Karaderi T, Barber TM, McCarthy MI, Franks S, Lindgren CM, Welt CK, Diamanti-Kandarakis E, Panidis D, Goodarzi MO, Azziz R, Zhang Y, James RG, Olivier M, Kissebah AH, Reproductive Medicine N, Stener-Victorin E, Legro RS, Dunaif A. Genome-wide association of polycystic ovary syndrome implicates alterations in gonadotropin secretion in European ancestry populations. Nat Commun 2015; 6:7502
  30. McAllister JM, Modi B, Miller BA, Biegler J, Bruggeman R, Legro RS, Strauss JF, 3rd. Overexpression of a DENND1A isoform produces a polycystic ovary syndrome theca phenotype. Proceedings of the National Academy of Sciences of the United States of America 2014; 111:E1519-1527
  31. Nelson-DeGrave VL, Wickenheisser JK, Cockrell JE, Wood JR, Legro RS, Strauss JFr, McAllister JM. Valproate potentiates androgen biosynthesis in human ovarian theca cells. Endocrinology 2004 Feb; 145:799-808
  32. Isojarvi JI, Laatikainen TJ, Pakarinen AJ, Juntunen KT, Myllyla VV. Polycystic ovaries and hyperandrogenism in women taking valproate for epilepsy. The New England journal of medicine 1993; 329:1383-1388
  33. Vagi SJ, Azziz-Baumgartner E, Sjodin A, Calafat AM, Dumesic D, Gonzalez L, Kato K, Silva MJ, Ye X, Azziz R. Exploring the potential association between brominated diphenyl ethers, polychlorinated biphenyls, organochlorine pesticides, perfluorinated compounds, phthalates, and bisphenol A in polycystic ovary syndrome: a case-control study. BMC Endocr Disord 2014; 14:86
  34. Rebar R, Judd HL, Yen SS, Rakoff J, Vandenberg G, Naftolin F. Characterization of the inappropriate gonadotropin secretion in polycystic ovary syndrome. The Journal of clinical investigation 1976; 57:1320-1329
  35. Barnes RB, Rosenfield RL, Burstein S, Ehrmann DA. Pituitary-ovarian responses to nafarelin testing in the polycystic ovary syndrome. The New England journal of medicine 1989; 320:559-565
  36. Taylor AE, McCourt B, Martin KA, Anderson EJ, Adams JM, Schoenfeld DH. Determinants of abnormal gonadotropin secretion in clinically defined women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1997; 82:2248-2256
  37. Morales AJ, Laughlin GA, Butzow T, Maheshwari H, Baumann G, Yen SSC. Insulin, somatotropic, and luteinizing hormone axes in lean and obese women with polycystic ovary syndrome - common and distinct features. The Journal of clinical endocrinology and metabolism 1996; 81:2854-2864
  38. McCartney CR, Blank SK, Prendergast KA, Chhabra S, Eagleson CA, Helm KD, Yoo R, Chang RJ, Foster CM, Caprio S, Marshall JC. Obesity and sex steroid changes across puberty: evidence for marked hyperandrogenemia in pre- and early pubertal obese girls. The Journal of clinical endocrinology and metabolism 2007; 92:430-436
  39. McGee WK, Bishop CV, Bahar A, Pohl CR, Chang RJ, Marshall JC, Pau FK, Stouffer RL, Cameron JL. Elevated androgens during puberty in female rhesus monkeys lead to increased neuronal drive to the reproductive axis: a possible component of polycystic ovary syndrome. Hum Reprod 2012; 27:531-540
  40. Schneyer AL, Fujiwara T, Fox J, Welt CK, Adams J, Messerlian GM, Taylor AE. Dynamic changes in the intrafollicular inhibin/activin/follistatin axis during human follicular development: relationship to circulating hormone concentrations. The Journal of clinical endocrinology and metabolism 2000; 85:3319-3330
  41. Nelson VL, Legro RS, Strauss JF, III, McAllister JM. Augmented androgen production is a stable steroidogenic phenotype of propagated theca cells from polycystic ovaries. Molecular Endocrinology 1999; 13:946-957
  42. Kahsar-Miller MD, Nixon C, Boots LR, Go RC, Azziz R. Prevalence of polycystic ovary syndrome (PCOS) in first-degree relatives of patients with PCOS. Fertility and sterility 2001; 75:53-58
  43. Sam S, Legro RS, Essah PA, Apridonidze T, Dunaif A. Evidence for metabolic and reproductive phenotypes in mothers of women with polycystic ovary syndrome. Proceedings of the National Academy of Sciences of the United States of America 2006; 103:7030-7035
  44. Legro RS, Driscoll D, Strauss JF, 3rd, Fox J, Dunaif A. Evidence for a genetic basis for hyperandrogenemia in polycystic ovary syndrome. Proceedings of the National Academy of Sciences of the United States of America 1998; 95:14956-14960
  45. Kumar A, Woods KS, Bartolucci AA, Azziz R. Prevalence of adrenal androgen excess in patients with the polycystic ovary syndrome (PCOS). Clinical endocrinology 2005; 62:644-649
  46. Legro RS, Kunselman AR, Demers L, Wang SC, Bentley-Lewis R, Dunaif A. Elevated dehydroepiandrosterone sulfate levels as the reproductive phenotype in the brothers of women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2002; 87:2134-2138
  47. Stewart DR, Dombroski B, Urbanek M, Ankener W, Ewens KG, Wood JR, Legro RS, Strauss JF, 3rd, Dunaif A, Spielman RS. Fine Mapping of Genetic Susceptibility to Polycystic Ovary Syndrome on Chromosome 19p13.2 and Tests for Regulatory Activity. The Journal of clinical endocrinology and metabolism 2006;
  48. Webber LJ, Stubbs S, Stark J, Trew GH, Margara R, Hardy K, Franks S. Formation and early development of follicles in the polycystic ovary. Lancet 2003 Sep 27; 362:1017-1021
  49. Webber LJ, Stubbs SA, Stark J, Margara RA, Trew GH, Lavery SA, Hardy K, Franks S. Prolonged survival in culture of preantral follicles from polycystic ovaries. The Journal of clinical endocrinology and metabolism 2007; 92:1975-1978
  50. Dunaif A, Segal KR, Futterweit W, Dobrjansky A. Profound peripheral insulin resistance, independent of obesity, in polycystic ovary syndrome. Diabetes 1989; 38:1165-1174
  51. Dunaif A, Finegood DT. Beta-cell dysfunction independent of obesity and glucose intolerance in the polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1996; 81:942-947
  52. Chang RJ, Nakamura RM, Judd HL, Kaplan SA. Insulin resistance in nonobese patients with polycystic ovarian disease. The Journal of clinical endocrinology and metabolism 1983; 57:356-359
  53. O'Meara NM, Blackman JD, Ehrmann DA, Barnes RB, Jaspan JB, Rosenfield RL, Polonsky KS. Defects in beta-cell function in functional ovarian hyperandrogenism. The Journal of clinical endocrinology and metabolism 1993; 76:1241-1247
  54. Ehrmann DA, Schneider DJ, Sobel BE, Cavaghan MK, Imperial J, Polonsky KS. Troglitazone improves defects in insulin action, insulin secretion, ovarian steroidogenesis, and fibrinolysis in women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1997; 82:2108-2116
  55. Adashi EY, Hsueh AJ, Yen SS. Insulin enhancement of luteinizing hormone and follicle-stimulating hormone release by cultured pituitary cells. Endocrinology 1981; 108:1441-1449
  56. Nandi A, Kitamura Y, Kahn CR, Accili D. Mouse models of insulin resistance. Physiol Rev 2004; 84:623-647
  57. Dunaif A. Insulin resistance and the polycystic ovary syndrome: mechanism and implications for pathogenesis. . Endocrine reviews 1997; 18:774-800
  58. Barbieri RL, Ryan KJ. Hyperandrogenism, insulin resistance, and acanthosis nigricans syndrome: a common endocrinopathy with distinct pathophysiologic features. [Review] [70 refs] American Journal of Obstetrics & Gynecology 1983 Sep 1; 147:90-101
  59. Ehrmann DA, Liljenquist DR, Kasza K, Azziz R, Legro RS, Ghazzi MN. Prevalence and predictors of the metabolic syndrome in women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2006; 91:48-53
  60. Polderman KH, Gooren LJ, Asscheman H, Bakker A, Heine RJ. Induction of insulin resistance by androgens and estrogens. The Journal of clinical endocrinology and metabolism 1994; 79:265-271
  61. Willis D, Franks S. Insulin action in human granulosa cells from normal and polycystic ovaries is mediated by the insulin receptor and not the type-I insulin-like growth factor receptor. The Journal of clinical endocrinology and metabolism 1995; 80:3788-3790
  62. Nestler JE, Powers LP, Matt DW, Steingold KA, Plymate SR, RittmasterRS., Clore JN, Blackard WG. A direct effect of hyperinsulinemia on serum sex hormone-binding globulin levels in obese women with the polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1991; 72:83-89
  63. Cumming DC, Wall SR. Non-sex hormone-binding globulin-bound testosterone as a marker for hyperandrogenism. The Journal of clinical endocrinology and metabolism 1985; 61:873-876
  64. Nestler JE. Metformin for the treatment of the polycystic ovary syndrome. The New England journal of medicine 2008; 358:47-54
  65. Legro RS, Dodson WC, Gnatuk CL, Estes SJ, Kunselman AR, Meadows JW, Kesner JS, Krieg EF, Jr., Rogers AM, Haluck RS, Cooney RN. Effects of Gastric Bypass Surgery on Female Reproductive Function. The Journal of clinical endocrinology and metabolism 2012;
  66. Stener-Victorin E, Holm G, Labrie F, Nilsson L, Janson PO, Ohlsson C. Are there any sensitive and specific sex steroid markers for polycystic ovary syndrome? The Journal of clinical endocrinology and metabolism 2010; 95:810-819
  67. Goldzieher JW, Axelrod LR. Clinical and biochemical features of polycystic ovarian disease. Fertility and sterility 1963; 14:631-653
  68. Teilmann G, Pedersen CB, Jensen TK, Skakkebaek NE, Juul A. Prevalence and incidence of precocious pubertal development in Denmark: an epidemiologic study based on national registries. Pediatrics 2005; 116:1323-1328
  69. Elting MW, Korsen TJ, Rekers-Mombarg LT, Schoemaker J. Women with polycystic ovary syndrome gain regular menstrual cycles when ageing. . Human Reproduction 2000 Jan; 15:24-28
  70. Winters SJ, Talbott E, Guzick DS, Zborowski J, McHugh KP. Serum testosterone levels decrease in middle age in women with the polycystic ovary syndrome. . Fertility & Sterility 2000 Apr; 73:724-729
  71. Elting MW, Kwee J, Korsen TJ, Rekers-Mombarg LT, Schoemaker J. Aging women with polycystic ovary syndrome who achieve regular menstrual cycles have a smaller follicle cohort than those who continue to have irregular cycles. Fertility and sterility 2003 May; 79:1154-1160
  72. Heijnen EM, Eijkemans MJ, Hughes EG, Laven JS, Macklon NS, Fauser BC. A meta-analysis of outcomes of conventional IVF in women with polycystic ovary syndrome. Human reproduction update 2006; 12:13-21
  73. Roland M. Problems of ovulation induction with clomiphene citrate with report of a case of ovarian hyperstimulation. Obstetrics and gynecology 1970; 35:55-62
  74. Giudice LC. Endometrium in PCOS: Implantation and predisposition to endocrine CA. Best Pract Res Clin Endocrinol Metab 2006; 20:235-244
  75. Diamond MP, Kruger M, Santoro N, Zhang H, Casson P, Schlaff W, Coutifaris C, Brzyski R, Christman G, Carr BR, McGovern PG, Cataldo NA, Steinkampf MP, Gosman GG, Nestler JE, Carson S, Myers EE, Eisenberg E, Legro RS. Endometrial shedding effect on conception and live birth in women with polycystic ovary syndrome. Obstetrics and gynecology 2012; 119:902-908
  76. Thiboutot D, Jabara S, McAllister JM, Sivarajah A, Gilliland K, Cong Z, Clawson G. Human skin is a steroidogenic tissue: steroidogenic enzymes and cofactors are expressed in epidermis, normal sebocytes, and an immortalized sebocyte cell line (SEB-1). J Invest Dermatol 2003; 120:905-914
  77. Legro RS, Carmina E, Stanczyk FZ, Gentzschein E, Lobo RA. Alterations in androgen conjugate levels in women and men with alopecia. Fertility and sterility 1994; 62:744-750
  78. Hardiman P, Pillay OS, Atiomo W. Polycystic ovary syndrome and endometrial carcinoma. Lancet 2003 May 24; 361:1810-1812
  79. Barry JA, Azizia MM, Hardiman PJ. Risk of endometrial, ovarian and breast cancer in women with polycystic ovary syndrome: a systematic review and meta-analysis. Human reproduction update 2014; 20:748-758
  80. Dahlgren E, Friberg LG, Johansson S, Lindstrom B, Oden A, Samsioe G. Endometrial carcinoma; ovarian dysfunction--a risk factor in young women. European Journal of Obstetrics,Gynecology,& Reproductive Biology 1991; 41:143-150
  81. Dahlgren E, Johansson S, Oden A, Lindstrom B, Janson PO. A model for prediction of endometrial cancer. Acta obstetricia et gynecologica Scandinavica 1989; 68:507-510
  82. Vgontzas AN, Legro RS, Bixler EO, Grayev A, Kales A, Chrousos GP. Polycystic ovary syndrome is associated with obstructive sleep apnea and daytime sleepiness: role of insulin resistance. . Journal of Clinical Endocrinology & Metabolism 2001 Feb; 86:517-520
  83. Fogel RB, Malhotra A, Pillar G, Pittman SD, Dunaif A, White DP. Increased prevalence of obstructive sleep apnea syndrome in obese women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2001; 86:1175-1180
  84. Tasali E, Van Cauter E, Hoffman L, Ehrmann DA. Impact of obstructive sleep apnea on insulin resistance and glucose tolerance in women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2008; 93:3878-3884
  85. Flum DR, Belle SH, King WC, Wahed AS, Berk P, Chapman W, Pories W, Courcoulas A, McCloskey C, Mitchell J, Patterson E, Pomp A, Staten MA, Yanovski SZ, Thirlby R, Wolfe B. Perioperative safety in the longitudinal assessment of bariatric surgery. The New England journal of medicine 2009; 361:445-454
  86. Tasali E, Chapotot F, Leproult R, Whitmore H, Ehrmann DA. Treatment of obstructive sleep apnea improves cardiometabolic function in young obese women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2011; 96:365-374
  87. Ramezani-Binabaj M, Motalebi M, Karimi-Sari H, Rezaee-Zavareh MS, Alavian SM. Are women with polycystic ovarian syndrome at a high risk of non-alcoholic Fatty liver disease; a meta-analysis. Hepat Mon 2014; 14:e23235
  88. Ehrmann DA, Kasza K, Azziz R, Legro RS, Ghazzi MN. Effects of race and family history of type 2 diabetes on metabolic status of women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2005; 90:66-71
  89. Legro RS, Kunselman AR, Dodson WC, Dunaif A. Prevalence and predictors of risk for type 2 diabetes mellitus and impaired glucose tolerance in polycystic ovary syndrome: a prospective, controlled study in 254 affected women. . Journal of Clinical Endocrinology & Metabolism 1999; 84:165-169
  90. Gambineri A, Pelusi C, Manicardi E, Vicennati V, Cacciari M, Morselli-Labate AM, Pagotto U, Pasquali R. Glucose intolerance in a large cohort of mediterranean women with polycystic ovary syndrome: phenotype and associated factors. Diabetes 2004; 53:2353-2358
  91. Legro RS, Gnatuk CL, Kunselman AR, Dunaif A. Changes in glucose tolerance over time in women with polycystic ovary syndrome: a controlled study. The Journal of clinical endocrinology and metabolism 2005; 90:3236-3242
  92. Celik C, Tasdemir N, Abali R, Bastu E, Yilmaz M. Progression to impaired glucose tolerance or type 2 diabetes mellitus in polycystic ovary syndrome: a controlled follow-up study. Fertility and sterility 2014; 101:1123-1128 e1121
  93. Lerchbaum E, Schwetz V, Giuliani A, Obermayer-Pietsch B. Assessment of glucose metabolism in polycystic ovary syndrome: HbA1c or fasting glucose compared with the oral glucose tolerance test as a screening method. Hum Reprod 2013; 28:2537-2544
  94. Li HW, Lam KS, Tam S, Lee VC, Yeung TW, Cheung PT, Yeung WS, Ho PC, Ng EH. Screening for dysglycaemia by oral glucose tolerance test should be recommended in all women with polycystic ovary syndrome. Hum Reprod 2015; 30:2178-2183
  95. Legro RS. Polycystic ovary syndrome and cardioivascular disease: A premature association? Endocrine Reviews 2003; In Press
  96. Solomon CG, Hu FB, Dunaif A, Rich-Edwards JE, Stampfer MJ, Willett WC, Speizer FE, Manson JE. Menstrual cycle irregularity and risk for future cardiovascular disease. Journal of Clinical Endocrinology & Metabolism 2002 May; 87:2013-2017
  97. Krentz AJ, von Muhlen D, Barrett-Connor E. Searching for polycystic ovary syndrome in postmenopausal women: evidence of a dose-effect association with prevalent cardiovascular disease. Menopause 2007; 14:284-292
  98. Shaw LJ, Bairey Merz CN, Azziz R, Stanczyk FZ, Sopko G, Braunstein GD, Kelsey SF, Kip KE, Cooper-Dehoff RM, Johnson BD, Vaccarino V, Reis SE, Bittner V, Hodgson TK, Rogers W, Pepine CJ. Postmenopausal women with a history of irregular menses and elevated androgen measurements at high risk for worsening cardiovascular event-free survival: results from the National Institutes of Health--National Heart, Lung, and Blood Institute sponsored Women's Ischemia Syndrome Evaluation. The Journal of clinical endocrinology and metabolism 2008; 93:1276-1284
  99. Merz CN, Shaw LJ, Azziz R, Stanczyk FZ, Sopko G, Braunstein GD, Kelsey SF, Kip KE, Cooper-DeHoff RM, Johnson BD, Vaccarino V, Reis SE, Bittner V, Hodgson TK, Rogers W, Pepine CJ. Cardiovascular Disease and 10-Year Mortality in Postmenopausal Women with Clinical Features of Polycystic Ovary Syndrome. J Womens Health (Larchmt) 2016; 25:875-881
  100. Hart R, Doherty DA. The potential implications of a PCOS diagnosis on a woman's long-term health using data linkage. The Journal of clinical endocrinology and metabolism 2015; 100:911-919
  101. Talbott EO, Guzick DS, Sutton-Tyrrell K, McHugh-Pemu KP, Zborowski JV, Remsberg KE, Kuller LH. Evidence for association between polycystic ovary syndrome and premature carotid atherosclerosis in middle-aged women. . Arteriosclerosis, Thrombosis & Vascular Biology 2000 Nov; 20:2414-2421
  102. Talbott EO, Zborowski JV, Rager JR, Boudreaux MY, Edmundowicz DA, Guzick DS. Evidence for an association between metabolic cardiovascular syndrome and coronary and aortic calcification among women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2004; 89:5454-5461
  103. Christian RC, Dumesic DA, Behrenbeck T, Oberg AL, Sheedy PFn, Fitzpatrick LA. Prevalence and predictors of coronary artery calcification in women with polycystic ovary syndrome. Journal of Clinical Endocrinology & Metabolism 2003 Jun; 88:2562-2568
  104. Roe A, Hillman J, Butts S, Smith M, Rader D, Playford M, Mehta NN, Dokras A. Decreased cholesterol efflux capacity and atherogenic lipid profile in young women with PCOS. The Journal of clinical endocrinology and metabolism 2014; 99:E841-847
  105. Reaven GM. Banting lecture 1988. Role of insulin resistance in human disease. Diabetes 1988; 37:1595-1607
  106. Legro RS, Kunselman AR, Dunaif A. Prevalence and predictors of dyslipidemia in women with polycystic ovary syndrome. The American journal of medicine 2001; 111:607-613
  107. Talbott E, Clerici A, Berga SL, Kuller L, Guzick D, Detre K, Daniels T, Engberg RA. Adverse lipid and coronary heart disease risk profiles in young women with polycystic ovary syndrome: results of a case-control study. . Journal of clinical epidemiology 1998; 51:415-422
  108. Dahlgren E, Janson PO, Johansson S, Lapidus L, Oden A. Polycystic ovary syndrome and risk for myocardial infarction. Evaluated from a risk factor model based on a prospective population study of women. Acta obstetricia et gynecologica Scandinavica 1992; 71:599-604
  109. Holte J, Gennarelli G, Berne C, Bergh T, Lithell H. Elevated ambulatory day-time blood pressure in women with polycystic ovary syndrome: a sign of a pre-hypertensive state? Hum Reprod 1996; 11:23-28
  110. Barnard L, Ferriday D, Guenther N, Strauss B, Balen AH, Dye L. Quality of life and psychological well being in polycystic ovary syndrome. Hum Reprod 2007; 22:2279-2286
  111. Dokras A, Clifton S, Futterweit W, Wild R. Increased prevalence of anxiety symptoms in women with polycystic ovary syndrome: systematic review and meta-analysis. Fertility and sterility 2011;
  112. Dokras A, Clifton S, Futterweit W, Wild R. Increased risk for abnormal depression scores in women with polycystic ovary syndrome: a systematic review and meta-analysis. Obstetrics and gynecology 2011; 117:145-152
  113. Hollinrake E, Abreu A, Maifeld M, Van Voorhis BJ, Dokras A. Increased risk of depressive disorders in women with polycystic ovary syndrome. Fertility and sterility 2007; 87:1369-1376
  114. Cronin L, Guyatt G, Griffith L, Wong E, Azziz R, Futterweit W, Cook D, Dunaif A. Development of a health-related quality-of-life questionnaire (PCOSQ) for women with polycystic ovary syndrome (PCOS). The Journal of clinical endocrinology and metabolism 1998; 83:1976-1987
  115. Dokras A, Sarwer DB, Allison KC, Milman L, Kris-Etherton PM, Kunselman AR, Stetter CM, Williams NI, Gnatuk CL, Estes SJ, Fleming J, Coutifaris C, Legro RS. Weight Loss and Lowering Androgens Predict Improvements in Health-Related Quality of Life in Women With PCOS. The Journal of clinical endocrinology and metabolism 2016; 101:2966-2974
  116. Robinson S, Rodin DA, Deacon A, Wheeler MJ, Clayton RN. Which hormone tests for the diagnosis of polycystic ovary syndrome? [see comments]. British Journal of Obstetrics & Gynaecology 1992; 99:232-238
  117. New MI, Speiser PW. Genetics of adrenal steroid 21-hydroxylase deficiency. [Review] [115 refs]. Endocrine reviews 1986; 7:331-349
  118. Azziz R, Hincapie LA, Knochenhauer ES, Dewailly D, Fox L, Boots LR. Screening for 21-hydroxylase-deficient nonclassic adrenal hyperplasia among hyperandrogenic women: a prospective study. . Fertility & Sterility 1999 Nov; 72:915-925
  119. Nieman LK, Biller BM, Findling JW, Newell-Price J, Savage MO, Stewart PM, Montori VM. The diagnosis of Cushing's syndrome: an Endocrine Society Clinical Practice Guideline. The Journal of clinical endocrinology and metabolism 2008; 93:1526-1540
  120. Waggoner W, Boots LR, Azziz R. Total testosterone and dheas levels as predictors of androgen-secreting neoplasms: a populational study. . Gynecological Endocrinology 1999 Dec; 13:394-400
  121. Legro RS, Chen G, Kunselman AR, Schlaff WD, Diamond MP, Coutifaris C, Carson SA, Steinkampf MP, Carr BR, McGovern PG, Cataldo NA, Gosman GG, Nestler JE, Myers ER, Zhang H, Foulds J, Reproductive Medicine N. Smoking in infertile women with polycystic ovary syndrome: baseline validation of self-report and effects on phenotype. Hum Reprod 2014; 29:2680-2686
  122. Rosner W, Auchus RJ, Azziz R, Sluss PM, Raff H. Position statement: Utility, limitations, and pitfalls in measuring testosterone: an Endocrine Society position statement. The Journal of clinical endocrinology and metabolism 2007; 92:405-413
  123. Wierman ME, Auchus RJ, Haisenleder DJ, Hall JE, Handelsman D, Hankinson S, Rosner W, Singh RJ, Sluss PM, Stanczyk FZ. Editorial: The new instructions to authors for the reporting of steroid hormone measurements. The Journal of clinical endocrinology and metabolism 2014; 99:4375
  124. Legro RS, Schlaff WD, Diamond MP, Coutifaris C, Casson PR, Brzyski RG, Christman GM, Trussell JC, Krawetz SA, Snyder PJ, Ohl D, Carson SA, Steinkampf MP, Carr BR, McGovern PG, Cataldo NA, Gosman GG, Nestler JE, Myers ER, Santoro N, Eisenberg E, Zhang M, Zhang H. Total Testosterone Assays in Women with Polycystic Ovary Syndrome: Precision and Correlation with Hirsutism. The Journal of clinical endocrinology and metabolism 2010;
  125. O'Reilly MW, Taylor AE, Crabtree NJ, Hughes BA, Capper F, Crowley RK, Stewart PM, Tomlinson JW, Arlt W. Hyperandrogenemia predicts metabolic phenotype in polycystic ovary syndrome: the utility of serum androstenedione. The Journal of clinical endocrinology and metabolism 2014; 99:1027-1036
  126. Cheung AP. Ultrasound and menstrual history in predicting endometrial hyperplasia in polycystic ovary syndrome. Obstetrics and gynecology 2001; 98:325-331
  127. Roth LW, Huang H, Legro RS, Diamond MP, Coutifaris C, Carson SA, Steinkampf MP, Carr BR, McGovern PG, Cataldo NA, Gosman GG, Nestler JE, Myers ER, Zhang H, Schlaff WD. Altering hirsutism through ovulation induction in women with polycystic ovary syndrome. Obstetrics and gynecology 2012; 119:1151-1156
  128. McGovern PG, Legro RS, Myers ER, Barnhart HX, Carson SA, Diamond MP, Carr BR, Schlaff WD, Coutifaris C, Steinkampf MP, Nestler JE, Gosman G, Leppert PC, Giudice LC. Utility of screening for other causes of infertility in women with "known" polycystic ovary syndrome. Fertility and sterility 2007; 87:442-444
  129. Consensus on infertility treatment related to polycystic ovary syndrome. Hum Reprod 2008; 23:462-477
  130. Consensus on infertility treatment related to polycystic ovary syndrome. Fertility and sterility 2008; 89:505-522
  131. Balen AH, Morley LC, Misso M, Franks S, Legro RS, Wijeyaratne CN, Stener-Victorin E, Fauser BC, Norman RJ, Teede H. The management of anovulatory infertility in women with polycystic ovary syndrome: an analysis of the evidence to support the development of global WHO guidance. Human reproduction update 2016; 22:687-708
  132. Franik S, Kremer JA, Nelen WL, Farquhar C. Aromatase inhibitors for subfertile women with polycystic ovary syndrome. Cochrane Database Syst Rev 2014; 2:CD010287
  133. Lundblad D, Eliasson M. Silent myocardial infarction in women with impaired glucose tolerance: the Northern Sweden MONICA study. Cardiovascular diabetology 2003; 2:9
  134. Casper RF, Mitwally MF. Review: aromatase inhibitors for ovulation induction. The Journal of clinical endocrinology and metabolism 2006; 91:760-771
  135. Roque M, Tostes AC, Valle M, Sampaio M, Geber S. Letrozole versus clomiphene citrate in polycystic ovary syndrome: systematic review and meta-analysis. Gynecological endocrinology : the official journal of the International Society of Gynecological Endocrinology 2015; 31:917-921
  136. Diamond MP, Legro RS, Coutifaris C, Alvero R, Robinson RD, Casson P, Christman GM, Ager J, Huang H, Hansen KR, Baker V, Usadi R, Seungdamrong A, Bates GW, Rosen RM, Haisenleder D, Krawetz SA, Barnhart K, Trussell JC, Ohl D, Jin Y, Santoro N, Eisenberg E, Zhang H, Network NRM. Letrozole, Gonadotropin, or Clomiphene for Unexplained Infertility. The New England journal of medicine 2015; 373:1230-1240
  137. Tulandi T, Martin J, Al-Fadhli R, Kabli N, Forman R, Hitkari J, Librach C, Greenblatt E, Casper RF. Congenital malformations among 911 newborns conceived after infertility treatment with letrozole or clomiphene citrate. Fertility and sterility 2006;
  138. Sharma S, Ghosh S, Singh S, Chakravarty A, Ganesh A, Rajani S, Chakravarty BN. Congenital malformations among babies born following letrozole or clomiphene for infertility treatment. PloS one 2014; 9:e108219
  139. Tatsumi T, Jwa SC, Kuwahara A, Irahara M, Kubota T, Saito H. No increased risk of major congenital anomalies or adverse pregnancy or neonatal outcomes following letrozole use in assisted reproductive technology. Hum Reprod 2016;
  140. Davies MJ, Moore VM, Willson KJ, Van Essen P, Priest K, Scott H, Haan EA, Chan A. Reproductive technologies and the risk of birth defects. The New England journal of medicine 2012; 366:1803-1813
  141. Tredway D, Schertz JC, Bock D, Hemsey G, Diamond MP. Anastrozole single-dose protocol in women with oligo- or anovulatory infertility: results of a randomized phase II dose-response study. Fertility and sterility 2011; 95:1725-1729 e1721-1728
  142. Tredway D, Schertz JC, Bock D, Hemsey G, Diamond MP. Anastrozole vs. clomiphene citrate in infertile women with ovulatory dysfunction: a phase II, randomized, dose-finding study. Fertility and sterility 2011; 95:1720-1724 e1721-1728
  143. Asch RH, Greenblatt RB. Update on the safety and efficacy of clomiphene citrate as a therapeutic agent. [Review] [59 refs] Journal of Reproductive Medicine 1976 Sep; 17:175-180
  144. Imani B, Eijkemans MJ, te Velde ER, Habbema JD, Fauser BC. A nomogram to predict the probability of live birth after clomiphene citrate induction of ovulation in normogonadotropic oligoamenorrheic infertility. Fertility and sterility 2002; 77:91-97
  145. Gysler M, March CM, Mishell DR, Jr., Bailey EJ. A decade's experience with an individualized clomiphene treatment regimen including its effect on the postcoital test. Fertility and sterility 1982; 37:161-167.
  146. Legro RS, Barnhart HX, Schlaff WD, Carr BR, Diamond MP, Carson SA, Steinkampf MP, Coutifaris C, McGovern PG, Cataldo NA, Gosman GG, Nestler JE, Giudice LC, Leppert PC, Myers ER. Clomiphene, metformin, or both for infertility in the polycystic ovary syndrome. The New England journal of medicine 2007; 356:551-566
  147. Imani B, Eijkemans MJ, te Velde ER, Habbema JD, Fauser BC. Predictors of chances to conceive in ovulatory patients during clomiphene citrate induction of ovulation in normogonadotropic oligoamenorrheic infertility. The Journal of clinical endocrinology and metabolism 1999; 84:1617-1622
  148. Rausch ME, Legro RS, Barnhart HX, Schlaff WD, Carr BR, Diamond MP, Carson SA, Steinkampf MP, McGovern PG, Cataldo NA, Gosman GG, Nestler JE, Giudice LC, Leppert PC, Myers ER, Coutifaris C. Predictors of pregnancy in women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2009; 94:3458-3466
  149. Lobo RA, Granger LR, Davajan V, Mishell DR, Jr. An extended regimen of clomiphene citrate in women unresponsive to standard therapy. Fertility and sterility 1982; 37:762-766
  150. Branigan EF, Estes MA. A randomized clinical trial of treatment of clomiphene citrate-resistant anovulation with the use of oral contraceptive pill suppression and repeat clomiphene citrate treatment. American journal of obstetrics and gynecology 2003; 188:1424-1428; discussion 1429-1430
  151. Lobo RA, Paul W, March CM, Granger L, Kletzky OA. Clomiphene and dexamethasone in women unresponsive to clomiphene alone. Obstetrics and gynecology 1982; 60:497-501
  152. Nestler JE, Jakubowicz DJ, Evans WS, Pasquali R. Effects of metformin on spontaneous and clomiphene-induced ovulation in the polycystic ovary syndrome. The New England journal of medicine 1998; 338:1876-1880
  153. Elnashar A, Abdelmageed E, Fayed M, Sharaf M. Clomiphene citrate and dexamethazone in treatment of clomiphene citrate-resistant polycystic ovary syndrome: a prospective placebo-controlled study. Hum Reprod 2006; 21:1805-1808
  154. Beck J, Boothroyd C, Proctor M, Farquhar C, Hughes E. Oral anti-oestrogens and medical adjuncts for subfertility associated with anovulation. Cochrane Database Syst Rev 2005:CD002249
  155. Hurst BS, Hickman JM, Matthews ML, Usadi RS, Marshburn PB. Novel clomiphene "stair-step" protocol reduces time to ovulation in women with polycystic ovarian syndrome. American journal of obstetrics and gynecology 2009; 200:510 e511-514
  156. Christin-Maitre S, Hugues JN. A comparative randomized multicentric study comparing the step-up versus step-down protocol in polycystic ovary syndrome. Hum Reprod 2003; 18:1626-1631
  157. Homburg R, Hendriks ML, Konig TE, Anderson RA, Balen AH, Brincat M, Child T, Davies M, D'Hooghe T, Martinez A, Rajkhowa M, Rueda-Saenz R, Hompes P, Lambalk CB. Clomifene citrate or low-dose FSH for the first-line treatment of infertile women with anovulation associated with polycystic ovary syndrome: a prospective randomized multinational study. Hum Reprod 2012; 27:468-473
  158. The Rotterdam ESHRE/ASRM sponsored PCOS consensus workshop group: Revised 2003 consensus on diagnostic criteria and long-term health risks related to polycystic ovary syndrome (PCOS). Hum Reprod 2004; 19:41-47
  159. Kalra SK, Ratcliffe SJ, Dokras A. Is the fertile window extended in women with polycystic ovary syndrome? Utilizing the Society for Assisted Reproductive Technology registry to assess the impact of reproductive aging on live-birth rate. Fertility and sterility 2013; 100:208-213
  160. Chen ZJ, Shi Y, Sun Y, Zhang B, Liang X, Cao Y, Yang J, Liu J, Wei D, Weng N, Tian L, Hao C, Yang D, Zhou F, Shi J, Xu Y, Li J, Yan J, Qin Y, Zhao H, Zhang H, Legro RS. Fresh versus Frozen Embryos for Infertility in the Polycystic Ovary Syndrome. The New England journal of medicine 2016; 375:523-533
  161. Farquhar C, Lilford RJ, Marjoribanks J, Vandekerckhove P. Laparoscopic 'drilling' by diathermy or laser for ovulation induction in anovulatory polycystic ovary syndrome. Cochrane Database Syst Rev 2007:CD001122
  162. Bayram N, van Wely M, Kaaijk EM, Bossuyt PM, van der Veen F. Using an electrocautery strategy or recombinant follicle stimulating hormone to induce ovulation in polycystic ovary syndrome: randomised controlled trial. Bmj 2004; 328:192
  163. Lemieux S, Lewis GF, Ben-Chetrit A, Steiner G, Greenblatt EM. Correction of hyperandrogenemia by laparoscopic ovarian cautery in women with polycystic ovarian syndrome is not accompanied by improved insulin sensitivity or lipid-lipoprotein levels. . Journal of Clinical Endocrinology & Metabolism 1999 Nov; 84:4278-4282
  164. Nahuis MJ, Kose N, Bayram N, van Dessel HJ, Braat DD, Hamilton CJ, Hompes PG, Bossuyt PM, Mol BW, van der Veen F, van Wely M. Long-term outcomes in women with polycystic ovary syndrome initially randomized to receive laparoscopic electrocautery of the ovaries or ovulation induction with gonadotrophins. Hum Reprod 2011; 26:1899-1904
  165. Nahuis MJ, Oude Lohuis E, Kose N, Bayram N, Hompes P, Oosterhuis GJ, Kaaijk EM, Cohlen BJ, Bossuyt PP, van der Veen F, Mol BW, van Wely M. Long-term follow-up of laparoscopic electrocautery of the ovaries versus ovulation induction with recombinant FSH in clomiphene citrate-resistant women with polycystic ovary syndrome: an economic evaluation. Hum Reprod 2012; 27:3577-3582
  166. Donesky BW, Adashi EY. Surgical ovulation induction: the role of ovarian diathermy in polycystic ovary syndrome. [Review]. Baillieres Clinical Endocrinology & Metabolism 1996; 10:293-309
  167. Tang T, Lord JM, Norman RJ, Yasmin E, Balen AH. Insulin-sensitising drugs (metformin, rosiglitazone, pioglitazone, D-chiro-inositol) for women with polycystic ovary syndrome, oligo amenorrhoea and subfertility. Cochrane Database Syst Rev 2012; 5:CD003053
  168. Morin-Papunen L, Rantala AS, Unkila-Kallio L, Tiitinen A, Hippelainen M, Perheentupa A, Tinkanen H, Bloigu R, Puukka K, Ruokonen A, Tapanainen JS. Metformin Improves Pregnancy and Live-Birth Rates in Women with Polycystic Ovary Syndrome (PCOS): A Multicenter, Double-Blind, Placebo-Controlled Randomized Trial. The Journal of clinical endocrinology and metabolism 2012;
  169. Palomba S, Falbo A, Orio F, Jr., Zullo F. Effect of preconceptional metformin on abortion risk in polycystic ovary syndrome: a systematic review and meta-analysis of randomized controlled trials. Fertility and sterility 2009; 92:1646-1658
  170. Moll E, Bossuyt PM, Korevaar JC, Lambalk CB, van der Veen F. Effect of clomifene citrate plus metformin and clomifene citrate plus placebo on induction of ovulation in women with newly diagnosed polycystic ovary syndrome: randomised double blind clinical trial. Bmj 2006; 332:1485
  171. Vanky E, Stridsklev S, Heimstad R, Romundstad P, Skogoy K, Kleggetveit O, Hjelle S, von Brandis P, Eikeland T, Flo K, Berg KF, Bunford G, Lund A, Bjerke C, Almas I, Berg AH, Danielson A, Lahmami G, Carlsen SM. Metformin versus placebo from first trimester to delivery in polycystic ovary syndrome: a randomized, controlled multicenter study. The Journal of clinical endocrinology and metabolism 2010; 95:E448-455
  172. Rowan JA, Rush EC, Obolonkin V, Battin M, Wouldes T, Hague WM. Metformin in gestational diabetes: the offspring follow-up (MiG TOFU): body composition at 2 years of age. Diabetes care 2011; 34:2279-2284
  173. Chiswick C, Reynolds RM, Denison F, Drake AJ, Forbes S, Newby DE, Walker BR, Quenby S, Wray S, Weeks A, Lashen H, Rodriguez A, Murray G, Whyte S, Norman JE. Effect of metformin on maternal and fetal outcomes in obese pregnant women (EMPOWaR): a randomised, double-blind, placebo-controlled trial. Lancet Diabetes Endocrinol 2015; 3:778-786
  174. Nestler JE, Jakubowicz DJ. Decreases in ovarian cytochrome P450C17-alpha activity and serum free testosterone after reduction of insulin secretion in polycystic ovary syndrome. The New England journal of medicine 1996; 335:617-623
  175. Knowler WC, Barrett-Connor E, Fowler SE, Hamman RF, Lachin JM, Walker EA, Nathan DM, Diabetes Prevention Program Research G. Reduction in the incidence of type 2 diabetes with lifestyle intervention or metformin. New England Journal of Medicine 2002; 346:393-403
  176. Moghetti P, Castello R, Negri C, Tosi F, Perrone F, Caputo M, Zanolin E, Muggeo M. Metformin effects on clinical features, endocrine and metabolic profiles, and insulin sensitivity in polycystic ovary syndrome: a randomized, double-blind, placebo-controlled 6-month trial, followed by open, long-term clinical evaluation. The Journal of clinical endocrinology and metabolism 2000; 85:139-146
  177. Lord JM, Flight IH, Norman RJ. Metformin in polycystic ovary syndrome: systematic review and meta-analysis. BMJ 2003 Oct 25; 327:951-953
  178. Hoeger KM, Kochman L, Wixom N, Craig K, Miller RK, Guzick DS. A randomized, 48-week, placebo-controlled trial of intensive lifestyle modification and/or metformin therapy in overweight women with polycystic ovary syndrome: a pilot study. Fertility and sterility 2004; 82:421-429
  179. Hoeger K, Davidson K, Kochman L, Cherry T, Kopin L, Guzick DS. The Impact of Metformin, Oral Contraceptives and Lifestyle Modification, on Polycystic Ovary Syndrome in Obese Adolescent Women in Two Randomized, Placebo-Controlled Clinical Trials. The Journal of clinical endocrinology and metabolism 2008;
  180. Tang T, Glanville J, Hayden CJ, White D, Barth JH, Balen AH. Combined lifestyle modification and metformin in obese patients with polycystic ovary syndrome. A randomized, placebo-controlled, double-blind multicentre study. Hum Reprod 2006; 21:80-89
  181. Ladson G, Dodson WC, Sweet SD, Archibong AE, Kunselman AR, Demers LM, Williams NI, Coney P, Legro RS. The effects of metformin with lifestyle therapy in polycystic ovary syndrome: a randomized double-blind study. Fertility and sterility 2011; 95:1059-1066 e1051-1057
  182. Ladson G, Dodson WC, Sweet SD, Archibong AE, Kunselman AR, Demers LM, Lee PA, Williams NI, Coney P, Legro RS. Effects of metformin in adolescents with polycystic ovary syndrome undertaking lifestyle therapy: a pilot randomized double-blind study. Fertility and sterility 2011; 95:2595-2598 e2591-2596
  183. Ghazeeri G, Kutteh WH, Bryer-Ash M, Haas D, Ke RW. Effect of rosiglitazone on spontaneous and clomiphene citrate-induced ovulation in women with polycystic ovary syndrome. Fertility and sterility 2003; 79:562-566
  184. Rouzi AA, Ardawi MS. A randomized controlled trial of the efficacy of rosiglitazone and clomiphene citrate versus metformin and clomiphene citrate in women with clomiphene citrate-resistant polycystic ovary syndrome. Fertility and sterility 2006; 85:428-435
  185. Dunaif A, Scott D, Finegood D, Quintana B, Whitcomb R. The insulin-sensitizing agent troglitazone improves metabolic and reproductive abnormalities in the polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1996; 81:3299-3306
  186. Azziz R, Ehrmann D, Legro RS, Whitcomb RW, Hanley R, Fereshetian AG, O'Keefe M, Ghazzi MN, PCOS/Troglitazone Study G. Troglitazone improves ovulation and hirsutism in the polycystic ovary syndrome: a multicenter, double blind, placebo-controlled trial. . Journal of Clinical Endocrinology & Metabolism 2001 Apr; 86:1626-1632
  187. Baillargeon JP, Jakubowicz DJ, Iuorno MJ, Jakubowicz S, Nestler JE. Effects of metformin and rosiglitazone, alone and in combination, in nonobese women with polycystic ovary syndrome and normal indices of insulin sensitivity. Fertility and sterility 2004; 82:893-902
  188. Elkind-Hirsch K, Marrioneaux O, Bhushan M, Vernor D, Bhushan R. Comparison of single and combined treatment with exenatide and metformin on menstrual cyclicity in overweight women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2008; 93:2670-2678
  189. Jensterle Sever M, Kocjan T, Pfeifer M, Kravos NA, Janez A. Short-term combined treatment with liraglutide and metformin leads to significant weight loss in obese women with polycystic ovary syndrome and previous poor response to metformin. European journal of endocrinology / European Federation of Endocrine Societies 2014; 170:451-459
  190. Vessey MP, Painter R. Endometrial and ovarian cancer and oral contraceptives--findings in a large cohort study. . British journal of cancer 1995 Jun; 71:1340-1342
  191. Schlesselman JJ. Risk of endometrial cancer in relation to use of combined oral contraceptives. A practitioner's guide to meta-analysis. Hum Reprod 1997; 12:1851-1863
  192. Bird ST, Hartzema AG, Brophy JM, Etminan M, Delaney JA. Risk of venous thromboembolism in women with polycystic ovary syndrome: a population-based matched cohort analysis. CMAJ : Canadian Medical Association journal = journal de l'Association medicale canadienne 2013; 185:E115-120
  193. Chasen-Taber L, Willett WC, Stampfer MJ, Hunter DJ, Colditz GA, Spielgelman D, Manson JE. A prospective study of oral contraceptives and NIDDM among U.S. women. Diabetes care 1997; 20:330-335
  194. Meyer C, McGrath BP, Teede HJ. Effects of medical therapy on insulin resistance and the cardiovascular system in polycystic ovary syndrome. Diabetes care 2007; 30:471-478
  195. Korytkowski MT, Mokan M, Horwitz MJ, Berga SL. Metabolic effects of oral contraceptives in women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 1995; 80:3327-3334
  196. Legro RS, Dodson WC, Kris-Etherton PM, Kunselman AR, Stetter CM, Williams NI, Gnatuk CL, Estes SJ, Fleming J, Allison KC, Sarwer DB, Coutifaris C, Dokras A. Randomized Controlled Trial of Preconception Interventions in Infertile Women With Polycystic Ovary Syndrome. The Journal of clinical endocrinology and metabolism 2015; 100:4048-4058
  197. Falsetti L, Pasinetti E. Effects of long-term administration of an oral contraceptive containing ethinylestradiol and cyproterone acetate on lipid metabolism in women with polycystic ovary syndrome. Acta obstetricia et gynecologica Scandinavica 1995; 74:56-60
  198. Costello MF, Shrestha B, Eden J, Johnson NP, Sjoblom P. Metformin versus oral contraceptive pill in polycystic ovary syndrome: a Cochrane review. Hum Reprod 2007;
  199. Bhattacharya SM, Jha A. Comparative study of the therapeutic effects of oral contraceptive pills containing desogestrel, cyproterone acetate, and drospirenone in patients with polycystic ovary syndrome. Fertility and sterility 2012; 98:1053-1059
  200. O'Brien RC, Cooper ME, Murray RM, Seeman E, Thomas AK, Jerums G. Comparison of sequential cyproterone acetate/estrogen versus spironolactone/oral contraceptive in the treatment of hirsutism. The Journal of clinical endocrinology and metabolism 1991; 72:1008-1013
  201. Legro RS, Pauli JG, Kunselman AR, Meadows JW, Kesner JS, Zaino RJ, Demers LM, Gnatuk CL, Dodson WC. Effects of continuous versus cyclical oral contraception: a randomized controlled trial. The Journal of clinical endocrinology and metabolism 2008; 93:420-429
  202. Anttila L, Koskinen P, Erkkola R, Irjala K, Ruutiainen K. Serum testosterone, androstenedione and luteinizing hormone levels after short-term medroxyprogesterone acetate treatment in women with polycystic ovarian disease. Acta obstetricia et gynecologica Scandinavica 1994; 73:634-636
  203. Palep-Singh M, Mook K, Barth J, Balen A. An observational study of Yasmin in the management of women with polycystic ovary syndrome. J Fam Plann Reprod Health Care 2004; 30:163-165
  204. Kriplani A, Singh BM, Lal S, Agarwal N. Efficacy, acceptability and side effects of the levonorgestrel intrauterine system for menorrhagia. International journal of gynaecology and obstetrics: the official organ of the International Federation of Gynaecology and Obstetrics 2007;
  205. Xiao B, Wu SC, Chong J, Zeng T, Han LH, Luukkainen T. Therapeutic effects of the levonorgestrel-releasing intrauterine system in the treatment of idiopathic menorrhagia. Fertility and sterility 2003; 79:963-969
  206. Tamaoka Y, Orikasa H, Sumi Y, Sakakura K, Kamei K, Nagatani M, Ezawa S. Treatment of endometrial hyperplasia with a danazol-releasing intrauterine device: a prospective study. Gynecologic and obstetric investigation 2004; 58:42-48
  207. Lin M, Xu X, Wang Y, Hu Y, Zhao Y. Evaluation of a levonorgestrel-releasing intrauterine system for treating endometrial hyperplasia in patients with polycystic ovary syndrome. Gynecologic and obstetric investigation 2014; 78:41-44
  208. Abu Hashim H, Ghayaty E, El Rakhawy M. Levonorgestrel-releasing intrauterine system vs oral progestins for non-atypical endometrial hyperplasia: a systematic review and metaanalysis of randomized trials. American journal of obstetrics and gynecology 2015; 213:469-478
  209. Laurelli G, Falcone F, Gallo MS, Scala F, Losito S, Granata V, Cascella M, Greggi S. Long-Term Oncologic and Reproductive Outcomes in Young Women With Early Endometrial Cancer Conservatively Treated: A Prospective Study and Literature Update. Int J Gynecol Cancer 2016; 26:1650-1657
  210. Sun J, Yuan Y, Cai R, Sun H, Zhou Y, Wang P, Huang R, Xia W, Wang S. An investigation into the therapeutic effects of statins with metformin on polycystic ovary syndrome: a meta-analysis of randomised controlled trials. BMJ Open 2015; 5:e007280
  211. Sathyapalan T, Kilpatrick ES, Coady AM, Atkin SL. The effect of atorvastatin in patients with polycystic ovary syndrome: a randomized double-blind placebo-controlled study. The Journal of clinical endocrinology and metabolism 2009; 94:103-108
  212. Raja-Khan N, Kunselman AR, Hogeman CS, Stetter CM, Demers LM, Legro RS. Effects of atorvastatin on vascular function, inflammation, and androgens in women with polycystic ovary syndrome: a double-blind, randomized, placebo-controlled trial. Fertility and sterility 2011; 95:1849-1852
  213. Banaszewska B, Pawelczyk L, Spaczynski RZ, Dziura J, Duleba AJ. Effects of simvastatin and oral contraceptive agent on polycystic ovary syndrome: prospective, randomized, crossover trial. The Journal of clinical endocrinology and metabolism 2007; 92:456-461
  214. Banaszewska B, Pawelczyk L, Spaczynski RZ, Duleba AJ. Effects of simvastatin and metformin on polycystic ovary syndrome after six months of treatment. The Journal of clinical endocrinology and metabolism 2011; 96:3493-3501
  215. Ridker PM, Danielson E, Fonseca FA, Genest J, Gotto AM, Jr., Kastelein JJ, Koenig W, Libby P, Lorenzatti AJ, MacFadyen JG, Nordestgaard BG, Shepherd J, Willerson JT, Glynn RJ. Rosuvastatin to prevent vascular events in men and women with elevated C-reactive protein. The New England journal of medicine 2008; 359:2195-2207
  216. Sjostrom L, Lindroos AK, Peltonen M, Torgerson J, Bouchard C, Carlsson B, Dahlgren S, Larsson B, Narbro K, Sjostrom CD, Sullivan M, Wedel H. Lifestyle, diabetes, and cardiovascular risk factors 10 years after bariatric surgery. The New England journal of medicine 2004; 351:2683-2693
  217. Moran LJ, Brinkworth G, Noakes M, Norman RJ. Effects of lifestyle modification in polycystic ovarian syndrome. Reproductive biomedicine online 2006; 12:569-578
  218. Moran LJ, Noakes M, Clifton PM, Wittert GA, Williams G, Norman RJ. Short-term meal replacements followed by dietary macronutrient restriction enhance weight loss in polycystic ovary syndrome. The American journal of clinical nutrition 2006; 84:77-87
  219. Moran LJ, Ko H, Misso M, Marsh K, Noakes M, Talbot M, Frearson M, Thondan M, Stepto N, Teede HJ. Dietary composition in the treatment of polycystic ovary syndrome: a systematic review to inform evidence-based guidelines. Journal of the Academy of Nutrition and Dietetics 2013; 113:520-545
  220. Sacks FM, Bray GA, Carey VJ, Smith SR, Ryan DH, Anton SD, McManus K, Champagne CM, Bishop LM, Laranjo N, Leboff MS, Rood JC, de Jonge L, Greenway FL, Loria CM, Obarzanek E, Williamson DA. Comparison of weight-loss diets with different compositions of fat, protein, and carbohydrates. The New England journal of medicine 2009; 360:859-873
  221. Vigorito C, Giallauria F, Palomba S, Cascella T, Manguso F, Lucci R, De Lorenzo A, Tafuri D, Lombardi G, Colao A, Orio F. Beneficial effects of a three-month structured exercise training program on cardiopulmonary functional capacity in young women with polycystic ovary syndrome. The Journal of clinical endocrinology and metabolism 2007; 92:1379-1384
  222. Hall KD, Sacks G, Chandramohan D, Chow CC, Wang YC, Gortmaker SL, Swinburn BA. Quantification of the effect of energy imbalance on bodyweight. Lancet 2011; 378:826-837
  223. Jensen MD, Ryan DH, Apovian CM, Ard JD, Comuzzie AG, Donato KA, Hu FB, Hubbard VS, Jakicic JM, Kushner RF, Loria CM, Millen BE, Nonas CA, Pi-Sunyer FX, Stevens J, Stevens VJ, Wadden TA, Wolfe BM, Yanovski SZ. 2013 AHA/ACC/TOS Guideline for the Management of Overweight and Obesity in Adults: A Report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines and The Obesity Society. Circulation 2013;
  224. Robinson MK. Surgical treatment of obesity--weighing the facts. The New England journal of medicine 2009; 361:520-521
  225. Escobar-Morreale HF, Botella-Carretero JI, Alvarez-Blasco F, Sancho J, San Millan JL. The polycystic ovary syndrome associated with morbid obesity may resolve after weight loss induced by bariatric surgery. The Journal of clinical endocrinology and metabolism 2005; 90:6364-6369
  226. Eid GM, Cottam DR, Velcu LM, Mattar SG, Korytkowski MT, Gosman G, Hindi P, Schauer PR. Effective treatment of polycystic ovarian syndrome with Roux-en-Y gastric bypass. Surgery for obesity and related diseases : official journal of the American Society for Bariatric Surgery 2005; 1:77-80
  227. Schauer PR, Bhatt DL, Kirwan JP, Wolski K, Brethauer SA, Navaneethan SD, Aminian A, Pothier CE, Kim ES, Nissen SE, Kashyap SR, Investigators S. Bariatric surgery versus intensive medical therapy for diabetes--3-year outcomes. The New England journal of medicine 2014; 370:2002-2013
  228. Johansson K, Cnattingius S, Naslund I, Roos N, Trolle Lagerros Y, Granath F, Stephansson O, Neovius M. Outcomes of pregnancy after bariatric surgery. The New England journal of medicine 2015; 372:814-824
  229. Coleman KJ, Huang YC, Hendee F, Watson HL, Casillas RA, Brookey J. Three-year weight outcomes from a bariatric surgery registry in a large integrated healthcare system. Surgery for obesity and related diseases : official journal of the American Society for Bariatric Surgery 2014; 10:396-403
  230. Lindholm A, Bixo M, Bjorn I, Wolner-Hanssen P, Eliasson M, Larsson A, Johnson O, Poromaa IS. Effect of sibutramine on weight reduction in women with polycystic ovary syndrome: a randomized, double-blind, placebo-controlled trial. Fertility and sterility 2008; 89:1221-1228
  231. Sabuncu T, Harma M, Nazligul Y, Kilic F. Sibutramine has a positive effect on clinical and metabolic parameters in obese patients with polycystic ovary syndrome. Fertility and sterility 2003; 80:1199-1204
  232. Panidis D, Farmakiotis D, Rousso D, Kourtis A, Katsikis I, Krassas G. Obesity, weight loss, and the polycystic ovary syndrome: effect of treatment with diet and orlistat for 24 weeks on insulin resistance and androgen levels. Fertility and sterility 2008; 89:899-906
  233. Diamanti-Kandarakis E, Piperi C, Alexandraki K, Katsilambros N, Kouroupi E, Papailiou J, Lazaridis S, Koulouri E, Kandarakis HA, Douzinas EE, Creatsas G, Kalofoutis A. Short-term effect of orlistat on dietary glycotoxins in healthy women and women with polycystic ovary syndrome. Metabolism: clinical and experimental 2006; 55:494-500
  234. Jayagopal V, Kilpatrick ES, Holding S, Jennings PE, Atkin SL. Orlistat is as beneficial as metformin in the treatment of polycystic ovarian syndrome. The Journal of clinical endocrinology and metabolism 2005; 90:729-733
  235. Swiglo BA, Cosma M, Flynn DN, Kurtz DM, Labella ML, Mullan RJ, Erwin PJ, Montori VM. Clinical review: Antiandrogens for the treatment of hirsutism: a systematic review and metaanalyses of randomized controlled trials. The Journal of clinical endocrinology and metabolism 2008; 93:1153-1160
  236. Helfer EL, Miller JL, Rose LI. Side-effects of spironolactone therapy in the hirsute woman. . Journal of Clinical Endocrinology & Metabolism 1988 Jan; 66:208-211
  237. Gambineri A, Patton L, Vaccina A, Cacciari M, Morselli-Labate AM, Cavazza C, Pagotto U, Pasquali R. Treatment with flutamide, metformin, and their combination added to a hypocaloric diet in overweight-obese women with polycystic ovary syndrome: a randomized, 12-month, placebo-controlled study. The Journal of clinical endocrinology and metabolism 2006; 91:3970-3980
  238. Wolf JES, D. Huber, F., Jackson, J. Lin, C.S., Mathes, B.M., Schrode, K. and the Eflornithine HCl Study Group. Randomized, double-blind clinical evaluation of the efficacy and safety of topical eflornithine HCL 13.9% cream in the treatment of women with facial hair. Int J Derm 2007; 207:94-98
  239. Khera R, Murad MH, Chandar AK, Dulai PS, Wang Z, Prokop LJ, Loomba R, Camilleri M, Singh S. Association of Pharmacological Treatments for Obesity With Weight Loss and Adverse Events: A Systematic Review and Meta-analysis. JAMA : the journal of the American Medical Association 2016; 315:2424-2434

 

Metabolism of Thyroid Hormone

ABSTRACT


Thyroid hormone is indispensable for normal development and metabolism of most cells and tissues. Thyroid hormones are metabolized by different pathways: glucuronidation, sulfation, and deiodination, the latter being the most important. Three enzymes catalyzing deiodination have been identified, called type 1 (D1), type 2 (D2) and type 3 (D3) iodothyronine deiodinases. D1 and D2 have outer ring deiodinase activity, converting the prohormone T4 to its bioactive form T3 and degrading rT3 to 3,3’-T2. D3 has inner ring deiodinase activity and degrades T4 to rT3 and T3 to 3,3’-T2.

D1 is largely expressed in liver and kidney. Its main role is clearance of rT3 from the circulation and it also contributes to production of plasma T3. D2 is importantly expressed in the central nervous system, pituitary, brown adipose tissue and muscle and, generally, its expression reciprocally responds to changes in thyroid state. D2 serves to adapt cellular thyroid state to changing physiological needs. D3 is importantly expressed in fetal tissues and in adult brain tissue. In addition, D3 can be re-expressed under certain pathological conditions such as critical illness or in specific cancers.

In recent years, the paradigm has evolved that D2 and D3 can locally modify thyroid hormone bioactivity independent of serum thyroid hormone concentrations. Its physiological relevance has been shown in various developmental and regenerative conditions. Future studies may reveal if modifying (local) deiodinase activity can be of use under certain circumstances. For complete coverage of all related areas of Endocrinology, please see our online FREE web-book, www.endotext.org. and WWW.THYROIDMANAGER.ORG.

 

 

CLINICAL SUMMARY

 

In healthy humans the thyroid gland produces predominantly the prohormone T4 together with a small amount of the bioactive hormone T3. Most T3 is produced by enzymatic outer ring deiodination (ORD) of T4 in peripheral tissues. Alternative, inner ring deiodination (IRD) of T4 yields the metabolite rT3, the thyroidal secretion of which is negligible. Normally, about one-third of T4 is converted to T3 and about one-third to rT3. The remainder of T4 is metabolized by different pathways, in particular glucuronidation and sulfation. T3 is further metabolized largely by IRD and rT3 largely by ORD, yielding in both cases the metabolite 3,3’T2. Thus, ORD is regarded as an activating pathway and IRD as an inactivating pathway.

Three enzymes catalyzing these deiodinations have been identified, called type 1 (D1), type 2 (D2) and type 3 (D3) iodothyronine deiodinases. All three deiodinases have been cloned and characterized in a variety of species. Together, they form a family of homologous selenoproteins which consist of »250-280 amino acids, including an essential selenocysteine residue in the active center. It is remarkable, therefore, that production and metabolism of thyroid hormone are dependent on two trace elements, namely iodine and selenium.

D1 is expressed mainly in the liver, the kidneys and the thyroid. In particular the hepatic enzyme is thought to contribute importantly to peripheral T3 production and to be the main site for the clearance of plasma rT3. These processes are mediated by the ORD activity of D1. However, D1 also has IRD activity, especially towards sulfated T4 and T3. Therefore, in addition to the bioactivation of T4 to T3, D1 also catalyzes the degradation of thyroid hormone. An important property distinguishing D1 from the other deiodinases is its sensitivity to inhibition by the anti-thyroid drug propylthiouracil (PTU). The important role of D1 in the peripheral production of plasma T3 has been demonstrated by the marked decrease in plasma T3 levels in T4-substituted athyreotic subjects treated with PTU.

D2 has been studied extensively in the central nervous system, the pituitary, brown adipose tissue and skeletal muscle of experimental animals. D2 has only ORD activity and its expression shows adaptive changes in response to alterations in thyroid state, which serves to maintain tissue T3 levels in the face of varying plasma T4 and T3 levels. Cell-specific modulation of D2 enables to adapt to physiological needs.

D3 mediates the degradation of thyroid hormone since it has only IRD activity. The brain is the predominant D3-expressing tissue in adult animals, and may thus be the main site for the clearance of plasma T3 and for the production of plasma rT3. However, high D3 activities have been demonstrated in the placenta and the pregnant uterus as well as in different fetal tissues. The high D3 activities at these sites appear to prevent exposure of fetal tissues to high T3 levels, allowing the growth of these tissues. T3 is only required at the differentiation stage of tissue development.

Whereas intitial studies focused on the role of the deiodinases in maintaining normal serum T3 concentrations, the paradigm has evolved that these enzymes can locally modify TH bioactivity independent of serum TH concentrations. An example is the critical role of D2 and D3 in cochlear development, since Dio2-/- as well as Dio3-/- mice have severe hearing loss. These enzymes prevent too little or too much hormonal stimulation at inappropriate stages in development. At immature stages, D3 limits stimulation by T3. Postnatally, a double switch occurs with a decline in D3 and an increase D2, resulting in a local T3 surge which is independent of serum T3 levels and triggers the onset of auditory function.

Clinically, the importance of the deiodinases in the regulation of thyroid hormone bioactivity is apparent when their activity is affected by patho-physiological conditions. Examples of such conditions are iodine insufficiency, thyroidal and non-thyroidal illness and malnutrition.

Expression of D1 and D3 is under positive control and that of D2 is under negative control of thyroid hormone. Therefore, the relative contribution of D1 and D2 to peripheral T3 production varies with thyroid state, with D1 prevailing in the hyperthyroid and D2 in the hypothyroid state. The proportions of T3 being produced via D1 or D2 in euthyroid subjects remain to be established.

In iodine deficiency, D1-mediated peripheral T3 production decreases but this is in part compensated by an increased thyroidal T3 production, which is mediated by an increased TSH secretion as well as by increased efficiency of D2-mediated T3 production. Simultaneously, neuronal D3 expression decreases thereby prolonging the local half-life of T3.

In non-thyroidal illness (NTI) plasma T3 is often decreased and plasma rT3 increased; plasma FT4 is still in the normal range depending on the severity of disease. The changes in plasma T3 and rT3 are explained by a diminished conversion of T4 to T3 and of rT3 to 3,3-T2 by D1 in the liver. Although this may be caused to some extent by decreased D1 expression or cofactor levels, a diminished activity of transporter(s) mediating hepatic uptake of T4 and rT3 appears to be another important mechanism. This also holds for the generation of the low T3 syndrome in malnutrition.

In addition to a decreased peripheral T3 production, the low T3 syndrome of NTI may also be caused by stimulated thyroid hormone degradation due to induction of D3 in different tissues. Pathological expression of D3 may be so high that this results in a state of consumptive hypothyroidism with low serum (F)T4 and T3 and very high rT3 levels. This has been shown in different patients with hemangiomas which express very high D3 activities.

Finally, peripheral production of T3 can be inhibited by a variety of drugs, including PTU, dexamethasone, propranolol, and iodinated compounds such as the radiographic agents iopanoic acid and ipodate and the anti-arrhythmic drug amiodarone. PTU is a specific uncompetitive inhibitor of D1, while iopanoic acid and ipodate are competitive inhibitors not only of D1 but also of D2. In addition, the radiographic agents inhibit hepatic uptake of thyroid hormone. Amiodarone and its metabolite desethylamidarone may also interfere with peripheral thyroid hormone levels by inhibition of deiodinase activities and tissue thyroid hormone transport. Little is known about the mechanisms by which propranolol and dexamethasone inhibit peripheral T3 production. Combinations of these drugs (e.g. PTU, ipodate, dexamethasone and/or propranolol) may be used to acutely decrease plasma T3 levels in patients with severe hyperthyroidism (toxic storm).

Thyroid hormone metabolism in humans

In healthy human subjects with an adequate iodine intake, the thyroid gland produces predominantly the prohormone T4 and a small amount of the bioactive thyroid hormone T3. Roughly 80% of T3 is produced by outer ring deiodination (ORD) of T4 in peripheral tissues. The relative contribution of T3 secretion increases in iodine deficiency and other conditions where the thyroid gland is stimulated by TSH or TSH receptor antibodies, since this is associated with increased de novo T3 synthesis and thyroidal expression of both D1 and D2, and thus increased intra-thyroidal T4 to T3 conversion (see below). Nevertheless, there is good agreement that about 1/3 of T4 daily produced (~130 nmol) in normal humans is converted to T3, which corresponds to about 40 nmol and thus 80% of the estimated total daily T3 production of 50 nmol. For recent comprehensive reviews of thyroid hormone metabolism and the role of the iodothyronine deiodinases therein, the reader is referred to (1-5)

 

That most plasma T3 is derived from peripheral conversion of T4 is supported by the fact that normal plasma T3 levels are obtained in athyreotic patients treated with sufficient T4 to achieve high-normal plasma (F)T4 levels. Administration of T4 to hypothyroid rats to achieve normal plasma T4 levels results in subnormal plasma T3 levels not only because of the lack of T3 secretion but also because of a decreased T3 production by D1 in peripheral tissues, since this enzyme is under positive control of T3 itself (6). Other studies in hypothyroid rats suggest that optimal restoration of serum and tissue thyroid hormone levels is achieved by the combined administration of specific amounts of T4 and T3 (7).

Also initial studies in humans suggested that replacement with a combination of T4 and T3 is better than replacement with T4 alone (8). However, this has not been confirmed in a large number of subsequent studies (reviewed in (9, 10)). A common drawback of these trials testing the possible beneficial effects of adding T3 to the T4 replacement therapy is that regular T3 tablets were used. Due to its short half-life, this results in substantial fluctuations of serum T3 levels. It remains to be investigated if administration of T3 in a slow-release formula which better mimics the continuous thyroidal T3 secretion (11) may improve the outcome of combined T4 and T3 replacement. Furthermore, psychological well-being and preference for L-T4 + L-T3 combination therapy may be influenced by polymorphisms in thyroid hormone pathway genes, specifically in thyroid hormone transporters and deiodinases (12-14).

Besides ORD to T3, T4 is converted by inner ring deiodination (IRD) to the metabolite rT3 (Fig. 1), which accounts for about 40% of T4 turnover, while thyroidal secretion of rT3 is negligible. T3 and rT3 undergo further deiodination, predominantly to the common metabolite 3,3'-diiodothyronine (3,3'T2), which is generated by IRD of T3 and by ORD of rT3 (1-5). Thus, ORD is an activating pathway by which the prohormone T4 is converted to active T3, whereas IRD is an inactivating pathway by which T4 and T3 are converted to the metabolites rT3 and 3,3'T2, respectively.

 
Figure 1. Pathways of Thyroid Hormone Metabolism

 

In addition to deiodination, iodothyronines are metabolized by conjugation of the phenolic hydroxyl group with sulfate or glucuronic acid (Fig. 1) (15, 16). Sulfation and glucuronidation are so-called phase II detoxification reactions, the general purpose of which is to increase the water-solubility of the substrates and, thus, to facilitate their biliary and/or urinary clearance. However, iodothyronine sulfate levels are normally very low in plasma, bile and urine, because these conjugates are rapidly degraded by D1, suggesting that sulfate conjugation is a primary step leading to the irreversible inactivation of thyroid hormone (17, 18). Plasma levels and, if investigated, biliary excretion of iodothyronine sulfates are increased by inhibition of D1 activity with PTU or iopanoic acid (IOP), and during fetal development, NTI and fasting (16, 18). Under these conditions, T3 sulfate (T3S) may function as a reservoir of inactive hormone from which active T3 may be recovered by action of tissue sulfatases and bacterial sulfatases in the intestine (15-17).

In contrast to the sulfates, iodothyronine glucuronides are rapidly excreted in the bile. However, this is not an irreversible pathway of hormone disposal. After hydrolysis of the glucuronides by bacterial ß-glucuronidases in the intestines, part of the liberated iodothyronines is reabsorbed, resulting in an enterohepatic cycle of iodothyronines (15, 16). Nevertheless, about 20% of daily T4 production appears in the feces, probably through biliary excretion of glucuronide conjugates.

Thyronamines (TAMs) are a novel class of iodothyronine-like endogenous signaling compounds (19). Their structure differs from T4 and T3 only with regard to the absence of the carboxylate group of the alanine side chain. THs and TAMs are designated Tx and TxAM, respectively, with “x” indicating the number of iodine atoms per molecule, thus following the same rules for nomenclature (see (20) for an excellent review). So far, only 3-iodothyronamine (3-T1AM) and thyronamine (T0AM) have been detected in vivo using liquid chromatography-tandem mass spectrometry (LC-MS/MS) (19, 21). 3T1AM and T0AM have been shown to exert acute and dramatic effects on heart rate, body temperature and physical activity, inducing a torpor-like state (19), but also more subtle effects on neurocognitive function (22). The physiological receptor(s) of TAMs has not been identified unambiguously, but despite their structural similarities, iodothyronines and TAMs appear to signal via different receptors. Initial studies suggested that the TAMs mediated their effects via the G-protein coupled trace amine receptor, TAR1 (19). However, the impressive hypothermic response to 3-T1AM administration is maintained in TAAR-1 knockout (23). Whether other members of the TAAR family or other plasma membrane receptors mediate the TAM response remains to be studied.

Studies in athyreotic patients provide evidence for extrathyroidal formation of 3-T1AM (24), but the pathways of TAM biosynthesis are still unknown (41). However, it has been shown that iodothyronamines are deiodinated by the different deiodinases (25), which may suggest a role in biosynthesis.

Interesting effects of other natural thyroid hormone derivatives have been described as well (26). Triac has significant thyromimetic activity and its affinity for the T3 receptor TRα1 is equal to that of T3 and for the TRβ receptor it is even higher than that of T3 (26). As a consequence, administration of Triac has successfully been used to suppress TSH secretion in patients with resistance to thyroid hormone due to mutations in TRβ (27). Interestingly, it was recently shown that the marine invertebrate Amphioxus expresses a TH receptor which is activated by Triac but not by T3 (28), as well as a non-selenoprotein that deiodinates Triac but not T3 (29). This may suggest that Triac is the primordial TH (29). A different natural TH derivative, 3,5-diiodo-L-thyronine (T2), has been shown to prevent adiposity and body weight gain when administered to rats receiving a high-fat diet (HFD) without the unfavorable side effects that are usually caused by T3 (30, 31).

However, the exact biological functions of these iodothyronine, iodothyronamine and iodothyroacetic acid metabolites remain to be established in future studies.

Cleavage of the ether bond connecting the inner and outer ring of iodothyronines represents a relatively minor pathway of thyroid hormone disposal (16) and will not be discussed here. In the following sections especially the biochemical aspects of the deiodination and conjugation pathways will be reviewed.

 

DEIODINATION

 

Three iodothyronine deiodinases have been identified, with distinct tissue distributions, catalytic specificities, physiological functions, and regulations (Fig. 2) (1-5). Whereas initial studies focused on the role of the deiodinases in maintaining normal serum T3 concentrations, the paradigm has evolved and it has now clearly been shown in different developmental and clinical conditions that these enzymes can locally modify TH bioactivity independent of serum TH concentrations. This is especially the case for D2 and D3 (see below).

 

Figure 2. Characteristics of the three types iodothyronine deiodinases

 

D1, D2 and D3 have been cloned in different species, including mammals, frog, chicken and fish. The deduced amino acid sequences of human D1, D2 and D3 are presented in Fig. 3. The deiodinases appear to be homologous proteins, consisting of 249-278 amino acids. A particular lipophilic sequence is present in the N terminal domain of all three deiodinases, which probably represents a membrane-spanning region.

 

The most remarkable feature of all three iodothyronine deiodinase is that they are selenoproteins, i.e. they contain a selenocysteine (Sec) residue in the center of the amino acid sequence (32). In all selenoproteins, Sec is encoded by a UGA triplet which is an opal stop codon because it usually signals termination of translation. However, if the 3' untranslated region (3'UTR) of the mRNA contains a particular stem loop structure, termed selenocysteine-insertion sequence (SECIS) element, the UGA codon specifies the insertion of Sec (33, 34). Interestingly, it was recently shown that the marine invertebrate Amphioxus expresses a non-selenoprotein that deiodinates Triac but not T3 (29).

 

 

Figure 3. Alignment of the amino acid sequences of human D1, D2 and D3

                U=selenocysteine (Sec)

 

Type I iodothyronine deiodinase (D1)

Biochemistry

D1 is expressed predominantly by liver parenchymal cells, kidney proximal tubular cells, and thyroid follicular cells. Most evidence points to the localization of D1 in the plasma membrane (35). D1 catalyzes the ORD and/or IRD of a variety of iodothyronine derivatives, although it is most effective in catalyzing the ORD of rT3, while the IRD of both T4 and T3 is strongly facilitated by sulfation of these iodothyronines (17). Therefore, although D1 is thought to be a major source of circulating T3, the enzyme shows particularly high activity towards TR-inactive metabolites such as rT3 and the different sulfo-conjugates. This suggests that D1 plays an important role in the recovery of iodide from inactive compounds for reutilization in thyroidal hormone synthesis (36). In the presence of dithiothreitol (DTT) as the cofactor, D1 displays high Km and Vmax values.

 

Studies of the topography of rat D1 have suggested that the major part of the protein is exposed on the cytoplasmic surface of the membrane (37). Older studies using detergent extracts of rat liver and kidney membranes have suggested that the native enzyme largely exists as a homodimer. This has been confirmed in a number of recent studies utilizing cells transfected with different D1 constructs (2, 38-40). These studies have also demonstrated that amino acids 148-163 constitute the dimerization domain of the D1 protein (DFLVIYIEEAHASDGW in human D1).

 

The D1 gene is located on human chromosome 1p32-33. It consists of four exons, with exon 1 coding for the 5’UTR and amino acids 1-112, exon 2 for amino acids 113-160, exon 3 for amino acids 161-227, and exon 4 for amino acids 228-249 and the 3’UTR, including the SECIS element. The Sec residue in D1 is essential for deiodinase activity since replacement of Sec by Cys results in a 100-fold decrease in catalytic activity, while substitution of Sec by Leu produces an enzymatically inactive protein (42). In addition, D1 is extremely sensitive to inactivation by iodoacetate due to carboxymethylation of a highly reactive residue, probably Sec, in the enzyme active center which is prevented in the presence of substrate (2, 15). Moreover, D1 activity is inhibited by very low concentrations (»10-8 M) of goldthioglucose (GTG), which is known to form very stable complexes with Sec residues, and this inhibition is also competitive with substrate (43). Therefore, Sec is probably the catalytic center of D1.

 

Two other observations have provided important clues about the possible catalytic mechanism of D1. Firstly, D1 shows ping-pong type reaction kinetics in catalyzing the deiodination of iodothyronines by DTT (2, 15), suggesting that reaction of iodothyronine substrate with D1 produces an enzyme intermediate, from which native enzyme is regenerated by reaction with thiol cofactor (DTT). Secondly, D1 is potently inhibited by PTU, and this inhibition is uncompetitive with substrate and competitive with cofactor, suggesting that PTU and cofactor react with the same enzyme intermediate. Thiouracil derivatives are particularly reactive towards protein sulfenyl iodide (SI) groups, and presumably even more reactive towards selenenyl iodide (SeI) groups, suggesting that such an intermediate is generated in the catalytic cycle of D1. Therefore, the selenolate (Se-) group of the native enzyme is thought to act as an acceptor of the iodonium (I+) ion which is substituted in the substrate by a proton, and the SeI intermediate thus generated is reduced back to native enzyme by thiols such as DTT or converted into a dead-end complex by PTU (Fig. 4).

 

 

Figure 4. Putative catalytic mechanism of D1 and inhibition by PTU, IAc and GTG.

 

 

 

Unlike the mammalian enzyme, D1 from tilapia was found to be insensitive to PTU inhibition (44). Like all other characterized deiodinases, tilapia D1 also contains a Sec residue in the corresponding position (44). However, two positions downstream from this Sec residue, tilapia D1 features a Pro residue, which is also the case in other fish species and in frog D1. In contrast, all known PTU-sensitive D1 enzymes have a Ser residue at this position. Remarkably, a Pro residue is also present at this position in all known D2 and D3 sequences, which are also PTU-insensitive. Substitution of Pro by Ser in tilapia D1 did not restore PTU sensitivity (44). However, substitution of Pro by Ser in frog D1 (4) as well as in human D2 and D3 not only made these enzymes susceptible to inhibition by PTU but also changed the kinetic mechanism of these enzymes (45). Therefore, in addition to the Sec residue the amino acid two positions downstream plays an important role in the catalytic mechanism of the deiodinases. The lack of PTU inhibition of the tilapia D1Pro>Ser mutant suggests that additional elements of the protein are important for effect of PTU.

Pathophysiology

D1 activity in liver and kidney is stimulated in hyperthyroidism and decreased in hypothyroidism, representing the regulation of D1 activity by T3 at the transcriptional level  (46). T3 response elements (TREs) have been identified in the upstream region of the D1 gene (47, 48). Studies in TR knockout mice have indicated that D1 expression in liver is primarily controlled by the TRb isoform (49). This agrees well with the colocalization of TRb and D1 in the pericentral zone of rat liver (50). In the thyroid, D1 expression is stimulated by T3, TSH and TSH receptor antibodies, where the effects of the latter are mediated by cAMP  (51, 52).

There is controversy about the contribution of D1 to peripheral T3 production. Different animal models have been studied which may provide a clue about this function of D1. Firstly, rats have been raised on a severely selenium-deficient diet, resulting in a dramatic reduction in liver and kidney D1 activity (53). These rats showed a significant decrease in serum T3 and increase in serum T4, compatible with an important role of D1 in peripheral T4 to T3 conversion. Other studies in rats have demonstrated that D2 and D3 activities in other tissues such as brain are much less affected directly by selenium deficiency (54).It should be noted that in mice lacking the plasma Se carrier selenoprotein P (SePP), thyroid hormone metabolism is preferentially maintained indicating that selenoenzymes have a priority in the organism with respect to selenium supply (55).

Findings in mice do not support an important function of liver and kidney D1 in peripheral T3 production as suggested by selenium deficiency in rats. C3H mice show a strongly reduced hepatic and renal D1 expression compared with other mouse strains (56-58). Yet, their serum T3 levels are comparable, although the C3H mice show some increase in serum T4 suggesting that an increased T4 production may compensate for the decreased T4 to T3 conversion. Serum rT3 levels are mildly elevated as well in C3H mice. In another mouse model, hepatic synthesis of selenoproteins, including D1, is disabled by inactivation of the Sec-specific tRNA  (59). This does not result in any change in circulating thyroid hormone concentrations. Finally, D1 knockout (D1KO) mice have been generated which do not express D1 in any tissue, including thyroid and kidneys (36). These D1KO mice also show normal serum T3 and TSH levels, but like the C3H mice they have elevated serum T4 and rT3 levels as well.

However, we should be careful to draw conclusions about the contribution of D1 to serum T3 homeostasis based on these knock-out mouse models, since even mice without any ORD (D1KO/D2KO mice) are able to maintain normal levels of serum T3 (60), pointing towards a major role played by the thyroid gland as well. Although these data in D1KO/D2KO mice suggest that D1 and D2 may not be essential for the maintenance of the serum T3 level, both enzymes do serve important roles in thyroid hormone homeostasis. Fecal excretion of endogenous iodothyronines was greatly increased in D1KO mice, pointing towards an important role in iodide conservation by serving as a scavenger enzyme in peripheral tissues and the thyroid (36). Similarly, despite normal serum T3 levels in D2KO mice, brain T3 levels as well as the expression of certain T3 responsive genes in the brain was reduced.

Many studies have addressed the question about the contribution of a diminished expression of hepatic D1 to the decrease in serum T3 in rats exposed to fasting or NTI. The results of these studies are confounded by the fact that D1 not only produces T3 from T4 but its expression is also stimulated by T3. In fact, D1 expression is a very sensitive indicator of the thyroid state of the liver (61).

So far, no patients with mutations in D1 have yet been identified. However, several candidate gene association studies have reported on significant associations of single nucleotide polymorphisms (SNPs) in D1 with reciprocal changes in serum T3 versus T4 and rT3 levels (62-64). Recently, a large genome wide association meta-analysis was conducted for serum FT4 levels, and a single nucleotide polymorphism in DIO1 was one of the five genome wide significant hits (65) strongly suggesting an important role for D1 in peripheral thyroid hormone metabolism in humans as well.

 

Type II iodothyronine deiodinase (D2)

 

Biochemistry

D2 is expressed primarily in the brain, the anterior pituitary gland and (rodent) brown adipose tissue (BAT) (1-5). D2 activity has also been shown in human thyroid (66-68) and skeletal muscle (69), while D2 mRNA is also expressed in human heart (70). Localization of D2 mRNA in rat brain by in situ hybridization has indicated that the enzyme is expressed in astrocytes, in particular in tanycytes lining the third ventricles (71). D2 activity is induced in cultured astrocytes by a variety of factors (73-74). Like the other deiodinases, D2 also forms functional homodimers (38, 39, 75). Regarding cellular localization, D2 is largely present in the endoplasmic reticulum (35).

D2 has only ORD activity, exhibiting low Km and Vmax values, and a slight preference for T4 over rT3 as the substrate. In contrast to D1, it does not catalyze the deiodination of sulfated iodothyronines. The amount of T3 in D2-expressing tissues is derived to a large extent from local conversion of T4 by this enzyme and to a minor extent from plasma T3. In general, D2 activity is increased in hypothyroidism and decreased in hyperthyroidism. Part of this negative control is explained by substrate-induced inactivation of the enzyme by T4 and rT3 (1-5). Reaction of these substrates with D2 induces the ubiquitination of the enzyme, which facilitates its degradation in the proteasomes. However, active D2 may also be recovered by de-ubiquitination of the modified enzyme. Thus, ubiquitination/de-ubiquitination is an important, dynamic mechanism in the regulation of D2 activity in many tissues, except hypothalamic D2 which is less ubiquitinated (72). For a more detailed discussion of this pathway, the reader is referred to excellent studies and reviews published in this area (38, 76-80). Furthermore, Dio2 is a cAMP-responsive gene and as a consequence the adrenergic/cAMP signaling pathway mediates the transcriptional control of D2 (81).In addition, D2 expression is also importantly regulated by ER stress reducing D2 activity by inhibition of  de novo synthesis of the D2 protein (82). Finally, presumably receptor-mediated inhibition of D2 activity by T3 has been demonstrated in pituitary tumor cells (83), and D2 mRNA levels in brain, pituitary and BAT are up-regulated in hypothyroid rats and down-regulated in hyperthyroid animals (84, 85).

The central Sec residue plays an important role in the catalysis and turnover of D2. Replacement of this Sec with Cys results in a 1000-fold increase in the Km value of the substrates T4 and rT3, and a 10-fold decrease in turnover number (86, 87). Substitution of Sec by Ala completely inactivates the enzyme. Also the mechanism of substrate-induced D2 degradation is strongly or completely impaired by replacement of Sec by Cys or Ala, respectively (88), suggesting that modification of this Sec residue during catalysis may be an essential step in the inactivation of the enzyme. Interestingly, mammalian and avian D2 also have a second Sec residue near the C-terminus which, however, is not important for catalytic activity (89).

The D2 gene is located on human chromosome 14q24.2-q24.3. It consists of 2 exons of 0.7 kb and 6.6 kb, seperated by a 7.4 kb intron (2). The SECIS element in the 3’UTR is separated by ~5 kb from the UGA triplet coding for the catalytic Sec residue, resulting in a poor translation efficiency of the D2 mRNA (90). This is even further hampered by the presence of multiple short open reading frames in the 5’UTR of human D2 mRNA (90).

 

Pathophysiology

D2 is expressed in human thyroid but not in rat thyroid. Both D2 mRNA and D2 activity in human thyroid are greatly stimulated by TSH and TSH receptor antibodies circulating in patients with Graves’ disease (66, 67). The expression of D2 in human thyroid has been associated with functional TTF-1 binding sites in the 5’ flanking region of the human D2 gene which are lacking in the 5’ flanking region of the rat D2 gene (91). The stimulatory effects of TSH and TSH receptor antibodies on D2 expression in human thyroid are mediated by cAMP, which has been associated with the presence of a cAMP response element (CRE) in the 5’ flanking region of the D2 gene (81, 92). Interestingly, follicular thyroid carcinoma may express high levels of D2, and in case of a large (metastatic) tumor mass this may results in strongly elevated serum T3 levels (93-95).

D2 knockout (D2KO) mice have been generated, showing modest phenotypic changes (96). The homozygous D2KO mice have increased serum T4 and increased TSH levels, but normal levels of T3. The combination of increased serum TSH and T4 is in agreement with an important role of D2 in the negative feedback of T4 at the hypothalamus and pituitary level. However, the normal serum T3 suggests that D2 is not essential for maintaining normal serum T3 levels. However, as mentioned above, even D1KO/D2KO mice are able to maintain normal levels of serum T3 (60), pointing towards a major role for the thyroid gland in serum T3 production as well. In skeletal muscle, D2 levels are higher in slow-twitch than fast-twitch mouse skeletal muscle and are increased in hypothyroidism (97). Specific deletion of D2 in skeletal muscle does not have large effects on thyroid hormone signaling or functional outcomes (98,99).

In contrast to the marked decrease in hepatic and renal (but not thyroidal) D1 activities, the unexpectedly small effects of Se deficiency on tissue D2 and D3 activities in rats, despite that they all appear to be Sec-containing proteins, may be explained by findings that the selenium state of different tissues varies greatly in Se-deficient animals. In addition, the efficiency of the SECIS element to complex with protein factors, such as SBP2, necessary for the read-through of the UGA codon may vary between different seleno­proteins. This could result in the preferred incorporation of Sec into some seleno­proteins, e.g. deiodinases, over others, e.g. glutathione peroxidase  (33).

 

Despite normal serum T3 levels in D2KO mice, brain T3 levels as well as the expression of certain T3 responsive genes in the brain is reduced, again pointing towards the crucial role of D2 in maintaining local T3 concentrations (96, 105). Several other studies point towards a crucial role for D2 (and D3, see below) in regulating local T3 concentrations, and as a consequence it is know well accepted that these deiodinases can regulate thyroid hormone action at the cellular level during development and tissue stress relatively independent of serum T4 and T3 concentrations (3).

One of the clearest examples of the role of D2 in development is its role in the inner ear. A sharp increase in D2 activity occurs in mouse cochlea at postnatal days 6-8, which is required for normal cochlear development (100). As a consequence, D2KO mice are deaf underlining the importance of D2 in producing local T3 in the cochlea during a critical period of its development (101). Another example of the important role of D2 in development is the observation that D2KO mice have an impaired embryonic BAT development, and as a consequence a permanent thermogenic defect (102, 103). D2KO mice exhibit an impaired thermogenesis in BAT, leading to hypothermia during cold exposure and a greater susceptibility to diet-induced obesity at thermoneutrality (104).

D2 is also essential for maintaining normal local concentrations of T3 in different physiological and pathophysiological situations. D2 has an important role in pituitary and hypothalamic feedback (96). It also plays an essential role for normal myogenesis (106) and in the optimization of bone strength and mineralization (107). Adult D2KO mice have a 50% reduction in bone formation and a generalized increase in skeletal mineralization resulting from a local deficiency of T3 in osteoblasts (101).

D2 is also required for the regeneration of skeletal muscle after injury (106), since regeneration after injury is markedly delayed in D2KO mice. The increase in muscle D2 is mediated via FoxO3, thereby locally increasing intracellular T3 concentrations. Muscle D2 expression during critical illness is differentially regulated, probably related to differences in the inflammatory response and type of pathology (108). In humans, skeletal muscle D2 mRNA expression is modulated by fasting and insulin, but not by hypothyroidism (109). Also in lung tissue, D2 activity increases upon injury. In a mouse model of ventilator-induced lung injury (VILI), lung D2 activity increased (110). D2KO mice had a greater susceptibility to VILI than WT mice, demonstrated by poorer alveoli integrity and quantified by lung chemokine and cytokine induction. Evidence accumulates that D2 is induced during inflammation (e.g. in macrophages) (198, 199).  The neonatal D2 peak in the liver appears relevant for susceptibility to diet-induced steatosis and obesity as shown in hepatocyte-specific D2KO (196).

No patients have been identified with mutations in D2. However, patients with homozygous or compound heterozygous mutations in the SECIS-binding protein SBP2, which is crucial for the synthesis of selenoproteins (111) have abnormal serum thyroid hormone levels: high (F)T4 and rT3, low T3, and somewhat elevated TSH levels. This resembles the changes in thyroid parameters in D2KO mice, although in patients with SBP2 mutations also the expression of functional D1 and D3 is probably affected. As SBP2 deficiency affects many selenoproteins, these patients have a multisystem disorder including growth delay in childhood, hearing loss, enhanced skin sensitivity and infertility (111,112).

Whether polymorphisms in D2 are associated with significant changes in serum thyroid hormone levels or other outcomes such as insulin resistance or osteoarthritis are controversial (62, 113-117, 200). Also uncommon D2 variants are not related to serum thyroid hormone levels (201).  The Thr92Ala polymorphism has been linked with local changes in a specific transcriptional fingerprint, although the relevance needs to be further studied (202).

Type III iodothyronine deiodinase (D3)

Biochemistry

D3 activity has been detected in a variety of tissues, i.e. brain, skin, liver, intestine, placenta, and the pregnant rat uterus (1-5, 118-120). D3 expression is usually much higher in fetal than in adult tissues. D3 activity is also highly expressed in certain tumors, including hepatocarcinomas, hemangiomas and basal cell carcinomas (121-124, 203) Because of its expression in fetal tissues and tumors, D3 has been named an oncofetal protein. The enzyme appears to be located in the plasma membrane in the form of a homodimer (38, 125, 126). D3 has only IRD activity, catalyzing the inactivation of T4 and T3 with intermediate Km and Vmax values (Fig. 2).

 

The expression of D3 in placenta, pregnant uterus, embryonic and fetal tissues may protect developing organs against undue exposure to active thyroid hormone. Also in adult subjects, D3 appears to be an important site for clearance of plasma T3 and production of plasma rT3. In brain and skin, but not in placenta, D3 activity is increased in hyperthyroidism and decreased in hypothyroidism, which in brain is associated with parallel changes in D3 mRNA levels (127).

 

The D3 gene is located on human chromosome 14q32 and consists of a single exon. In all species, D3 is a selenoprotein homologous with the amino acid sequences of D1 and D3, including the essential Sec residue positioned in a strongly conserved region (Fig. 2). It has been shown that D3 expression is predominantly regulated by TRα1 (128), and studies in TRα1-/- mice have demonstrated a reduced clearance rate of TH due to an impaired regulation of D3 (129).

 

The presence of Sec in a strongly conserved region of the proteins strongly suggests the same mechanism of deiodination for the different deiodinases. This seems to be contradicted by the widely different susceptibilities of D1 versus D2 and D3 to the different mechanism-based inhibitors PTU, IAc and GTG (Fig. 2). It also seems to be in conflict with previous findings that, in contrast to the ping-pong kinetics of D1, the other two enzymes appear to follow sequential-type kinetics, suggesting the formation of a ternary enzyme-substrate-cofactor complex during catalysis. The differences in enzyme kinetics and PTU inhibition between the deiodinases are determined by the presence of Ser (D1) or Pro (D2,D3) two positions downstream of  Sec, which may somehow influence the reactivity of the catalytic Sec residue (see above). The crystal structure of the catalytic part of D3 suggests a close similarity to 2-Cys peroxiredoxin(s) (Prx) with an  important resolving role for Cys239 by forming a selenenyl-sulfide with Sec170 (204).

 

Pathophysiology

D3 plays a very important role in the regulation of local and systemic thyroid hormone bioactivity (1, 123). It has been shown that region-specific expression of D3 in fetal and adult human brain is negatively associated with local tissue T3 levels (130, 131). High expression of D3 in vascular tumors may result in subclinical or even severe hypothyroidism in patients with such tumors, which condition has been termed consumptive hypothyroidism (122, 123, 132). Induction of D3 expression has also been demonstrated in liver and skeletal muscle biopsies from patients who died after severe illness, and D3 activities were correleated to both local and serum rT3 concentrations in these severely sick patients (133-135). Therefore, tissue and circulating iodothyronine levels are regulated not only by changes in the T3-producing deiodinases D1 and D2 but also importantly by reciprocal changes in the T3-degrading deiodinase D3.

 

D3 knockout (D3KO) mice have been generated, showing remarkable neonatal mortality and growth retardation, althought the severity of the phenotype depends on the genetic background (136-138). In addition, they show largely abnormal thyroid hormone levels, dependent on the age of the animals. Compared with wild-type mice, serum T4 is very low in D3KO mice at all ages, T3 is higher in neonatal mice but much lower in older D3KO mice, while TSH varies between very low in younger to low in older knockout mice. This picture represents a state of central hypothyroidism, suggesting that the setpoint of the hypothalamus-pituitary-thyroid axis is strongly affected by inactivation of D3, which could be due to overexposure of tissues (e.g. the developing hypothalamus) to T3. This is reminiscent of the reports of congenital central hypothyroidism in newborns from mothers who were hyperthyroid during pregnancy (139).

 

Heterozygous D3KO mice show either almost normal or strongly decreased D3 expression, depending on whether the defective allele is inherited from the mother or the father, respectively, indicating paternal imprinting of the DIO3 gene (136, 205). However, D3 expressed from the maternal D3 allele is important in pancreatic islets to maintain normal glucose homeostasis (206). The DIO3 gene is located in an imprinted region on human chromosome 14 or mouse chromosome 12 which is about 1 Mb in size and comprises the paternally expressed genes DLK1 (delta-like 1) and DIO3, and a large number of in particular maternally expressed non-coding genes. Both Dlk1 and Dio3 expression are elevated in cultured brown pre-adipocytes and down-regulated during differentiation, suggesting that imprinting might control the dosage of these genes to regulate thermogenesis (140). Interestingly, transgenic animals with partial loss of imprinting of this locus show significant lethality in the third postnatal week, associated with developmental delay and failure to maintain UCP1 expression in BAT (141). This defect is the combined result of prolonged elevated expression of Dlk1, leading to a failure in BAT differentiation and subsequent reduced expression of β-adrenergic receptors, and hypothyroidism due to dysregulation of D3.

 

 

The important role of hepatic D3 in the regulation of circulating thyroid hormone during development has been investigated in detail in the embryonic chicken (142, 143). These studies have demonstrated that during the last (third) week of incubation there is a gradual increase in plasma T4 levels paralleled by a steady increase in hepatic D1 activity although hepatic D1 mRNA levels do not change much. D3 mRNA and D3 activity show a parallel increase to maximum levels at day 17 of embryonic development, followed by a steep decrease in both parameters in particular immediately before hatching. This is associated with an equally steep increase in plasma T3, strongly suggesting that the latter is importantly and negatively regulated by hepatic D3 activity (142, 143).

A study of the ontogeny of hepatic D1 and D3 during human development has indicated similar profiles of deiodinase expression, with substantial and relatively constant D1 activities from mid-gestation onwards, and high D3 activities at mid-gestation declining to very low levels around term (144). Since in rat liver D1 is not expressed until the last days of gestation, while hepatic D3 expression is low at all stages of rat development (118), these results indicate that the embryonic chicken is a better model than the fetal rat for the regulation of hepatic deiodinases during human development. Injection of the chicken embryo with growth hormone or glucocorticoids induces an acute down-regulation of hepatic D3 mRNA levels and D3 activities, suggesting that the D3 mRNA in the embryonic chicken has a very short half-life, and that transcription of the D3 gene is acutely blocked by these treatments (142). If D3 expression in the fetal human liver is also rapidly down-regulated by GH and glucocorticoids remains to be determined. It is likely that the high D3 activities expressed in the fetal liver, in addition to the high D3 activities in the placenta (145, 146) and perhaps the uterus (119), plays an important role in the regulation of fetal circulating T3 levels and protect the fetus against early T3 exposure.

 

In recent years, several studies have addressed the role of D3 in regulating local T3 concentrations. It is now well accepted that D3 plays a crucial role in regulating thyroid hormone action at the cellular level during development, relatively independent of serum T4 and T3 concentrations. During development, D3 is expressed in the immature cochlea before D2 (147). Like D2KO mice, D3KO mice display auditory deficits as well. However, in contrast to the retarded cochlear development in D2KO mice, D3KO mice display an accelerated cochlear differentiation due to premature stimulation of TRb. The additional deletion of TRb converts the accelerated cochlear phenotype in D3KO mice to one of delayed differentiation (147), indicating a protective role for D3 in hearing development. This clearly illustrates how different tissues can auto-regulate their developmental response to thyroid hormone through both D2 and D2. D3 also plays a crucial role in cerebellar development, since D3KO mice display abnormally accelerated cerebellar differentiation and locomotor behavioral defects, suggesting that D3 protects cerebellar tissues from inappropriate, premature stimulation by thyroid hormone (148, 207). This cerebellar phenotype results specifically from inappropriate stimulation of the TRa1 receptor isoform, since the additional deletion of TRa1 reversed the cerebellar phenotype. Also, additional deletion of MCT8 in the D3KO mice ameliorates the phenotype indicating the relevance of MCT8 for intracellular T3 levels (208). Similarly, D3 protects cones to unlimited T3 exposure in the immature mouse retina. As a consequence, approximately 80% of cones are lost through neonatal cell death in D3KO mice (149). Similar results were obtained in zebrafish (209). D3 appears also a critical factor in testis development via influencing local thyroid hormone bioavailability (210).  Furthermore, protection against untimed T3 exposure by D3 in pancreatic β-cells during development is essential for normal islet function and glucose homeostasis (150). As a consequence, D3KO mice have impaired insulin secretion in response to glucose stimulation. In contrast to most tissues, D3 expression remains throughout adulthood in human and mouse β-cells. However, whether dysregulation of Dio3 might play a role in different states of impaired insulin secretion remains to be explored in future studies (150). In addition, less fat tissue is seen in D3KO mice, which is mediated through in the leptin-melanocortin system (211).

In addition to its crucial role during development, D3 activity is also important in regulating thyroid hormone action at the cellular level in different pathophysiological conditions. Induction of D3 expression has been documented in the hypertrophic or failing heart resulting from pressure overload or myocardial infarction (151-153). Hypoxia-inducible factor 1a (HIF-1a) induces local thyroid hormone inactivation by inducing D3 during hypoxia (152), suggesting a mechanism of down-regulating metabolism during ischemia. In neuronal hypoxia, translocation of D3 to the nucleus is mediated by Hsp-40, thereby facilitating local inactivation of thyroid hormone and reducing ischemia-induced hypoxic brain damage (154). Heterozygous D3KO mice constitute a model of cardiac D3 inactivation in an otherwise systemically euthyroid animal (155). These mice have normal hearts but later develop restrictive cardiomyopathy due to cardiac-specific increase in thyroid hormone signaling. In addition, heterozygous D3KO mice are more vulnerable to isoproterenol, further worsening the restrictive cardiomyopathy and leading to congestive heart failure and increased mortality (155). D3 activity is also induced in liver and muscle of critically ill patients (133-135). See (156) for an excellent overview of the changes in local thyroid hormone metabolism during illness and inflammation. Interestingly, in a mouse model of turpentine-induced tissue inflammation, high D3 expression in invading granulocytes has also been reported (157, 158). Recent studies also documented D3 in human neutrophils (212). Furthermore, D1 decreases and D3 increases are seen in livers of premature and normal aging mice, hinting that changes in deiodinases are mediated via DNA damage and might contribute to the beneficial survival response (213).

Several recent studies have demonstrated that local regulation of thyroid hormone action also plays a crucial role in repair mechanisms, for example D3 in liver (159) and brain (160), and D2 and D3 in muscle (106, 214). The reciprocal changes in D2 and D3 are shown in elegant studies demonstrating that D2 is induced to allow proper differentiation after muscle injury, while D3 induction in proliferating muscle cells protects against excessive local thyroid hormone concentrations, preventing apoptosis (106, 214). Furthermore, D2 and D3 activities are regulated by a variety of growth factors and morphogens, which are important mediators of tissue injury repair (161).  After a large hepatectomy, ‘stem-like’ cells switch from a quiescent state to a proliferative state. During these processes, many fetal genes are reactivated (162). Among them, D3 activity was increased 10-fold and D3 mRNA expression was increased 3-fold 20 h after partial hepatectomy in rats. No significant effects on D1 and D2 activities or mRNA expression were found after partial hepatectomy in mice (159). This leads to the concept that a coordinated regulation of thyroid hormone action is essential in the control of the tight balance between proliferation and differentiation in the regeneration processes. Induction of D3 expression in the early phases of regeneration may therefore very well correlate with a requirement of increased cellular proliferation in these circumstances (3, 161).

 

The balance between proliferation and differentiation is disturbed in cancer, and D3 is turned on in several malignant cell lines and human cancers (3, 163). D3 activity in these cancers can be very high and may even lead to so-called consumptive hypothyroidism (123, 132, 203) . In basal cell carcinomas, as well as in primary proliferating keratinocytes, Sonic hedgehog  (Shh) increases the expression of D3, acting via a conserved Gli2 binding site on the human Dio3 promoter (121, 215). This suggests that Shh may induce local down-regulation of thyroid hormone activity. Interestingly, knockdown of D3 caused a 5-fold reduction in the growth of basal cell carcinoma xenografts in nude mice (121), suggesting that D3 up-regulation provides an advantage for proliferating tumor cells. This appears to be mediated by miR21 that reduces the tumor suppressor gene GRHL3 which in turn increases D3 expression (216). Interestingly, a recent study in papillary thyroid carcinoma demonstrated an  association between increased levels of D3 activity and advanced disease (164). However, since only a few tumors over-express D3, D3 expression seems not be a necessary step in tumorigenesis.

Sulfation

Iodothyronine sulfotransferases

Sulfotransferases represent a family of enzymes with a monomer molecular weight of »34 kDa, located in the cytoplasmic fraction of different tissues, in particular liver, kidney, intestine and brain (165). They catalyze the transfer of sulfate from 3'-phosphoadenosine-5'-phosphosulfate (PAPS) to usually a hydroxyl group of the substrate (165). On the basis of substrate specificity and amino acid sequence homology, mainly two sulfotransferase families have been recognized in human tissues, i.e. the phenol sulfo-transferases (SULT1 family), including estrogen sulfotransferase, and the hydroxysteroid sulfotransferases (SULT2 family) (165). Different phenol sulfotransferases have been identified with significant activity towards iodothyronines. These include human SULT1A1, 1A2, 1A3, 1B1 and 1C2 (Table 1) (166-174). These studies have indicated a large substrate preference of the recombinant enzymes as well as the native enzymes in human liver and kidney for 3,3'T2, the sulfation of which is catalyzed orders of magnitude faster than that of T3 or rT3, while sulfation of T4 is hardly detectable (168).

 

Surprisingly, it has also been demonstrated that human estrogen sulfo-transferase (SULT1E1) is an important isoenzyme for sulfation of thyroid hormone. Although human SULT1E1 shows much higher affinities for estrogens (Km »nM) than for iodothyronines (Km »μM), it is about as efficient as other isoenzymes in sulfating 3,3'T2 and T3, and much more efficient in sulfating rT3 and T4 (169). Human tissues known to express SULT1E1 include liver, uterus, and mammary gland (175). In particular the enzyme expressed in the endometrium may be a significant source for the high levels of iodothyronine sulfates in human fetal plasma (see below). Recently, different human SULTs have also been shown to catalyze the sulfation of iodothyronamines (Table 1) (172).

 

Deiodination of iodothyronine sulfates

Although D1 is capable of converting T4 with similar efficiency by ORD to T3 and by IRD to rT3, this is changed dramatically after sulfate conjugation, i.e. IRD of T4S by rat D1 is accelerated »200-fold, whereas ORD of T4S becomes undetectable (Fig. 5) (17). IRD of T3 by rat and human D1 is also markedly stimulated (»40-fold) by sulfation (Fig. 5)(17). As mentioned before, rT3 is by far the preferred D1 substrate; its ORD is not influenced by sulfation, suggesting that the catalytic efficiency of D1 is already optimal with nonsulfated rT3 (17). While sulfation inhibits ORD of T4 and is without effect on ORD of rT3, it markedly stimulates ORD of 3,3’-T2 (Fig. 5). Thus, sulfation facilitates the IRD of T4 and T3, while it either inhibits (T4), does not affect (rT3) or markedly stimulates (3,3’T2) the ORD of other substrates (17).

 

 

Figure 5. Efficiency of deiodination of iodothyronines and their sulfates by rat liver D1.

   *) Vmax (pmol/min/mg protein)/Km (µM) ratio.

 

 

The mechanism by which sulfation stimulates especially the IRD of different substrates remains unclear. In some cases sulfation primarily effects an increase in Vmax, while in others there is a predominant decrease in apparent Km value. The facilitated deiodination of sulfated iodothyronines by rat liver D1 may be due to interaction of the negatively charged sulfate group with protonated residues in the active center of this basic protein. The effect of sulfation on deiodination of iodothyronines is both conjugation type and deiodinase type-specific since D1 does not catalyze the deiodination of T3 glucuronide, while D2 and D3 do not accept T4S and/or T3S as substrates.

 

Importance of thyroid hormone sulfation

Serum concentrations of T4S, T3S, rT3S and 3,3’T2S are low in normal human subjects but they are high in fetal and cord blood, in patients with NTI, and in patients treated with the D1 inhibitor (16, 176). The serum T3S/T3 ratio is also increased in hypothyroid patients (177, 178). High iodothyronine sulfate levels have also been detected in human fetal and neonatal serum and amniotic fluid (16). The high serum iodothyronine sulfate levels during NTI, hypothyroidism and fetal development have been ascribed to a low peripheral D1 activity in these conditions (17, 18). These results are in accordance with studies in rats, showing marked increases in the serum concentration and biliary excretion of iodothyronine sulfates when hepatic and renal D1 activities are decreased by D1 inhibitors or selenium deficiency (17). These changes are not caused by an increased sulfation of iodothyronines but by a decreased clearance of the sulfated iodothyronines by D1.

Thus, sulfation is a primary step leading to the irreversible degradation of T4 and T3 by D1. However, if D1 activity is low, inactivation of thyroid hormone by sulfation is reversible due to expression of sulfatases in different tissues and by intestinal bacteria (179). It has been speculated that especially in the fetus T3S has an important function as a reservoir from which active T3 may be released in a tissue-specific and time-dependent manner(17, 180).

Wu and coworkers have demonstrated the presence of a 3,3’T2S cross-reacting substance, termed compound W, in the serum and urine of pregnant women  (16, 181). Interestingly, compound W is derived from the fetus and its concentration in maternal serum may reflect fetal thyroid state (16, 181). The structure of compound W remains to be identified.

 

Glucuronidation

Like sulfation, glucuronidation is a phase II metabolic reaction that increases the water-solubility of endogenous and exogenous compounds to increase their biliary or urinary excretion. Glucuronidation is catalyzed by UDP-glucuronyltransferases (UGTs) that utilize UDP-glucuronic acid (UDPGA) as cofactor. UGTs are localized in the endoplasmic reticulum of predominantly liver, kidney and intestine. Most UGTs are members of the UGT1A and UGT2B families (182).

Iodothyronines are also metabolized by glucuronidation, although this appears more important in rodents than in humans (183). Especially in rodents, metabolism of thyroid hormone is accelerated through induction of T4-glucuronidating UGTs by different classes of compounds, including barbiturates, fibrates and PCBs (184-186). This may result in a hypothyroid state as the thyroid gland is not capable of compensating for the increased hormone loss. In humans, thyroid function may be affected by induction of T4 glucuronidation by anti-epileptics, but development of overt hypothyroidism is rare (187).

Glucuronidation of T4 and T3 is catalyzed by different members of the UGT1A family (Table 2) (188-191). Usually, this involves the glucuronidation of the hydroxyl group, but human UGT1A3 also catalyzes the glucuronidation of the side-chain carboxyl group, with formation of so-called acyl glucuronides (189). Interestingly, Tetrac and Triac are much more rapidly glucuronidated in human liver than T4 and T3, and this occurs predominantly by acyl glucuronidation (192). Acyl glucuronides are reactive compounds that may form covalent complexes with proteins. It is unknown if this is a significant route for the formation of covalent iodothyronine-protein complexes.

Integrated physiological role of thyroid hormone metabolism

Since most actions of thyroid hormone are initiated by binding of T3 to its nuclear receptors, it is important to consider the role of the processes discussed above in the regulation of nuclear T3. There are two sources of intracellular T3, i.e. T3 derived from plasma T3, and T3 produced locally from T4, and the degree to which they contribute to the occupied receptors varies among the different tissues in different physiological and pathophysiological states (1, 3, 5, 76, 98). The liver and kidneys are typical of most tissues in the body in which most of the T3 specifically bound to the T3 receptor is derived directly from plasma. In cerebral cortex, BAT, and anterior pituitary there is a substantial contribution to nuclear T3 from locally produced T3. Local T3 production may be an autocrine process, where T3 is produced in the same cells where it acts, or a paracrine mechanism, where T3 production and action take place in neighboring cells. The latter appears very important for T3 action in the brain, where neurons are the primary target cells for T3 produced by D2 expressed in nearby astrocytes (193-195).

 

D3 plays an additional important role in maintaining intracellular T3 concentrations in these tissues by catalyzing the degradation of T3 in case of excess or by diverting the metabolism of T4 to rT3. Indeed, the adaptations of deiodinase activities in response to changes in thyroid state are thought to serve the purpose of keeping intracellular T3 in the brain constant. Thus, when T4 supply is decreased in hypothyroidism, both D1 and D3 activities are down-regulated, so that relatively more T4 is available for conversion to T3 by D2 in the brain, the activity of which is up-regulated. Opposite changes occur in hyperthyroidism. These adaptations are not only important for the optimal function of the brain in adult life, they are also essential for the development of the brain which is critically dependent on thyroid hormone. Although the adaptations in deiodinase activities during hypo- or hyper­thyroidism go a long way in securing T3 availability in the brain, in severe iodine deficiency they may not fully compensate for the extreme decrease in T4 supply. This may result in severe impairment of neurological development in the child even when plasma T3 levels in the mother are sufficient to maintain a euthyroid state.

 

The critical role of deiodination in regulating local thyroid hormone action is clearly illustrated by the developing cochlea, where D3 is expressed before the onset of D2 activity (101, 147), preventing too much or too little hormonal stimulation at inappropriate stages in development. At immature stages, D3 limits stimulation by T3. Postnatally, a double switch occurs with a decline in D3 and an increase D2, resulting in a local T3 surge which is independent of serum T3 levels and triggers the onset of auditory function. A similar double switch, preventing premature T3 stimulation, occurs in the developing cerebellum (148), and D3 expression has also been shown to be crucial for normal retinal (149) and pancreatic b-cell development (150). Similarly, local thyroid hormone activation by D2 has been shown to be essential for normal BAT development (102) and myogenesis (106) as well.

Deiodinases are not only essential in controlling local thyroid hormone action during development, but also for normal function of adult tissues such as hypothalamus, pituitary, bone, and brown adipose tissue (96, 102, 107). Finally, deiodination is also important in regulating thyroid hormone bioactivity in different pathophysiological conditions, such as hypoxia, myocardial infarction, neuronal ischemia, critical illness, tissue injury, regeneration, and cancer (106, 121, 134, 152-154, 156, 157, 159). D2KO mice are more vulnerable to ventilator induced lung injury (110), whereas heterozygous D3KO mice are more vulnerable to a chemically induced worsening of restrictive cardiomyopathy, leading to congestive heart failure and increased mortality (155). The high expression of D3 in regenerating liver tissue and certain tumors and the crucial role of D2 and D3 in muscle regeneration (3, 106, 123, 159, 197, 214) suggest that coordinated regulation of thyroid hormone action is essential in the control of the tight balance between proliferation and differentiation in the regeneration processes, and that high expression of D3 may also be an advantage in proliferating tumor cells (98, 101, 147). The joint coordination between the different deiodinases is seen in mice lacking all deiodinases (D1/D2/D3 KO) versus individual deiodinase KO mice. D1/D2/D3 KO mice are viable and some features resulting from deficiency of either of the deiodinases is mitigated by the simultaneous lack of all deiodinases (217).

REFERENCES

 

  1. Bianco AC, Kim BW. Deiodinases: implications of the local control of thyroid hormone action. J Clin Invest. 2006 Oct;116(10):2571-9.
  2. 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 Feb;23(1):38-89.
  3. Gereben B, Zavacki AM, Ribich S, Kim BW, Huang SA, Simonides WS, et al. Cellular and molecular basis of deiodinase-regulated thyroid hormone signaling. Endocr Rev. 2008 Dec;29(7):898-938.
  4. Gereben B, McAninch EA, Ribeiro MO, Bianco AC. Scope and limitations of iodothyronine deiodinases in hypothyroidism. Nat Rev Endocrinol. 2015 Nov;11(11):642-52.
  5. Larsen PR, Zavacki AM. Role of the Iodothyronine Deiodinases in the Physiology and Pathophysiology of Thyroid Hormone Action. Eur Thyroid J. 2012.
  6. Escobar-Morreale HF, del Rey FE, Obregon MJ, de Escobar GM. Only the combined treatment with thyroxine and triiodothyronine ensures euthyroidism in all tissues of the thyroidectomized rat. Endocrinology. 1996 Jun;137(6):2490-502.
  7. Escobar-Morreale HF, Obregon MJ, Escobar del Rey F, Morreale de Escobar G. Tissue-specific patterns of changes in 3,5,3'-triiodo-L-thyronine concentrations in thyroidectomized rats infused with increasing doses of the hormone. Which are the regulatory mechanisms? Biochimie. 1999 May;81(5):453-62.
  8. Bunevicius R, Jakubonien N, Jurkevicius R, Cernicat J, Lasas L, Prange AJ, Jr. Thyroxine vs thyroxine plus triiodothyronine in treatment of hypothyroidism after thyroidectomy for Graves' disease. Endocrine. 2002 Jul;18(2):129-33.
  9. 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 Jul;91(7):2592-9.
  10. Escobar-Morreale HF, Botella-Carretero JI, Gomez-Bueno M, Galan JM, Barrios V, Sancho J. Thyroid hormone replacement therapy in primary hypothyroidism: a randomized trial comparing L-thyroxine plus liothyronine with L-thyroxine alone. Ann Intern Med. 2005 Mar 15;142(6):412-24.
  11. Hennemann G, Docter R, Visser TJ, Postema PT, Krenning EP. Thyroxine plus low-dose, slow-release triiodothyronine replacement in hypothyroidism: proof of principle. Thyroid. 2004 Apr;14(4):271-5.
  12. Appelhof BC, Peeters RP, Wiersinga WM, Visser TJ, Wekking EM, Huyser J, et al. Polymorphisms 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 Nov;90(11):6296-9.
  13. Panicker V, Saravanan P, Vaidya B, Evans J, Hattersley AT, Frayling TM, et al. Common variation in the DIO2 gene predicts baseline psychological well-being and response to combination thyroxine plus triiodothyronine therapy in hypothyroid patients. J Clin Endocrinol Metab. 2009 May;94(5):1623-9.
  14. van der Deure WM, Appelhof BC, Peeters RP, Wiersinga WM, Wekking EM, Huyser J, et al. Polymorphisms in the brain-specific thyroid hormone transporter OATP1C1 are associated with fatigue and depression in hypothyroid patients. Clin Endocrinol (Oxf). 2008 Nov;69(5):804-11.
  15. Visser TJ. Pathways of thyroid hormone metabolism. Acta Med Austriaca. 1996;23(1-2):10-6.
  16. Wu SY, Green WL, Huang WS, Hays MT, Chopra IJ. Alternate pathways of thyroid hormone metabolism. Thyroid. 2005 Aug;15(8):943-58.
  17. Visser TJ. Role of sulfation in thyroid hormone metabolism. Chem Biol Interact. 1994 Jun;92(1-3):293-303.
  18. Peeters RP, Kester MH, Wouters PJ, Kaptein E, van Toor H, Visser TJ, et al. Increased thyroxine sulfate levels in critically ill patients as a result of a decreased hepatic type I deiodinase activity. J Clin Endocrinol Metab. 2005 Dec;90(12):6460-5.
  19. Scanlan TS, Suchland KL, Hart ME, Chiellini G, Huang Y, Kruzich PJ, et al. 3-Iodothyronamine is an endogenous and rapid-acting derivative of thyroid hormone. Nat Med. 2004 Jun;10(6):638-42.
  20. Piehl S, Hoefig CS, Scanlan TS, Kohrle J. Thyronamines--past, present, and future. Endocr Rev. 2011 Feb;32(1):64-80.
  21. DeBarber AE, Geraci T, Colasurdo VP, Hackenmueller SA, Scanlan TS. Validation of a liquid chromatography-tandem mass spectrometry method to enable quantification of 3-iodothyronamine from serum. J Chromatogr A. 2008 Nov 7;1210(1):55-9.
  22. Manni ME, De Siena G, Saba A, Marchini M, Landucci E, Gerace E, et al. Pharmacological effects of 3-iodothyronamine (T1AM) in mice include facilitation of memory acquisition and retention and reduction of pain threshold. Br J Pharmacol. 2012 Aug 13.
  23. Panas HN, Lynch LJ, Vallender EJ, Xie Z, Chen GL, Lynn SK, et al. Normal thermoregulatory responses to 3-iodothyronamine, trace amines and amphetamine-like psychostimulants in trace amine associated receptor 1 knockout mice. J Neurosci Res. 2010 Jul;88(9):1962-9.
  24. Hoefig CS, Kohrle J, Brabant G, Dixit K, Yap B, Strasburger CJ, et al. Evidence for extrathyroidal formation of 3-iodothyronamine in humans as provided by a novel monoclonal antibody-based chemiluminescent serum immunoassay. J Clin Endocrinol Metab. 2011 Jun;96(6):1864-72.
  25. Piehl S, Heberer T, Balizs G, Scanlan TS, Smits R, Koksch B, et al. Thyronamines are isozyme-specific substrates of deiodinases. Endocrinology. 2008 Jun;149(6):3037-45.
  26. Moreno M, de Lange P, Lombardi A, Silvestri E, Lanni A, Goglia F. Metabolic effects of thyroid hormone derivatives. Thyroid. 2008 Feb;18(2):239-53.
  27. Radetti G, Persani L, Molinaro G, Mannavola D, Cortelazzi D, Chatterjee VK, et al. Clinical and hormonal outcome after two years of triiodothyroacetic acid treatment in a child with thyroid hormone resistance. Thyroid. 1997 Oct;7(5):775-8.
  28. Paris M, Escriva H, Schubert M, Brunet F, Brtko J, Ciesielski F, et al. Amphioxus postembryonic development reveals the homology of chordate metamorphosis. Curr Biol. 2008 Jun 3;18(11):825-30.
  29. Klootwijk W, Friesema EC, Visser TJ. A nonselenoprotein from amphioxus deiodinates triac but not T3: is triac the primordial bioactive thyroid hormone? Endocrinology. 2011 Aug;152(8):3259-67.
  30. de Lange P, Cioffi F, Senese R, Moreno M, Lombardi A, Silvestri E, et al. Nonthyrotoxic prevention of diet-induced insulin resistance by 3,5-diiodo-L-thyronine in rats. Diabetes. 2011 Nov;60(11):2730-9.
  31. Lanni A, Moreno M, Lombardi A, de Lange P, Silvestri E, Ragni M, et al. 3,5-diiodo-L-thyronine powerfully reduces adiposity in rats by increasing the burning of fats. FASEB J. 2005 Sep;19(11):1552-4.
  32. Berry MJ, Banu L, Larsen PR. Type I iodothyronine deiodinase is a selenocysteine-containing enzyme. Nature. 1991 Jan 31;349(6308):438-40.
  33. Berry MJ. Insights into the hierarchy of selenium incorporation. Nat Genet. 2005 Nov;37(11):1162-3.
  34. Hoffmann PR, Berry MJ. Selenoprotein synthesis: a unique translational mechanism used by a diverse family of proteins. Thyroid. 2005 Aug;15(8):769-75.
  35. Baqui MM, Gereben B, Harney JW, Larsen PR, Bianco AC. Distinct subcellular localization of transiently expressed types 1 and 2 iodothyronine deiodinases as determined by immunofluorescence confocal microscopy. Endocrinology. 2000 Nov;141(11):4309-12.
  36. Schneider MJ, Fiering SN, Thai B, Wu SY, St Germain E, Parlow AF, et al. Targeted disruption of the type 1 selenodeiodinase gene (Dio1) results in marked changes in thyroid hormone economy in mice. Endocrinology. 2006 Jan;147(1):580-9.
  37. Toyoda N, Berry MJ, Harney JW, Larsen PR. Topological analysis of the integral membrane protein, type 1 iodothyronine deiodinase (D1). J Biol Chem. 1995 May 19;270(20):12310-8.
  38. Curcio-Morelli C, Gereben B, Zavacki AM, Kim BW, Huang S, Harney JW, et al. In vivo dimerization of types 1, 2, and 3 iodothyronine selenodeiodinases. Endocrinology. 2003 Mar;144(3):937-46.
  39. Leonard JL, Simpson G, Leonard DM. Characterization of the protein dimerization domain responsible for assembly of functional selenodeiodinases. J Biol Chem. 2005 Mar 25;280(12):11093-100.
  40. Leonard JL, Visser TJ, Leonard DM. Characterization of the subunit structure of the catalytically active type I iodothyronine deiodinase. J Biol Chem. 2001 Jan 26;276(4):2600-7.
  41. Hoefig CS, Renko K, Piehl S, Scanlan TS, Bertoldi M, Opladen T et al. Does the aromatic L-amino acid decarboxylase contribute to thyronamine biosynthesis? Mol Cell Endocrinol. 2012 Feb 26;349(2):195-201.
  42. Berry MJ, Maia AL, Kieffer JD, Harney JW, Larsen PR. Substitution of cysteine for selenocysteine in type I iodothyronine deiodinase reduces the catalytic efficiency of the protein but enhances its translation. Endocrinology. 1992 Oct;131(4):1848-52.
  43. Berry MJ, Kieffer JD, Harney JW, Larsen PR. Selenocysteine confers the biochemical properties characteristic of the type I iodothyronine deiodinase. J Biol Chem. 1991 Aug 5;266(22):14155-8.
  44. Sanders JP, Van der Geyten S, Kaptein E, Darras VM, Kuhn ER, Leonard JL, et al. Characterization of a propylthiouracil-insensitive type I iodothyronine deiodinase. Endocrinology. 1997 Dec;138(12):5153-60.
  45. Callebaut I, Curcio-Morelli C, Mornon JP, Gereben B, Buettner C, Huang S, et al. The iodothyronine selenodeiodinases are thioredoxin-fold family proteins containing a glycoside hydrolase clan GH-A-like structure. J Biol Chem. 2003 Sep 19;278(38):36887-96.
  46. O'Mara BA, Dittrich W, Lauterio TJ, St Germain DL. Pretranslational regulation of type I 5'-deiodinase by thyroid hormones and in fasted and diabetic rats. Endocrinology. 1993 Oct;133(4):1715-23.
  47. Jakobs TC, Schmutzler C, Meissner J, Kohrle J. The promoter of the human type I 5'-deiodinase gene--mapping of the transcription start site and identification of a DR+4 thyroid-hormone-responsive element. Eur J Biochem. 1997 Jul 1;247(1):288-97.
  48. Toyoda N, Zavacki AM, Maia AL, Harney JW, Larsen PR. A novel retinoid X receptor-independent thyroid hormone response element is present in the human type 1 deiodinase gene. Mol Cell Biol. 1995 Sep;15(9):5100-12.
  49. Amma LL, Campos-Barros A, Wang Z, Vennstrom B, Forrest D. Distinct tissue-specific roles for thyroid hormone receptors beta and alpha1 in regulation of type 1 deiodinase expression. Mol Endocrinol. 2001 Mar;15(3):467-75.
  50. Zandieh Doulabi B, Platvoet-ter Schiphorst M, van Beeren HC, Labruyere WT, Lamers WH, Fliers E, et al. TR(beta)1 protein is preferentially expressed in the pericentral zone of rat liver and exhibits marked diurnal variation. Endocrinology. 2002 Mar;143(3):979-84.
  51. Toyoda N, Nishikawa M, Horimoto M, Yoshikawa N, Mori Y, Yoshimura M, et al. Synergistic effect of thyroid hormone and thyrotropin on iodothyronine 5'-deiodinase in FRTL-5 rat thyroid cells. Endocrinology. 1990 Sep;127(3):1199-205.
  52. Toyoda N, Nishikawa M, Horimoto M, Yoshikawa N, Mori Y, Yoshimura M, et al. Graves' immunoglobulin G stimulates iodothyronine 5'-deiodinating activity in FRTL-5 rat thyroid cells. J Clin Endocrinol Metab. 1990 Jun;70(6):1506-11.
  53. Beckett GJ, MacDougall DA, Nicol F, Arthur R. Inhibition of type I and type II iodothyronine deiodinase activity in rat liver, kidney and brain produced by selenium deficiency. Biochem J. 1989 May 1;259(3):887-92.
  54. Chanoine JP, Safran M, Farwell AP, Dubord S, Alex S, Stone S, et al. Effects of selenium deficiency on thyroid hormone economy in rats. Endocrinology. 1992 Oct;131(4):1787-92.
  55. Schomburg L, Riese C, Michaelis M, Griebert E, Klein MO, Sapin R, et al. Synthesis and metabolism of thyroid hormones is preferentially maintained in selenium-deficient transgenic mice. Endocrinology. 2006 Mar;147(3):1306-13.
  56. Berry MJ, Grieco D, Taylor BA, Maia AL, Kieffer JD, Beamer W, et al. Physiological and genetic analyses of inbred mouse strains with a type I iodothyronine 5' deiodinase deficiency. J Clin Invest. 1993 Sep;92(3):1517-28.
  57. Maia AL, Berry MJ, Sabbag R, Harney JW, Larsen PR. Structural and functional differences in the dio1 gene in mice with inherited type 1 deiodinase deficiency. Mol Endocrinol. 1995 Aug;9(8):969-80.
  58. Schoenmakers CH, Pigmans IG, Poland A, Visser TJ. Impairment of the selenoenzyme type I iodothyronine deiodinase in C3H/He mice. Endocrinology. 1993 Jan;132(1):357-61.
  59. Streckfuss F, Hamann I, Schomburg L, Michaelis M, Sapin R, Klein MO, et al. Hepatic deiodinase activity is dispensable for the maintenance of normal circulating thyroid hormone levels in mice. Biochem Biophys Res Commun. 2005 Nov 18;337(2):739-45.
  60. Galton VA, Schneider MJ, Clark AS, St Germain DL. Life without thyroxine to 3,5,3'-triiodothyronine conversion: studies in mice devoid of the 5'-deiodinases. Endocrinology. 2009 Jun;150(6):2957-63.
  61. Zavacki AM, Ying H, Christoffolete MA, Aerts G, So E, Harney JW, et al. Type 1 iodothyronine deiodinase is a sensitive marker of peripheral thyroid status in the mouse. Endocrinology. 2005 Mar;146(3):1568-75.
  62. de Jong FJ, Peeters RP, den Heijer T, van der Deure WM, Hofman A, Uitterlinden AG, et al. The association of polymorphisms in the type 1 and 2 deiodinase genes with circulating thyroid hormone parameters and atrophy of the medial temporal lobe. J Clin Endocrinol Metab. 2007 Feb;92(2):636-40.
  63. Panicker V, Cluett C, Shields B, Murray A, Parnell KS, Perry JR, et al. A common variation in deiodinase 1 gene DIO1 is associated with the relative levels of free thyroxine and triiodothyronine. J Clin Endocrinol Metab. 2008 Aug;93(8):3075-81.

64        Medici M, van der Deure WM, Verbiest M, Vermeulen SH, Hansen PS, Kiemeney LA, et al. A large-scale association analysis of 68 thyroid hormone pathway genes with serum TSH and FT4 levels. Eur J Endocrinol. 2011 May;164(5):781-8.

  1. Porcu E, Medici M, Pistis G, Volpato CB, Wilson SG, Cappola AR et al. A meta-analysis of thyroid-related traits reveals novel loci and gender-specific differences in the regulation of thyroid function. PLoS Genet. 2013;9(2):e1003266.
  2. Imai Y, Toyoda N, Maeda A, Kadobayashi T, Fangzheng G, Nishikawa M, et al. Type 2 iodothyronine deiodinase expression is upregulated by the protein kinase A-dependent pathway and is downregulated by the protein kinase C-dependent pathway in cultured human thyroid cells. Thyroid. 2001 Oct;11(10):899-907.
  3. Murakami M, Araki O, Hosoi Y, Kamiya Y, Morimura T, Ogiwara T, et al. Expression and regulation of type II iodothyronine deiodinase in human thyroid gland. Endocrinology. 2001 Jul;142(7):2961-7.
  4. Salvatore D, Tu H, Harney JW, Larsen PR. Type 2 iodothyronine deiodinase is highly expressed in human thyroid. J Clin Invest. 1996 Aug 15;98(4):962-8.
  5. Hosoi Y, Murakami M, Mizuma H, Ogiwara T, Imamura M, Mori M. Expression and regulation of type II iodothyronine deiodinase in cultured human skeletal muscle cells. J Clin Endocrinol Metab. 1999 Sep;84(9):3293-300.
  6. Dentice M, Morisco C, Vitale M, Rossi G, Fenzi G, Salvatore D. The different cardiac expression of the type 2 iodothyronine deiodinase gene between human and rat is related to the differential response of the Dio2 genes to Nkx-2.5 and GATA-4 transcription factors. Mol Endocrinol. 2003 Aug;17(8):1508-21.
  7. Guadano-Ferraz A, Obregon MJ, St Germain DL, Bernal J. The type 2 iodothyronine deiodinase is expressed primarily in glial cells in the neonatal rat brain. Proc Natl Acad Sci U S A. 1997 Sep 16;94(19):10391-6.
  8. Werneck de Castro JP, Fonseca TL, Ueta CB, McAninch EA, Abdalla S, Wittmann G et al. Differences in hypothalamic type 2 deiodinase ubiquitination explain localized sensitivity to thyroxine. J Clin Invest. 2015 Feb;125(2):769-81.
  9. Courtin F, Liva P, Gavaret JM, Toru-Delbauffe D, Pierre M. Induction of 5-deiodinase activity in astroglial cells by 12-O-tetradecanoylphorbol 13-acetate and fibroblast growth factors. J Neurochem. 1991 Apr;56(4):1107-13.
  10. Lamirand A, Mercier G, Ramauge M, Pierre M, Courtin F. Hypoxia stabilizes type 2 deiodinase activity in rat astrocytes. Endocrinology. 2007 Oct;148(10):4745-53.
  11. Simpson GI, Leonard DM, Leonard JL. Identification of the key residues responsible for the assembly of selenodeiodinases. J Biol Chem. 2006 May 26;281(21):14615-21.
  12. Bianco AC, Larsen PR. Cellular and structural biology of the deiodinases. Thyroid. 2005 Aug;15(8):777-86.
  13. Dentice M, Bandyopadhyay A, Gereben B, Callebaut I, Christoffolete MA, Kim BW, et al. The Hedgehog-inducible ubiquitin ligase subunit WSB-1 modulates thyroid hormone activation and PTHrP secretion in the developing growth plate. Nat Cell Biol. 2005 Jul;7(7):698-705.
  14. Kim BW, Zavacki AM, Curcio-Morelli C, Dentice M, Harney JW, Larsen PR, et al. Endoplasmic reticulum-associated degradation of the human type 2 iodothyronine deiodinase (D2) is mediated via an association between mammalian UBC7 and the carboxyl region of D2. Mol Endocrinol. 2003 Dec;17(12):2603-12.
  15. Sagar GD, Gereben B, Callebaut I, Mornon JP, Zeold A, da Silva WS, et al. Ubiquitination-induced conformational change within the deiodinase dimer is a switch regulating enzyme activity. Mol Cell Biol. 2007 Jul;27(13):4774-83.
  16. Arrojo EDR, Fonseca TL, Werneck-de-Castro JP, Bianco AC. Role of the type 2 iodothyronine deiodinase (D2) in the control of thyroid hormone signaling. Biochim Biophys Acta. 2012 Aug 29.
  17. Canettieri G, Celi FS, Baccheschi G, Salvatori L, Andreoli M, Centanni M. Isolation of human type 2 deiodinase gene promoter and characterization of a functional cyclic adenosine monophosphate response element. Endocrinology. 2000 May;141(5):1804-13.
  18. Arrojo EDR, Fonseca TL, Castillo M, Salathe M, Simovic G, Mohacsik P, et al. Endoplasmic reticulum stress decreases intracellular thyroid hormone activation via an eIF2a-mediated decrease in type 2 deiodinase synthesis. Mol Endocrinol. 2011 Dec;25(12):2065-75.
  19. Halperin Y, Shapiro LE, Surks MI. Down-regulation of type II L-thyroxine, 5'-monodeiodinase in cultured GC cells: different pathways of regulation by L-triiodothyronine and 3,3',5'-triiodo-L-thyronine. Endocrinology. 1994 Oct;135(4):1464-9.
  20. Burmeister LA, Pachucki J, St Germain DL. Thyroid hormones inhibit type 2 iodothyronine deiodinase in the rat cerebral cortex by both pre- and posttranslational mechanisms. Endocrinology. 1997 Dec;138(12):5231-7.
  21. Kim SW, Harney JW, Larsen PR. Studies of the hormonal regulation of type 2 5'-iodothyronine deiodinase messenger ribonucleic acid in pituitary tumor cells using semiquantitative reverse transcription-polymerase chain reaction. Endocrinology. 1998 Dec;139(12):4895-905.
  22. Buettner C, Harney JW, Larsen PR. The role of selenocysteine 133 in catalysis by the human type 2 iodothyronine deiodinase. Endocrinology. 2000 Dec;141(12):4606-12.
  23. Kuiper GG, Klootwijk W, Visser TJ. Substitution of cysteine for a conserved alanine residue in the catalytic center of type II iodothyronine deiodinase alters interaction with reducing cofactor. Endocrinology. 2002 Apr;143(4):1190-8.
  24. Steinsapir J, Bianco AC, Buettner C, Harney J, Larsen PR. Substrate-induced down-regulation of human type 2 deiodinase (hD2) is mediated through proteasomal degradation and requires interaction with the enzyme's active center. Endocrinology. 2000 Mar;141(3):1127-35.
  25. Salvatore D, Harney JW, Larsen PR. Mutation of the Secys residue 266 in human type 2 selenodeiodinase alters 75Se incorporation without affecting its biochemical properties. Biochimie. 1999 May;81(5):535-8.
  26. Gereben B, Kollar A, Harney JW, Larsen PR. The mRNA structure has potent regulatory effects on type 2 iodothyronine deiodinase expression. Mol Endocrinol. 2002 Jul;16(7):1667-79.
  27. Gereben B, Salvatore D, Harney JW, Tu HM, Larsen PR. The human, but not rat, dio2 gene is stimulated by thyroid transcription factor-1 (TTF-1). Mol Endocrinol. 2001 Jan;15(1):112-24.
  28. Bartha T, Kim SW, Salvatore D, Gereben B, Tu HM, Harney JW, et al. Characterization of the 5'-flanking and 5'-untranslated regions of the cyclic adenosine 3',5'-monophosphate-responsive human type 2 iodothyronine deiodinase gene. Endocrinology. 2000 Jan;141(1):229-37.
  29. Kim BW, Daniels GH, Harrison BJ, Price A, Harney JW, Larsen PR, et al. Overexpression of type 2 iodothyronine deiodinase in follicular carcinoma as a cause of low circulating free thyroxine levels. J Clin Endocrinol Metab. 2003 Feb;88(2):594-8.
  30. Miyauchi A, Takamura Y, Ito Y, Miya A, Kobayashi K, Matsuzuka F, et al. 3,5,3'-Triiodothyronine thyrotoxicosis due to increased conversion of administered levothyroxine in patients with massive metastatic follicular thyroid carcinoma. J Clin Endocrinol Metab. 2008 Jun;93(6):2239-42.
  31. Takano T, Miyauchi A, Ito Y, Amino N. Thyroxine to triiodothyronine hyperconversion thyrotoxicosis in patients with large metastases of follicular thyroid carcinoma. Thyroid. 2006 Jun;16(6):615-8.
  32. Schneider MJ, Fiering SN, Pallud SE, Parlow AF, St Germain DL, Galton VA. Targeted disruption of the type 2 selenodeiodinase gene (DIO2) results in a phenotype of pituitary resistance to T4. Mol Endocrinol. 2001 Dec;15(12):2137-48.
  33. Marsili A, Ramadan W, Harney JW, Mulcahey M, Castroneves LA, Goemann IM, et al. Type 2 iodothyronine deiodinase levels are higher in slow-twitch than fast-twitch mouse skeletal muscle and are increased in hypothyroidism. Endocrinology. 2010 Dec;151(12):5952-60.
  34. Werneck-de-Castro JP, Fonseca TL, Ignacio DL, Fernandes GW, Andrade-Feraud CM, Lartey LJ et al. Thyroid hormone signaling in male mouse skeletal muscle is largely independent of D2 in myocytes. Endocrinology. 2015 Oct;156(10):3842-52.
  35. Ignacio DL, Silvestre DH, Palmer E, Bocco B, Fonseca T, Gereben B et al. Early developmental disruption of type 2 deiodinase pathway in mouse skeletal muscle does not impair muscle function. Thyroid. 2016 Dec 14. [Epub ahead of print]
  36. Campos-Barros A, Amma LL, Faris JS, Shailam R, Kelley MW, Forrest D. Type 2 iodothyronine deiodinase expression in the cochlea before the onset of hearing. Proc Natl Acad Sci U S A. 2000 Feb 1;97(3):1287-92.
  37. Ng L, Goodyear RJ, Woods CA, Schneider MJ, Diamond E, Richardson GP, et al. Hearing loss and retarded cochlear development in mice lacking type 2 iodothyronine deiodinase. Proc Natl Acad Sci U S A. 2004 Mar 9;101(10):3474-9.
  38. Hall JA, Ribich S, Christoffolete MA, Simovic G, Correa-Medina M, Patti ME, et al. Absence of thyroid hormone activation during development underlies a permanent defect in adaptive thermogenesis. Endocrinology. 2010 Sep;151(9):4573-82.
  39. de Jesus LA, Carvalho SD, Ribeiro MO, Schneider M, Kim SW, Harney JW, et al. The type 2 iodothyronine deiodinase is essential for adaptive thermogenesis in brown adipose tissue. J Clin Invest. 2001 Nov;108(9):1379-85.
  40. Castillo M, Hall JA, Correa-Medina M, Ueta C, Kang HW, Cohen DE, et al. Disruption of thyroid hormone activation in type 2 deiodinase knockout mice causes obesity with glucose intolerance and liver steatosis only at thermoneutrality. Diabetes. 2011 Apr;60(4):1082-9.
  41. Bocco BM, Werneck-de-Castro JP, Oliveira KC, Fernandes GW, Fonseca TL, Nascimento BP et al. Type 2 deiodinase disruption in astrocytes results in anxiety-depressive-like behavior in male mice. Endocrinology. 2016 Sep;157(9):3682-95.
  42. Dentice M, Marsili A, Ambrosio R, Guardiola O, Sibilio A, Paik JH, et al. The FoxO3/type 2 deiodinase pathway is required for normal mouse myogenesis and muscle regeneration. J Clin Invest. 2010 Nov;120(11):4021-30.
  43. Bassett JH, Boyde A, Howell PG, Bassett RH, Galliford TM, Archanco M, et al. Optimal bone strength and mineralization requires the type 2 iodothyronine deiodinase in osteoblasts. Proc Natl Acad Sci U S A. Apr 20;107(16):7604-9.
  44. Kwakkel J, van Beeren HC, Ackermans MT, Platvoet-Ter Schiphorst MC, Fliers E, Wiersinga WM, et al. Skeletal muscle deiodinase type 2 regulation during illness in mice. J Endocrinol. 2009 Nov;203(2):263-70.
  45. Heemstra KA, Soeters MR, Fliers E, Serlie MJ, Burggraaf J, van Doorn MB, et al. Type 2 iodothyronine deiodinase in skeletal muscle: effects of hypothyroidism and fasting. J Clin Endocrinol Metab. 2009 Jun;94(6):2144-50.
  46. Barca-Mayo O, Liao XH, DiCosmo C, Dumitrescu A, Moreno-Vinasco L, Wade MS, et al. Role of type 2 deiodinase in response to acute lung injury (ALI) in mice. Proc Natl Acad Sci U S A. 2011 Dec 6;108(49):E1321-9.
  47. Schoenmakers E, Agostini M, Mitchell C, Schoenmakers N, Papp L, Rajanayagam O, et al. Mutations in the selenocysteine insertion sequence-binding protein 2 gene lead to a multisystem selenoprotein deficiency disorder in humans. J Clin Invest. 2010 Dec;120(12):4220-35.
  48. Dumitrescu AM, Liao XH, Abdullah MS, Lado-Abeal J, Majed FA, Moeller LC, et al. Mutations in SECISBP2 result in abnormal thyroid hormone metabolism. Nat Genet. 2005 Nov;37(11):1247-52.
  49. Canani LH, Capp C, Dora JM, Meyer EL, Wagner MS, Harney JW, et al. The type 2 deiodinase A/G (Thr92Ala) polymorphism is associated with decreased enzyme velocity and increased insulin resistance in patients with type 2 diabetes mellitus. J Clin Endocrinol Metab. 2005 Jun;90(6):3472-8.
  50. Grarup N, Andersen MK, Andreasen CH, Albrechtsen A, Borch-Johnsen K, Jorgensen T, et al. Studies of the common DIO2 Thr92Ala polymorphism and metabolic phenotypes in 7342 Danish white subjects. J Clin Endocrinol Metab. 2007 Jan;92(1):363-6.
  51. Peeters RP, van den Beld AW, Attalki H, Toor H, de Rijke YB, Kuiper GG, et al. A new polymorphism in the type II deiodinase gene is associated with circulating thyroid hormone parameters. Am J Physiol Endocrinol Metab. 2005 Jul;289(1):E75-81.
  52. Peeters RP, van Toor H, Klootwijk W, de Rijke YB, Kuiper GG, Uitterlinden AG, et al. Polymorphisms in thyroid hormone pathway genes are associated with plasma TSH and iodothyronine levels in healthy subjects. J Clin Endocrinol Metab. 2003 Jun;88(6):2880-8.
  53. Meulenbelt I, Min JL, Bos S, Riyazi N, Houwing-Duistermaat JJ, van der Wijk HJ, et al. Identification of DIO2 as a new susceptibility locus for symptomatic osteoarthritis. Hum Mol Genet. 2008 Jun 15;17(12):1867-75.
  54. Bates JM, St Germain DL, Galton VA. Expression profiles of the three iodothyronine deiodinases, D1, D2, and D3, in the developing rat. Endocrinology. 1999 Feb;140(2):844-51.
  55. Galton VA, Martinez E, Hernandez A, St Germain EA, Bates JM, St Germain DL. Pregnant rat uterus expresses high levels of the type 3 iodothyronine deiodinase. J Clin Invest. 1999 Apr;103(7):979-87.
  56. Santini F, Vitti P, Chiovato L, Ceccarini G, Macchia M, Montanelli L, et al. Role for inner ring deiodination preventing transcutaneous passage of thyroxine. J Clin Endocrinol Metab. 2003 Jun;88(6):2825-30.
  57. Dentice M, Luongo C, Huang S, Ambrosio R, Elefante A, Mirebeau-Prunier D, et al. Sonic hedgehog-induced type 3 deiodinase blocks thyroid hormone action enhancing proliferation of normal and malignant keratinocytes. Proc Natl Acad Sci U S A. 2007 Sep 4;104(36):14466-71.
  58. Huang SA, Dorfman DM, Genest DR, Salvatore D, Larsen PR. Type 3 iodothyronine deiodinase is highly expressed in the human uteroplacental unit and in fetal epithelium. J Clin Endocrinol Metab. 2003 Mar;88(3):1384-8.
  59. Huang SA, Tu HM, Harney JW, Venihaki M, Butte AJ, Kozakewich HP, et al. Severe hypothyroidism caused by type 3 iodothyronine deiodinase in infantile hemangiomas. N Engl J Med. 2000 Jul 20;343(3):185-9.
  60. Sato K, Robbins J. Thyroid hormone metabolism in cultured monkey hepatocarcinoma cells. Monodeiodination activity in relation to cell growth. J Biol Chem. 1980 Aug 10;255(15):7347-52.
  61. Baqui M, Botero D, Gereben B, Curcio C, Harney JW, Salvatore D, et al. Human type 3 iodothyronine selenodeiodinase is located in the plasma membrane and undergoes rapid internalization to endosomes. J Biol Chem. 2003 Jan 10;278(2):1206-11.
  62. Sagar GD, Gereben B, Callebaut I, Mornon JP, Zeold A, Curcio-Morelli C, et al. The thyroid hormone-inactivating deiodinase functions as a homodimer. Mol Endocrinol. 2008 Jun;22(6):1382-93.
  63. Tu HM, Legradi G, Bartha T, Salvatore D, Lechan RM, Larsen PR. Regional expression of the type 3 iodothyronine deiodinase messenger ribonucleic acid in the rat central nervous system and its regulation by thyroid hormone. Endocrinology. 1999 Feb;140(2):784-90.
  64. Barca-Mayo O, Liao XH, Alonso M, Di Cosmo C, Hernandez A, Refetoff S, et al. Thyroid hormone receptor alpha and regulation of type 3 deiodinase. Mol Endocrinol. 2011 Apr;25(4):575-83.
  65. Macchia PE, Takeuchi Y, Kawai T, Cua K, Gauthier K, Chassande O, et al. Increased sensitivity to thyroid hormone in mice with complete deficiency of thyroid hormone receptor alpha. Proc Natl Acad Sci U S A. 2001 Jan 2;98(1):349-54.
  66. Kester MH, Martinez de Mena R, Obregon MJ, Marinkovic D, Howatson A, Visser TJ, et al. Iodothyronine levels in the human developing brain: major regulatory roles of iodothyronine deiodinases in different areas. J Clin Endocrinol Metab. 2004 Jul;89(7):3117-28.
  67. Santini F, Pinchera A, Ceccarini G, Castagna M, Rosellini V, Mammoli C, et al. Evidence for a role of the type III-iodothyronine deiodinase in the regulation of 3,5,3'-triiodothyronine content in the human central nervous system. Eur J Endocrinol. 2001 Jun;144(6):577-83.
  68. Huang SA, Fish SA, Dorfman DM, Salvatore D, Kozakewich HP, Mandel SJ, 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 Oct;87(10):4457-61.
  69. Peeters RP, van der Geyten S, Wouters PJ, Darras VM, van Toor H, Kaptein E, et al. Tissue thyroid hormone levels in critical illness. J Clin Endocrinol Metab. 2005 Dec;90(12):6498-507.
  70. Peeters RP, Wouters PJ, Kaptein E, van Toor H, Visser TJ, Van den Berghe G. Reduced activation and increased inactivation of thyroid hormone in tissues of critically ill patients. J Clin Endocrinol Metab. 2003 Jul;88(7):3202-11.
  71. Peeters RP, Wouters PJ, van Toor H, Kaptein E, Visser TJ, Van den Berghe G. Serum 3,3',5'-triiodothyronine (rT3) and 3,5,3'-triiodothyronine/rT3 are prognostic markers in critically ill patients and are associated with postmortem tissue deiodinase activities. J Clin Endocrinol Metab. 2005 Aug;90(8):4559-65.
  72. Hernandez A, Fiering S, Martinez E, Galton VA, St Germain D. The gene locus encoding iodothyronine deiodinase type 3 (Dio3) is imprinted in the fetus and expresses antisense transcripts. Endocrinology. 2002 Nov;143(11):4483-6.
  73. Hernandez A, Martinez ME, Fiering S, Galton VA, St Germain D. Type 3 deiodinase is critical for the maturation and function of the thyroid axis. J Clin Invest. 2006 Feb;116(2):476-84.
  74. Hernandez A, Martinez ME, Liao XH, Van Sande J, Refetoff S, Galton VA, et al. Type 3 deiodinase deficiency results in functional abnormalities at multiple levels of the thyroid axis. Endocrinology. 2007 Dec;148(12):5680-7.
  75. Kempers MJ, van Tijn DA, van Trotsenburg AS, de Vijlder JJ, Wiedijk BM, Vulsma T. Central congenital hypothyroidism due to gestational hyperthyroidism: detection where prevention failed. J Clin Endocrinol Metab. 2003 Dec;88(12):5851-7.
  76. Hernandez A, Garcia B, Obregon MJ. Gene expression from the imprinted Dio3 locus is associated with cell proliferation of cultured brown adipocytes. Endocrinology. 2007 Aug;148(8):3968-76.
  77. Charalambous M, Ferron SR, da Rocha ST, Murray AJ, Rowland T, Ito M, et al. Imprinted gene dosage is critical for the transition to independent life. Cell Metab. 2012 Feb 8;15(2):209-21.
  78. Van der Geyten S, Buys N, Sanders JP, Decuypere E, Visser TJ, Kuhn ER, et al. Acute pretranslational regulation of type III iodothyronine deiodinase by growth hormone and dexamethasone in chicken embryos. Mol Cell Endocrinol. 1999 Jan 25;147(1-2):49-56.
  79. Van der Geyten S, Sanders JP, Kaptein E, Darras VM, Kuhn ER, Leonard JL, et al. Expression of chicken hepatic type I and type III iodothyronine deiodinases during embryonic development. Endocrinology. 1997 Dec;138(12):5144-52.
  80. Richard K, Hume R, Kaptein E, Sanders JP, van Toor H, De Herder WW, et al. Ontogeny of iodothyronine deiodinases in human liver. J Clin Endocrinol Metab. 1998 Aug;83(8):2868-74.
  81. Koopdonk-Kool JM, de Vijlder JJ, Veenboer GJ, Ris-Stalpers C, Kok JH, Vulsma T, et al. Type II and type III deiodinase activity in human placenta as a function of gestational age. J Clin Endocrinol Metab. 1996 Jun;81(6):2154-8.
  82. Roti E, Gnudi A, Braverman LE. The placental transport, synthesis and metabolism of hormones and drugs which affect thyroid function. Endocr Rev. 1983 Spring;4(2):131-49.
  83. Ng L, Hernandez A, He W, Ren T, Srinivas M, Ma M, et al. A protective role for type 3 deiodinase, a thyroid hormone-inactivating enzyme, in cochlear development and auditory function. Endocrinology. 2009 Apr;150(4):1952-60.
  84. Peeters RP, Hernandez A, Ng L, Ma M, Sharlin DS, Pandey M, et al. Cerebellar Abnormalities in Mice Lacking Type 3 Deiodinase and Partial Reversal of Phenotype by Deletion of Thyroid Hormone Receptor alpha1. Endocrinology. 2012 Nov 16.
  85. Ng L, Lyubarsky A, Nikonov SS, Ma M, Srinivas M, Kefas B, et al. Type 3 deiodinase, a thyroid-hormone-inactivating enzyme, controls survival and maturation of cone photoreceptors. J Neurosci. 2010 Mar 3;30(9):3347-57.
  86. Medina MC, Molina J, Gadea Y, Fachado A, Murillo M, Simovic G, et al. The thyroid hormone-inactivating type III deiodinase is expressed in mouse and human beta-cells and its targeted inactivation impairs insulin secretion. Endocrinology. 2011 Oct;152(10):3717-27.
  87. Olivares EL, Marassi MP, Fortunato RS, da Silva AC, Costa-e-Sousa RH, Araujo IG, et al. Thyroid function disturbance and type 3 iodothyronine deiodinase induction after myocardial infarction in rats a time course study. Endocrinology. 2007 Oct;148(10):4786-92.
  88. Simonides WS, Mulcahey MA, Redout EM, Muller A, Zuidwijk MJ, Visser TJ, et al. Hypoxia-inducible factor induces local thyroid hormone inactivation during hypoxic-ischemic disease in rats. J Clin Invest. 2008 Mar;118(3):975-83.
  89. Wassen FW, Schiel AE, Kuiper GG, Kaptein E, Bakker O, Visser TJ, et al. Induction of thyroid hormone-degrading deiodinase in cardiac hypertrophy and failure. Endocrinology. 2002 Jul;143(7):2812-5.
  90. Jo S, Kallo I, Bardoczi Z, Arrojo e Drigo R, Zeold A, Liposits Z, et al. Neuronal hypoxia induces Hsp40-mediated nuclear import of type 3 deiodinase as an adaptive mechanism to reduce cellular metabolism. J Neurosci. 2012 Jun 20;32(25):8491-500.
  91. Ueta CB, Oskouei BN, Olivares EL, Pinto JR, Correa MM, Simovic G, et al. Absence of myocardial thyroid hormone inactivating deiodinase results in restrictive cardiomyopathy in mice. Mol Endocrinol. 2012 May;26(5):809-18.
  92. 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.
  93. Boelen A, Kwakkel J, Alkemade A, Renckens R, Kaptein E, Kuiper G, et al. Induction of type 3 deiodinase activity in inflammatory cells of mice with chronic local inflammation. Endocrinology. 2005 Dec;146(12):5128-34.
  94. Boelen A, Boorsma J, Kwakkel J, Wieland CW, Renckens R, Visser TJ, et al. Type 3 deiodinase is highly expressed in infiltrating neutrophilic granulocytes in response to acute bacterial infection. Thyroid. 2008 Oct;18(10):1095-103.
  95. Kester MH, Toussaint MJ, Punt CA, Matondo R, Aarnio AM, Darras VM, et al. Large induction of type III deiodinase expression after partial hepatectomy in the regenerating mouse and rat liver. Endocrinology. 2009 Jan;150(1):540-5.
  96. Li WW, Le Goascogne C, Ramauge M, Schumacher M, Pierre M, Courtin F. Induction of type 3 iodothyronine deiodinase by nerve injury in the rat peripheral nervous system. Endocrinology. 2001 Dec;142(12):5190-7.
  97. Dentice M, Salvatore D. Deiodinases: the balance of thyroid hormone: local impact of thyroid hormone inactivation. J Endocrinol. 2011 Jun;209(3):273-82.
  98. Tanimizu N, Miyajima A. Molecular mechanism of liver development and regeneration. Int Rev Cytol. 2007;259:1-48.
  99. Kester MH, Kuiper GG, Versteeg R, Visser TJ. Regulation of type III iodothyronine deiodinase expression in human cell lines. Endocrinology. 2006 Dec;147(12):5845-54.
  100. Romitti M, Wajner SM, Zennig N, Goemann IM, Bueno AL, Meyer EL, et al. Increased type 3 deiodinase expression in papillary thyroid carcinoma. Thyroid. 2012 Sep;22(9):897-904.
  101. Pacifici GM, Coughtrie MW. Human Cytosolic Sulfotransferases. Baco Raton: Taylor & Francis; 2005.
  102. Blanchard RL, Freimuth RR, Buck J, Weinshilboum RM, Coughtrie MW. A proposed nomenclature system for the cytosolic sulfotransferase (SULT) superfamily. Pharmacogenetics. 2004 Mar;14(3):199-211.
  103. Fujita K, Nagata K, Ozawa S, Sasano H, Yamazoe Y. Molecular cloning and characterization of rat ST1B1 and human ST1B2 cDNAs, encoding thyroid hormone sulfotransferases. J Biochem. 1997 Nov;122(5):1052-61.
  104. Kester MH, Kaptein E, Roest TJ, van Dijk CH, Tibboel D, Meinl W, et al. Characterization of human iodothyronine sulfotransferases. J Clin Endocrinol Metab. 1999 Apr;84(4):1357-64.
  105. Kester MH, van Dijk CH, Tibboel D, Hood AM, Rose NJ, Meinl W, et al. Sulfation of thyroid hormone by estrogen sulfotransferase. J Clin Endocrinol Metab. 1999 Jul;84(7):2577-80.
  106. Li X, Anderson RJ. Sulfation of iodothyronines by recombinant human liver steroid sulfotransferases. Biochem Biophys Res Commun. 1999 Oct 5;263(3):632-9.
  107. Li X, Clemens DL, Anderson RJ. Sulfation of iodothyronines by human sulfotransferase 1C1 (SULT1C1)*. Biochem Pharmacol. 2000 Dec 1;60(11):1713-6.
  108. Pietsch CA, Scanlan TS, Anderson RJ. Thyronamines are substrates for human liver sulfotransferases. Endocrinology. 2007 Apr;148(4):1921-7.
  109. Visser TJ, Kaptein E, Glatt H, Bartsch I, Hagen M, Coughtrie MW. Characterization of thyroid hormone sulfotransferases. Chem Biol Interact. 1998 Feb 20;109(1-3):279-91.
  110. Wang J, Falany JL, Falany CN. Expression and characterization of a novel thyroid hormone-sulfating form of cytosolic sulfotransferase from human liver. Mol Pharmacol. 1998 Feb;53(2):274-82.
  111. Venkatachalam KV, Akita H, Strott CA. Molecular cloning, expression, and characterization of human bifunctional 3'-phosphoadenosine 5'-phosphosulfate synthase and its functional domains. J Biol Chem. 1998 Jul 24;273(30):19311-20.
  112. Eelkman Rooda SJ, Kaptein E, Visser TJ. Serum triiodothyronine sulfate in man measured by radioimmunoassay. J Clin Endocrinol Metab. 1989 Sep;69(3):552-6.
  113. Chopra MFI. Nonthyroidal illness syndrome or euthyroid sick syndrome? Endocr Pract. 1996;2(1):45-52.
  114. Chopra IJ, Nguyen D. Demonstration of thyromimetic effects of 3,5,3'-triiodothyronine sulfate (T3S) in euthyroid rats. Thyroid. 1996 Jun;6(3):229-32.
  115. Kester MH, Kaptein E, Van Dijk CH, Roest TJ, Tibboel D, Coughtrie MW, et al. Characterization of iodothyronine sulfatase activities in human and rat liver and placenta. Endocrinology. 2002 Mar;143(3):814-9.
  116. Santini F, Chopra IJ, Wu SY, Solomon DH, Chua Teco GN. Metabolism of 3,5,3'-triiodothyronine sulfate by tissues of the fetal rat: a consideration of the role of desulfation of 3,5,3'-triiodothyronine sulfate as a source of T3. Pediatr Res. 1992 Jun;31(6):541-4.
  117. Wu SY, Huang WS, Ho E, Wu ES, Fisher DA. Compound W, a 3,3'-diiodothyronine sulfate cross-reactive substance in serum from pregnant women--a potential marker for fetal thyroid function. Pediatr Res. 2007 Mar;61(3):307-12.
  118. Mackenzie PI, Bock KW, Burchell B, Guillemette C, Ikushiro S, Iyanagi T, et al. Nomenclature update for the mammalian UDP glycosyltransferase (UGT) gene superfamily. Pharmacogenet Genomics. 2005 Oct;15(10):677-85.
  119. Hennemann G, Visser TJ. Thyroid hormone synthesis, plasma membrane transport, and metabolism.1997.
  120. Brouwer A, Morse DC, Lans MC, Schuur AG, Murk AJ, Klasson-Wehler E, et al. Interactions of persistent environmental organohalogens with the thyroid hormone system: mechanisms and possible consequences for animal and human health. Toxicol Ind Health. 1998 Jan-Apr;14(1-2):59-84.
  121. Klaassen CD, Hood AM. Effects of microsomal enzyme inducers on thyroid follicular cell proliferation and thyroid hormone metabolism. Toxicol Pathol. 2001 Jan-Feb;29(1):34-40.
  122. Visser TJ, Kaptein E, van Toor H, van Raaij JA, van den Berg KJ, Joe CT, et al. Glucuronidation of thyroid hormone in rat liver: effects of in vivo treatment with microsomal enzyme inducers and in vitro assay conditions. Endocrinology. 1993 Nov;133(5):2177-86.
  123. Benedetti MS, Whomsley R, Baltes E, Tonner F. Alteration of thyroid hormone homeostasis by antiepileptic drugs in humans: involvement of glucuronosyltransferase induction. Eur J Clin Pharmacol. 2005 Dec;61(12):863-72.
  124. Kato Y, Ikushiro S, Emi Y, Tamaki S, Suzuki H, Sakaki T, et al. Hepatic UDP-glucuronosyltransferases responsible for glucuronidation of thyroxine in humans. Drug Metab Dispos. 2008 Jan;36(1):51-5.
  125. Tong Z, Li H, Goljer I, McConnell O, Chandrasekaran A. In vitro glucuronidation of thyroxine and triiodothyronine by liver microsomes and recombinant human UDP-glucuronosyltransferases. Drug Metab Dispos. 2007 Dec;35(12):2203-10.
  126. Yamanaka H, Nakajima M, Katoh M, Yokoi T. Glucuronidation of thyroxine in human liver, jejunum, and kidney microsomes. Drug Metab Dispos. 2007 Sep;35(9):1642-8.
  127. Yoder Graber AL, Ramirez J, Innocenti F, Ratain MJ. UGT1A1*28 genotype affects the in-vitro glucuronidation of thyroxine in human livers. Pharmacogenet Genomics. 2007 Aug;17(8):619-27.
  128. Moreno M, Kaptein E, Goglia F, Visser TJ. Rapid glucuronidation of tri- and tetraiodothyroacetic acid to ester glucuronides in human liver and to ether glucuronides in rat liver. Endocrinology. 1994 Sep;135(3):1004-9.
  129. Bernal J. Thyroid hormone receptors in brain development and function. Nat Clin Pract Endocrinol Metab. 2007 Mar;3(3):249-59.
  130. Bernal J. Thyroid hormone transport in developing brain. Curr Opin Endocrinol Diabetes Obes. 2011 Oct;18(5):295-9.
  131. Heuer H, Maier MK, Iden S, Mittag J, Friesema EC, Visser TJ, et al. The monocarboxylate transporter 8 linked to human psychomotor retardation is highly expressed in thyroid hormone-sensitive neuron populations. Endocrinology. 2005 Apr;146(4):1701-6.
  132. Fonseca TL, Fernandes GW, McAninch EA, Bocco BM, Abdalla SM, Ribeiro MO et al. Perinatal deiodinase 2 expression in hepatocytes defines epigenetic susceptibility to liver steatosis and obesity. Proc Natl Acad Sci U S A. 2015 Nov 10;112(45):14018-23.
  133. Huang SA. Deiodination and cellular proliferation: parallels between development, differentiation, tumorigenesis, and now regeneration. Endocrinology. 2009 Jan;150(1):3-4.
  134. Wittmann G, Harney JW, Singru PS, Nouriel SS, Reed Larsen P, Lechan RM. Inflammation-inducible type 2 deiodinase expression in the leptomeninges, choroid plexus, and at brain blood vessels in male rodents. Endocrinology. 2014 May;155(5):2009-19.
  135. Kwakkel J, Surovtseva OV, de Vries EM, Stap J, Fliers E, Boelen A. A novel role for the thyroid hormone-activating enzyme type 2 deiodinase in the inflammatory response of macrophages. Endocrinology. 2014;155:2725–2734.
  136. Medici M, Visser WE, Visser TJ, Peeters RP. Genetic determination of the hypothalamic-pituitary-thyroid axis: where do we stand? Endocr Rev. 2015 Apr;36(2):214-44.
  137. Zevenbergen C, Klootwijk W, Peeters RP, Medici M, de Rijke YB, Huisman SA, et al. Functional analysis of novel genetic variation in the thyroid hormone activating type 2 deiodinase.

J Clin Endocrinol Metab. 2014 Nov;99(11):E2429-36.

  1. McAninch EA, Jo S, Preite NZ, Farkas E, Mohácsik P, Fekete C, Egri P et al. Prevalent polymorphism in thyroid hormone-activating enzyme leaves a genetic fingerprint that underlies associated clinical syndromes. J Clin Endocrinol Metab. 2015 Mar;100(3):920-33.
  2. Maynard MA, Marino-Enriquez A, Fletcher JA, Dorfman DM, Raut CP, Yassa L et al. Thyroid hormone inactivation in gastrointestinal stromal tumors. N Engl J Med. 2014 Apr 3;370(14):1327-34
  3. Schweizer U, Schlicker C, Braun D, Köhrle J, Steegborn C5. Crystal structure of mammalian selenocysteine-dependent iodothyronine deiodinase suggests a peroxiredoxin-like catalytic mechanism. Proc Natl Acad Sci U S A. 2014 Jul 22;111(29):10526-31.
  4. Martinez ME, Charalambous M, Saferali A, Fiering S, Naumova AK, St Germain D et al. Genomic imprinting variations in the mouse type 3 deiodinase gene between tissues and brain regions. Mol Endocrinol. 2014 Nov;28(11):1875-86.
  5. Medina MC, Fonesca TL, Molina J, Fachado A, Castillo M, Dong L et al. Maternal inheritance of an inactive type III deiodinase gene allele affects mouse pancreatic β-cells and disrupts glucose homeostasis. Endocrinology. 2014 Aug;155(8):3160-71.
  6. Stohn JP, Martinez ME, Hernandez A. Decreased anxiety- and depression-like behaviors and hyperactivity in a type 3 deiodinase-deficient mouse showing brain thyrotoxicosis and peripheral hypothyroidism. Psychoneuroendocrinology. 2016 Aug 24;74:46-56.
  7. Stohn JP, Martinez ME, Matoin K, Morte B, Bernal J, Galton VA et al. Mct8 deficiency in male mice mitigates the phenotypic abnormalities associated with the absence of a functional type 3 deiodinase. Endocrinology. 2016 Aug;157(8):3266-77.
  8. Houbrechts AM, Vergauwen L, Bagci E, Van Houcke J, Heijlen M, Kulemeka B et al. Deiodinase knockdown affects zebrafish eye development at the level of gene expression, morphology and function. Mol Cell Endocrinol. 2016 Mar 15;424:81-93.
  9. Martinez ME, Karaczyn A, Stohn JP, Donnelly WT, Croteau W, Peeters RP et al. The type 3 deiodinase is a critical determinant of appropriate thyroid hormone action in the developing testis. Endocrinology. 2016 Mar;157(3):1276-88.
  10. Wu Z, Martinez ME, St Germain DL, Hernandez A. Type 3 deiodinase role on central thyroid hormone action affects the leptin-melanocortin system and circadian activity. Endocrinology. 2016 Dec 2:en20161680. [Epub ahead of print]
  11. Van der Spek AH, Bloise FF, Tigchelaar W, Dentice M, Salvatore D, van der Wel NN et al. The Thyroid Hormone Inactivating Enzyme Type 3 Deiodinase is Present in Bactericidal Granules and the Cytoplasm of Human Neutrophils. Endocrinology. 2016 Aug;157(8):3293-305.
  12. Visser WE, Bombardieri CR, Zevenbergen C, Barnhoorn S, Ottaviani A, van der Pluijm I et al. Tissue-specific suppression of thyroid hormone signaling in various mouse models of aging. PLoS One. 2016 Mar 8;11(3):e0149941.
  13. Dentice M, Ambrosio R, Damiano V, Sibilio A, Luongo C, Guardiola O et al. Intracellular inactivation of thyroid hormone is a survival mechanism for muscle stem cell proliferation and lineage progression. Cell Metab. 2014 Dec 2;20(6):1038-48.
  14. Luongo C, Ambrosio R, Salzano S, Dlugosz AA, Missero C, Dentice M. The sonic hedgehog-induced type 3 deiodinase facilitates tumorigenesis of basal cell carcinoma by reducing Gli2 inactivation. Endocrinology. 2014 Jun;155(6):2077-88.
  15. Di Girolamo D, Ambrosio R, De Stefano MA, Mancino G, Porcelli T, Luongo C et al. Reciprocal interplay between thyroid hormone and microRNA-21 regulates hedgehog pathway–driven skin tumorigenesis. J Clin Invest. 2016 Jun 1; 126(6): 2308–2320.
  16. Galton VA, de Waard E, Parlow AF, St Germain DL, Hernandez A. Life without the iodothyronine deiodinases. Endocrinology. 2014 Oct;155(10):4081-7.

 

 

 

 

 

Testing for Endocrine Hypertension

ABSTRACT

Endocrine hypertension belongs to the group of secondary forms of hypertension and mostly is caused by disorders of the adrenal gland. There are also nonadrenal endocrine conditions that can lead to hypertension secondary to hormonal imbalance and this chapter reviews how to best identify patients with and test for such disorders. For complete coverage of all related areas of Endocrinology, please see our FREE on line web-book, www.endotext.org.

INTRODUCTION

Hypertension (HTN) is a preventable contributor to disease and death in humans. HTN is commonly defined as a blood pressure (BP) ≥140/90 mm Hg for adults ≥18 years old based on the mean of ≥2 properly measured seated BP readings on each of 2 or more office visits. The overall prevalence of HTN among U.S. adults has not changed appreciably since 2009-2010, affecting ~29.1% of the population (1). The majority of patients with HTN (82.8%) are aware of their HTN and taking medication to lower it (75.7%), although BP control (<140/90 mm Hg) is suboptimal (51.9%) (1). Other studies have demonstrated BP control in < 1 in 3 patients (2, 3).

The prevalence of resistant HTN varies from 34-53% in different large studies: ALLHAT (34%), NHANES (53%) or Framingham Heart Study (48%) (4). Control to <140/90 mm Hg has been a matter of debate over the past decade. Recently, a large randomized trial of intensive versus standard BP control (SPRINT) among patients at high risk for cardiovascular events but without diabetes mellitus, showed that targeting a systolic BP <120 mm Hg, as compared with <140 mm Hg, resulted in lower rates of fatal and nonfatal major cardiovascular events and death from any cause (5). However, significantly higher rates of some adverse events, including hypotension and kidney injury were observed in the intensive-treatment group (5). In another study, the Heart Outcomes Prevention Evaluation (HOPE)-3 trial randomly assigned 12,705 participants at intermediate risk who did not have cardiovascular disease to rosuvastatin or placebo and to candesartan plus hydrochlorothiazide or placebo, and showed that there was a clear benefit of antihypertensive therapy in persons with a systolic BP of ≥140 mm Hg but no benefit on cardiovascular events (as a composite) in those with an initial systolic BP of <140 mm Hg (6). SPRINT and (HOPE)-3 have implications for antihypertensive therapy and the cutoff of BP target in general.

The exact prevalence of resistant HTN is unknown, but likely increasing as older age, obstructive sleep apnea, and obesity, three of the strongest risk factors for resistant HTN, are becoming more prevalent. Obstructive sleep apnea (OSA) is the commonest and underappreciated cause of resistant HTN (7, 8). The Joint National Committee (JNC 8) (9, 10) did not address the definition of HTN. However, HTN is commonly classified for adults ≥18 years old into pre-hypertension, Stage 1 and Stage 2 (see Table 1).

Table 1. Classification of Hypertension

 

  Systolic (mm Hg) Diastolic (mm Hg)
Normal < 120 < 80
Pre-hypertension 120-139 80-89
Stage 1 hypertension 140-159 90-99
Stage 2 hypertension ≥ 160 ≥ 100

 

 

The definition, diagnostic method and therapeutic targets for HTN differ between several recently published international and national guidelines (see Table 2).

 

Table 2. Selected Guidelines for the Diagnosis of Hypertension

 

Guidelines

Definition, Diagnostic Method and Targets

 

2011 UK National Institute for Health and Clinical Excellence (NICE) (11)

 

•  ≥ 140/90 mm Hg and ABPM (or home BP) >135/85

•  Target: <140/90. ≥80 years old, <150/90

 

2013 European Society of Hypertension (12)

•  ≥140/90 on office ABPM (gold standard)

•  Ambulatory or home BPM for resistant HTN

•  Target: <140/90. ≥80 years old, SBP <140-150

 

2014 Joint National Committee (JNC 8) (9, 10)

•  Did not define HTN

•  Office ABPM favored

•  Ambulatory BPM is not mentioned

•  Target: <60 year, <140/90. ≥60 year, <150/90

 

2014 American Society of Hypertension - International Society of Hypertension (13)

•  ≥140/90 on office ABPM (gold standard)

•  Ambulatory or home BPM for resistant HTN

•  Target: <140/90. ≥80 years old, <150/90

2015 Canadian Hypertension Education Program (CHEP) (14)

 

•  Daytime mean ≥135/85, 24 hour mean 130/80 on ABPM. Home BPM mean ≥135/85. Mean office BP >180/110.

•  Target: <140/90. ≥80 years old, SBP ≤150

 

Abbreviations: ABPM, ambulatory blood pressure measurement; HTN, hypertension; SBP, systolic blood pressure.

Adapted from Stergiou et al. (15)

Several new terms and definitions are proposed to further classify HTN, which has added more confusion to clinicians. Refractory HTN, difficult to control HTN, controlled or uncontrolled resistant HTN with ≥3, ≥4, or ≥5 drugs, pseudo-resistant HTN, and apparent and true resistant HTN (16), are among the many others used in the literature. Common terminologies used in clinical practice are:

  • Resistant HTN: The diagnosis of resistant HTN should by definition rely on the inability of a well-constructed antihypertensive regimen to control BP (16). The American Heart Association defines resistant HTN as BP that remains above goal in spite of full doses of at least 3 antihypertensive medications, including a diuretic (17). This definition identifies patients who are at high risk of having reversible causes of HTN, such as primary aldosteronism. The definition also includes patients whose BP is controlled but require 4 or more medications to do so (“controlled resistant HTN”) and pseudo-resistant HTN. A lower cutoff for the diagnosis of resistant HTN (<130/80 mm Hg) could be used in patients with diabetes mellitus or kidney disease (i.e., with a creatinine level of >1.5 mg per deciliter [133 μmol per liter] or urinary protein excretion of >300 mg/24-hour period), despite adherence to treatment with full doses of at least 3 antihypertensive medications, including a diuretic (18).

 

  • Pseudo-resistant HTN: This should defined as an 'apparent' lack of BP control in spite of full doses of at least 3 antihypertensive medications, including a diuretic.

 

  • Masked HTN: This is defined as a normal BP in the office (<140/90 mm Hg), but an elevated BP out of the office (ambulatory daytime BP or home BP >135/85 mmHg) (19). Higher prevalence of masked HTN with poor cardiovascular score was found in both untreated subjects and treated hypertensives (20).

 

  • White-coat HTN: This is defined as subjects with office BP ≥140/90 mm Hg and a 24-hour BP <130/80 mm Hg (21).

 

  • Refractory HTN: This is defined as BP that remained uncontrolled in spite of maximal medical therapy after ≥3 visits to a hypertension clinic within a minimum 6-month follow-up period (22).

 

  • Isolated systolic HTN: This is defined as a systolic BP >140 mm Hg and diastolic BP <90 mm Hg, which is predominantly present in elderly patients, although not uncommonly in young and middle-aged adults (23).

 

  • Malignant/accelerated HTN: This is characterized by HTN with multi-organ involvement appearing over a short period of time, from a few weeks to a few months.

 

  • Sustained HTN: This is defined as BP >150/100 mm Hg on each of three measurements obtained on different days without antihypertensive therapy, necessating screening for secondary causes, including primary aldosteronism (24).

Several factors may induce episodes of resistance to therapy in treated hypertensive patients. Common factors include poor adherence to drug therapy, periodic changes in dietary factors (e.g.: high sodium intake, high alcohol consumption, natural licorice [glycyrrhiza glabra]), rapid weight gain, periodic changes in concomitant drugs (nonsteroidal anti-inflammatory agents [NSAIDs], selective COX-2 inhibitors, sympathomimetic agents [decongestants, diet pills, and cocaine], stimulants), weight loss medications, herbal compounds (e.g.: arnica, bitter orange, ephedra, ginkgo, and ginseng), and secondary forms of HTN (16).

The causes of HTN are broadly divided into primary (essential, due to an unclear etiology) and secondary. Secondary forms refer to an underlying, potentially correctable diagnosis, and are listed in Table 3. Approximately 10-20% of adults with HTN have a secondary form, although there is compelling evidence that the figure is likely higher (25). The prevalence of secondary forms of HTN is dependent on age, co-existing comorbidities, such as atherosclerosis, and the criteria used for screening. One study found the prevalence of secondary HTN was 10.2%, divided into renovascular HTN (3.1%), primary aldosteronism (1.4%), Cushing syndrome (0.5%), pheochromocytoma (0.3%), primary hypothyroidism (3.0%) and chronic kidney disease (defined as a serum creatinine > 2.0 mg/dl) (1.8%) (26). The presence of atherosclerosis significantly increased the prevalence of renovascular HTN (9.5%) and chronic kidney disease (8.0%) (26).

 

 

Table 3. Common and rare causes of Endocrine Hypertension

 

Common causes

 

Rare causes

 

·       Renal vascular hypertension

·       Primary aldosteronism

·       Hypothyroidism

·       Thyrotoxicosis

·       Hypercalcemia

·       Pheochromocytoma and paraganglioma

·       Cushing Syndrome

 

 

 

·       CAH: 11β-hydroxylase deficiency

·       CAH: 17α-hydroxylase deficiency

·       Familial hyperaldosteornism (e.g.: Type 1: Glucocorticoid-Remediable Aldosteronism)

·       Apparent mineralocorticoid excess

·       Liddle syndrome

·       Pseudohypaldosteronism Type 1

·       Pseudohypaldosteronism Type 2

·       Glucocorticoid Resistance syndrome

·       Growth hormone excess

·       Neuroendocrine tumors (“Carcinoid”)

·       Medullary thyroid cancer

·       Geller syndrome

 

Abbreviations: CAH, congenital adrenal hyperplasia.

 

 

CLINICAL FINDINGS AND OTHER CLUES TO IDENTIFY PATIENTS WITH ENDOCRINE HYPERTENSION

 

Several physical examination findings may point to the etiology for secondary forms of HTN. Table 4 provides a specific description of the clinical presentation of endocrine conditions related to HTN. Symptoms such as flushing and sweating may point toward pheochromocytoma and paraganglioma (27), while a renal bruit has been demonstrated in 87% of patients with renal artery stenosis (28). Laboratory abnormalities such as hypokalemia and metabolic alkalosis suggest primary aldosteronism due to increased renal hydrogen ion loss. Other clues that may point to the presence of endocrine forms of HTN are:

 

  • The onset of HTN in young individuals (< 40 years) or after the age of 50 years
  • Patients of African American descent that are at an increased risk for ARMC5 mutations leading to primary aldosteronism
  • Worsening HTN despite maximum drug treatment (failing triple-drug therapy including a diuretic) or controlled BP (<140/90 mm Hg) on four or more antihypertensive drugs
  • Sustained HTN, >150/100 mm Hg on each of three measurements obtained on different days
  • HTN and spontaneous or diuretic-induced hypokalemia
  • HTN and adrenal incidentaloma (bilateral or unilateral), defined as an asymptomatic adrenal mass detected on imaging not performed for suspected adrenal disease
  • HTN and obstructive sleep apnea
  • HTN and a family history of early-onset HTN or cerebrovascular accident at a young age (<40 years)
  • First-degree relatives of patients with endocrine HTN
  • Skin lesions (pheochromocytoma and paragangliomas [PPGLs], neurofibromatosis 1 [NF1], multiple endocrine neoplasia types 1 or 2 [MEN1, MEN2]): retinal angiomas, marfanoid body habitus, mucosal neuromas on eyelids/tongue, café-au-lait spots, axillary frecking, angiofibromas, or collagenomas
  • Worsening glycemic control and/or spinal osteoporosis (Cushing syndrome)
  • MEN2 with symptoms and signs of medullary thyroid carcinoma, PPGLs and/or primary hyperparathyroidism
  • Hemangioblastomas (brain, spinal cord, retina), endolymphatic sac tumors, renal cancer, pancreatic/renal/male genital tract cysts in von Hippel-Lindau (VHL) syndrome associated with PPGLs
  • Renal cancer associated with PPGLs in patients with germline mutations in SDHx
  • Gastrointestinal stromal tumors and PPGLs
  • Signs related to MEN1: primary hyperparathyroidism (hypercalcemia and/ or nephrolithiasis, osteoporosis), pituitary tumors (visual field defect, galactorrhea, impotence, headache) etc.

 

Table 4. Clinical Findings in Patients with Endocrine Hypertension

 

Condition

Clinical presentation

 

Renal Vascular Hypertension

Sudden increase in BP while controlled by medication, abdominal bruit (holosystolic, high-pitched) with radiation to the flanks

 

Primary Aldosteronism

Sustained, resistant or mild hypertension, muscle weakness, hypokalemia, metabolic alkalosis

 

Thyrotoxicosis

 

Isolated systolic hypertension, tremors, palpitations, atrial fibrillation, thyroid bruit, Graves’ extrathyroidal manifestations (orbitopathy, dermopathy, acropachy)

 

Cushing syndrome Fatigue, weight gain, round face,
proximal myopathy, skin thinning, ecchymosis, hirsutism, fat pads (supraclavicular, dorsocervical, temporal fossae), plethora, purple or red striae (>1cm)
Pheochromocytoma and Paraganglioma

Headache, palpitations, sweating, pallor, paroxysmal, resistant, mild hypertension, hypotension

 

CAH: 11β-hydroxylase deficiency

Classic form: virilization of   the external genitalia   in 46,XX newborn females, and precocious pseudopuberty in both sexes.

 

Nonclassic form: extremely rare, presents with hyperandrogenism during childhood

 

CAH: 17α-hydroxylase deficiency

Classic presentation: phenotypic female (46,XX or 46,XY) with hypertension, hypokalemia and   absence of secondary   sexual characteristics.

 

Partial 17-OHD:   46,XY with undervirilization and ambiguous   genitalia.

 

Familial Hyperaldosteronism

E.g.: Type I: Glucocorticoid-Remediable Aldosteronism (GRA)

Early onset of hypertension, presence of family history of mortality or morbidity from early hemorrhagic stroke

 

Liddle syndrome Severe hypertension, hypokalemia,
and metabolic alkalosis
Apparent Mineralocorticoid Excess Growth retardation, short stature,
hypokalemia
Pseudohypaldosteronism Type 2 Short stature, hyperkalemic metabolic
acidosis, normal aldosterone levels

Glucocorticoid Resistance Syndrome

 

Ambiguous genitalia, precocious
puberty, acne, hirsutism, oligo/anovulation, hypertension

 

 

COMMON CAUSES OF ENDOCRINE HYPERTENSION

Renal Vascular Hypertension (RVH)

RVH is one of the commonest types of secondary HTN that can be correctable with specific therapy. Most cases of RVH are caused by atherosclerotic renal artery stenosis (ARAS), fibromuscular dysplasia (FMD), vasculitis, thromboembolism and aneurysms. The renal arterial lumen must be decreased ≥ 70% to cause or worsen HTN. Most patients with RVH or renal artery stenosis (RAS) are >50 years old and have ARAS, while younger patients usually have renal artery FMD. Prevalence of RAS is estimated to be between 2% in unselected hypertensives and 40% in older patients with other atherosclerotic comorbidities (26, 29, 30). RVH is probably the third most common correctable cause of secondary HTN after OSA and primary aldosteronism. Table 5 lists the clinical and biochemical findings that should alert the clinician to screen for RVH:

 

 

Table 5. Clinical and biochemical findings associated with an increased possibility of RVH (31-34)

 

•  Moderate - severe or refractory HTN in patients with:

·       Multidrug (>3 agents) antihypertensive therapy

·       Recurrent episodes of flash pulmonary edema

·       Evidence of generalized vascular disease (CAD) or diffuse atherosclerosis

·       Unilateral small kidney (≤9 cm), or a difference in renal size of more than 1.5 cm that have no other explanation

•  Age/sex (young women <30 years old, think FMD, older men, think ARAS)

•  Sudden onset of severe HTN (≥160/100 mm Hg) after the age of 55 years

•  Elevation of serum creatinine ≥30% after administration of ACEi or ARB

•  Malignant HTN

•  Abdominal bruit (86% of cases)

 

Abbreviations: ACEi, angiotensin-converting enzyme inhibitor; ARAS, atherosclerotic renal artery stenosis; ARB, angiotensin II receptor blocker; CAD, coronary artery disease; FMD, fibromuscular dysplasia; HTN, hypertension.

 

The suggested work-up for RVH/RAS greatly depends on the degree of clinical suspicion. In general, a low index of suspicion does not require work-up. However, it is important to note that RVH is underdiagnosed and may be asymptomatic for several years. The American College of Cardiology/American Heart Association Guidelines (35) for screening of RVH/RAS recommends an invasive work-up when a corrective procedure is intended to be employed for clinically significant RVH. Individuals with moderate suspicion of RVH/RAS should undergo screening and confirmatory testing (see Figure 1).

 

The gold standard test for screening of RVS is renal arteriography. Noninvasive tests for diagnosing of RVH with a good sensitivity and specificity are Gadolinium enhanced magnetic resonance angiography (MRA), computed tomography angiography (CTA) and duplex ultrasonography. In presence of a positive screening test, renal arteriogram would help confirm the diagnosis and localize the site of the stenotic renal blood vessel. A schematic screening/diagnostic approach based on the degree of probability of clinical suspicion is described in Figure 1.

 

Plasma renin activity (PRA) as a screening test in patients with suspected RVH is underappreciated. PRA are within reference range in ~50% of patients with RVH while, conversely, increased levels may be found in ≤10% of patients with primary HTN (36-38). RVH leads to a low-pressure state within the afferent renal arterioles at the site of renin secretion, resulting in a log-unit (>10-fold) increase in plasma renin. Although no studies to date have evaluated the overall sensitivity and specificity of plasma renin as a diagnostic test in patients with RVH, the recognition of a plasma renin that is many-fold higher than the normal range, after accounting for false positives such as antihypertensive therapy, may be useful, particularly if the diagnosis was not appreciated on imaging (37). The sensitivity and specificity (39-47) of various tests in the workup of RVH/RAS are outlined in Table 6.

Figure 1. Flow chart for the diagnostic work-up of RVH.

Figure 1. Flow chart for the diagnostic work-up of RVH.

Table 6. Tests used in the diagnosis of RVH    
Test Sensitivity % Specificity %
Renal artery angiography 100 100
Intra-arterial digital subtraction angiogram (DSA) 94 97
Carbon dioxide digital angiography 83 99
Computed tomography angiography (CTA) 91 93
Gadolinium enhanced Magnetic resonance angiography (MRA) 96-100 71-96
Duplex ultrasonography 85 92
Captopril renography 57-94 44-98
Captopril test 15-68 76-93

The gold standard for confirming the diagnosis of RAS is a renal angiogram. Renal angiogram, intravenous subtraction angiography, intra-arterial digital subtraction angiography (DSA), or carbon dioxide angiography are imaging tests used in the diagnosis of RVH. Conventional aortography and intra-arterial digital subtraction angiogram (DSA) are considered the best tests offering high-quality radiographic images of the renal artery. Both, intra-arterial DSA (use less iodinated contrast ~25 mL) and carbon dioxide angiography are recommended in patients with deranged renal function. Intravenous DSA has a lower sensitivity and specificity compared with DSA and is falling out of favor. As an invasive procedure renal angiogram carry the risk of infections, cholesterol emboli and contrast-induced nephropathy.

Computed tomography angiography is a first line screening modality of RVH. Single breath-hold image detection and multidetector (MDCT) imaging have consistently improved the image acquisition and resolution with a better visualization of proximal/distal renal arteries. Spiral CT scan with angiography uses small amounts of IV iodinated contrast and is less invasive than arteriography. The use of iodinated compounds in subjects with poor renal function can result in contrast induced nephropathy. In patients with kidney failure the accuracy of this technique is impaired. Contrast allergies, anaphylaxis, and radiation exposure have to be considered in selected patients.

Gadolinium-enhanced MRA has one of the highest diagnostic performances for the detection of RAS among the non-invasive tests for RAS. The method is sensitive for imaging of proximal renal arteries but is suboptimal for significant distal lesions, accessory renal arteries and FMD (45). The use of 3-D gadolinium-enhanced MR (GEMR) imaging has improved the sensitivity and specificity of the method (>90%). Although MRA is considered as a first line investigation in patients with RVH, there are contraindications that limit its use in certain groups of patients: history of claustrophobia, metallic implant (pacemaker, surgical clips on vascular aneurysms), and pregnancy. The administration of gadolinium during MR imaging has caused nephrogenic systemic fibrosis in cases with acute/or chronic kidney disease and eGFR <30 mL/min (48, 49).

Duplex ultrasound is used as a screening test or for detection of recurrent stenosis of renal arteries in patients who underwent angioplasty/surgery. However, it is less sensitive in obese patients and the results are operator dependent (47). Using B-mode imaging/Doppler the operator can visualize main renal arteries and assess intrarenal pressures and velocities. Peak systolic velocity (>200 cm/sec.), renal-aortic ratio (≥3.5) and acceleration index (> .07 sec) are used to evaluate the presence of absence of a renal stenosis of >60% and end diastolic renal artery velocity of >150 cm/s for a stenosis of ≥80% (41, 44, 45).

Captopril renography is a noninvasive test to assess renal function. The administration of captopril orally (25-50 mg) 1 hour before the isotope injection increases the sensitivity of this test. Glomerular filtration can be estimated by measuring the excretion of Tc99m DTPA, Tc99m MAG3 or OIH (41, 46, 50). The test is considered positive (especially in unilateral RAS) when there is decreased relative uptake of the isotope with one kidney accounting for <40% of the total eGFR or a delayed peak uptake of the isotope of ≥10-11 minutes (50, 51). Limitation of the method include: creatinine ≥2 mg/dL and bilateral RAS.

Captopril test is used as a screening test. ~50% of patients with RVH have an increased PRA. PRA is measured before and 2 hours after oral administration of captopril 25 mg in seated position. Patients with RVH/RAS respond by increasing PRA >12 ng/ml/hr with an absolute increase >10 ng/ml/hr. Some antihypertensive therapy (ACEi, β-blockers, diuretics) should be stopped before testing. This method has a decreased sensitivity and specificity when compared with captopril renography, duplex ultrasound, CTA and MRA.

Cushing Syndrome

Endogenous Cushing syndrome (CS) is characterized by a constellation of signs and symptoms due to prolonged and high exposure of a variety of tissues to glucocorticoids. The incidence of endogenous CS is ~2-3 cases per 1 million inhabitants per year (52, 53). Endogenous CS is broadly divided into ACTH-dependent (~85% pituitary adenoma [Cushing disease], <5% ectopic ACTH secretion (54)) or ACTH-independent (~10-15% adrenal overproduction of glucocorticoids from an adrenocortical adenoma, hyperplasia or carcinoma). ACTH-independent causes of CS are classified on the basis of their radiographic and biochemical characteristics as being either functional or nonfunctional and benign or malignant. Approximately 75–90% of ACTH-independent causes of CS are due to  unilateral and benign cortisol-producing adenomas, with the remaining majority due to bilateral adrenocortical hyperplasias (BAH) (55, 56).

BAH are divided into micronodular (<1 cm in diameter), macrocronodular (>1cm in diameter) or non-nodular. Briefly, the micronodular subtypes are usually diagnosed in children and young adults, and are either pigmented (primary pigmented nodular adrenocortical disease [c-PPNAD]) as seen in familial cases in the context of Carney complex, or isolated (i-PPNAD) when nonsyndromic, and not pigmented (iMAD; isolated massive adrenocortical disease). The macronodular subtypes, which are usually diagnosed in adults > 50 years old, may be sporadic or familial. Primary bilateral macronodular adrenocortical hyperplasia (PBMAH) was first described in 1964 (57), and was previously called massive macronodular adrenocortical disease (MMAD), bilateral macronodular adrenal hyperplasia (BMAH), or ACTH-independent   macronodular   adrenocortical   hyperplasia   (AIMAH) (58). PBMAH may be syndromic, as seen with mutations in ARMC5, APC, MEN1, FH and the Carney triad, Carney-Stratakis syndrome, and hereditary nonpolyposis colorectal cancer (55, 58-61). Other subtypes of macronodular PBMAH include primary bimorphic adrenocortical disease (PBAD), as seen in McCune-Albright syndrome, and lesions with G-protein-coupled receptors that produce excess cortisol only in response to certain endogenous factors (e.g.: gastrointestinal inhibitory polypeptide, GIP), as seen   with   food-dependent   Cushing   syndrome   (FDCS).

CS is associated with HTN in ~80% of adult cases and ~50% of children (62-65); CS is more likely if the onset of HTN (among other signs or symptoms) is at a younger age. Several clinical features should be considered when evaluating individuals for the presence of this CS (62-64):

Patients with features very suggestive of hypercortisolism:

  • Abnormal fat distribution, particularly in the supraclavicular, dorsocervical and temporal fossae
  • Facial rounding with plethora
  • Proximal myopathy
  • Easy bruising
  • Wide (>1 cm) purple striae
  • Decreased growth rate with weight gain in children
  • Menstrual irregularities

 

Patients with unusual features for their age group:

  • HTN in young individuals or resistant to therapy
  • Adrenal incidentalomas
  • Metabolic syndrome X
  • Hypogonadotropic hypogonadism
  • Spontaneous fracture in a young individual
  • Kidney stones
  • Cyclicity in symptoms
  • Hypokalemia
  • Peripheral edema
  • Psychiatric comorbidities
  • Obesity
  • Impaired short term memory
  • Female balding
  • Type 2 diabetes
  • Unusual infections

 

Clinical characteristics of CS vary based on the duration and cyclicity of high cortisol level exposure. When cortisol levels are mildly/or intermittently increased (cyclical), clinicians face a diagnostic challenge. It is important to distinguish these symptoms/signs in order of their frequency: increased fatigue (sensitivity ~100%), generalized obesity (sensitivity ~51-90%), round face (sensitivity ~88-92%), plethora (sensitivity ~78-94%), HTN (sensitivity ~74-90%), weakness, especially of proximal muscles (sensitivity ~56-90%), thinness and fragility of skin (sensitivity ~84%), and hirsutism (sensitivity ~64-84%) (62, 65). Facial plethora is also one of the earliest described clinical features of CS (27, 66).

 

Common screening/diagnostic tests:

The initial screening test for CS should be based on the suitability for a given patient (see Figure 2). The tests recommended by the Clinical Practice Guidelines of the Endocrine Society (65) are: late-night salivary cortisol (LNSC, two measurements), 1-mg overnight dexamethasone suppression test (ODST), urine free cortisol (UFC; at least two measurements) and the longer low-dose DST (LLDST, 2 mg/d for 48 hours). A random serum cortisol or plasma ACTH levels, 8-mg DST, urinary 17-ketosteroids or the insulin tolerance test should not be used to screen for CS. Screening for aberrant expression of GPCRs in adrenocortical tumors and hyperplasia could be considered (67) in a select group of patients. The clinician should be aware of any current or recent use of oral, skin creams, rectal, inhaled, topical, herbal or injected glucocorticoids before biochemical testing to avoid false positives.

 

Assays differ widely in their accuracy, and should be chosen on the basis of their performance. Thus, knowledge of assay variability, functional limit of detection, precision and normal ranges should be carefully assessed to assist in the interpretation of the test results. Antibody-based immunoassays (RIA and ELISA) can cross-react with cortisol metabolites and synthetic glucocorticoids while structurally based assays (HPLC and LC-MS/MS) do not pose this problem and are the method of choice for detection of cortisol and/or other metabolites.

 

Late night salivary cortisol (LNSC) - Patient with CS have an impaired diurnal variation of cortisol. The loss of circadian rhythm with absence of a late-night cortisol nadir is a consistent biochemical abnormality in patients with CS (65, 68, 69). Since biologically active free cortisol in the blood is in equilibrium with cortisol in the saliva, then measurement of a late night salivary cortisol (LNSC) level by liquid chromatography–tandem mass spectrometry (LC-MS/MS) can be employed as a screening test for CS. 0.5 mL (minimum 0.2 mL) of saliva is necessary for the test. Basic instructions for collection includes: no food, smoking (ideally avoided on the day of testing), chewing tobacco/licorice (contains the 11β-hydroxysteroid dehydrogenase type 2 inhibitor glycyrrhizic acid) or fluids for 30 minutes to 1 hour prior to collection; avoid any activity that can cause gums to bleed, including brushing and flossing of teeth, or stress; the saliva should be collected 10 minutes after rinsing the mouth with water; the swab is placed under the tongue until well saturated approximately one minute; the specimen can be placed in room air for up to 5 days, and refrigerated for 7 days. Two saliva samples on two separate evenings between 2300 and 2400 hours should be collected because the hypercortisolism of CS can be variable, and this strategy increases confidence in the test results. Levels at midnight ≤0.09 mcg/dL (see questdiagnostics.com) are considered normal. The timing of the collection should be adjusted to the time of sleeping for shift workers or those with variable bedtimes. One study found that in men ≥60 years, 20% of all participants and 40% of diabetic hypertensive subjects had at least one elevated LNSC (70), which questions its utility as a screening test in this age group. LNSC is useful in detection early recurrence from CS in the postoperative period where urinary free cortisol and morning cortisol levels may be normal. If there is a normal diurnal rhythm (i.e. an appropriately low LNSC), then remission is likely (62, 64, 65). LNSC yields a 92–100% sensitivity and a 93–100% specificity for the diagnosis of CS (65).

 

1-mg Overnight Dexamethasone Suppression Test (ODST) - Patients with CS fail to suppress ACTH secretion from the pituitary gland when low doses of the synthetic glucocorticoid dexamethasone are given. This test entails administration of 1mg of dexamethasone at 2300 hours the night before a morning (0800 hours) blood sample for serum cortisol is drawn, simultaneously with a dexamethasone level (if feasible) to ensure adequate plasma concentrations [>5.6 nmol/L (0.22 μg/dL)] (71). Variable absorption and metabolism of dexamethasone may influence the result of both the 1-mg ODST and the longer low-dose DST (LDDST; 2 mg/d for 48 h). Patients should avoid eating or drinking for 12 hours before the morning blood test. Drugs such as phenytoin, phenobarbital, carbamazepine, rifampicin, and alcohol induce hepatic enzymatic clearance of dexamethasone, mediated through CYP3A4, thereby reducing the plasma dexamethasone concentrations leading to false positives (65). Dexamethasone clearance may be reduced in patients with liver and/or kidney failure. Interpretation of the serum cortisol has many caveats. The serum cortisol assay measures total cortisol, which is not an adequate representation of the biologically relevant free cortisol levels in conditions that cause cortisol binding globulin (CBG) deficiency (e.g.: nephrotic syndrome, cirrhosis, critical illness, postoperative period, CBG deficiency or malnourished states) or excess (e.g.: obesity, pregnancy, oral contraceptives, and estrogen   therapy). False-positives for ODST are seen in 50% of women taking oral contraceptives, and should be withdrawn for 6 weeks before testing or retesting (72). Certain conditions associated with abnormal cortisol levels need to be excluded: alcoholism, major depression, stress, thyrotoxicosis, poorly controlled diabetes mellitus, pregnancy or kidney failure. Morning cortisol levels >1.8 mcg/dL (50 nmol/L) are considered positive (65). If an increased specificity (95-100%) is sought, the longer LDDST (2 mg/d for 48 h) could be employed (73), or a higher serum cortisol threshold for the 1mg ODST is used (74). This is particularly useful in the evaluation of adrenal incidentalomas where a cutoff of >5 mcg/dL (137.95 nmol/L) increases specificity for the detection of autonomous cortisol secretion (75). Fast-acetylators of dexamethasone may have a false positive test with the 1 mg ODST, which can be overcome with the longer LDDST. DST is not the screening test of choice in pregnancy, epilepsy, and cyclic CS. DST is the test of choice in renal failure and in the evaluation of an adrenal incidentaloma for autonomous cortisol secretion (so called mild or subclinical CS).

 

Urine Free Cortisol (UFC) - Unlike serum cortisol, UFC provides an integrated assessment of cortisol secretion that is not bound to CBG over a 24-hour period. Therefore, UFC is not affected by conditions and medications that alter CBG. Two UFC samples should be collected, with the first morning void discarded so that the collection begins with an empty bladder, up to and including the first morning void on the second day (65). Patients should not drink excessive amounts of fluid and to avoid the use of any glucocorticoid preparations. Because the hypercortisolism of CS can be variable, at least two collections should be performed, which increases confidence in the test results (65). Values above the upper limit of normal for the particular assay is considered positive, provided the creatinine shows that the collection is complete and that the urine volume is not excessive (>5L) (65). Pseudo-Cushing syndrome is associated with false positive UFC’s and should be considered on the differential. UFC appears to be less sensitive than the 1 mg DST or LNSC for the identification of autonomous cortisol secretion in the setting of an adrenal incidentaloma (65). Upper limits of normal are much lower with HPLC or LC-MS/MS than in antibody-based assays (as low as 40% of the value measured by RIA) (76).

 

Plasma ACTH - A serum ACTH level could help narrow the differential diagnosis of hypercortisolemia (ACTH-dependent vs. independent) after the diagnosis has been established. Immunochemiluminometric assays detect intact ACTH; 1.5 ml frozen EDTA plasma (0.3 ml minimum) is collected between 7:00-10:00 a.m., transferred on ice, and centrifuged immediately after collection to separate plasma from cells. The reference range for 3-17 years is 9-57 pg/mL, and ≥18 years is 6-50 pg/mL (see questdiagnostics.com). Elevated levels are seen in ectopic ACTH and Cushing’s disease, unless cyclicity is present, while suppressed levels are seen in ACTH-independent causes, such as CS due to adrenocortical tumors and hyperplasia. An ectopic ACTH-secreting pheochromocytoma from the adrenal glands is an exception to the rule. False-positives are not uncommon, and could be from errors in sample transfer and processing, assay interference (e.g.: 5 mg/day of biotin or presence of monoclonal mouse antibodies), and stress.

 

Corticotropin-releasing hormone (CRH) stimulation test - This test is useful for differentiating between ACTH-dependent from ACTH-independent CS. Human and ovine CRH are commercially available and are given intravenously (bolus) at a dose of 1 µg/kg body weight. ACTH and cortisol levels are measured before (-5, 0 minutes) and after (15, 30, 45, 60, 90, and 120 minutes) the administration of CRH. Some studies suggested that the measurements of ACTH and cortisol before and after 15, 30 minutes and 45, 60 minutes, respectively, are sufficient to diagnose patients with ACTH-dependent CS (77, 78). A rise in cortisol >20% and ACTH >35% in comparison with baseline levels is diagnostic for Cushing disease, with a sensitivity of 93% and a specificity of 100% (77).

 

Inferior petrosal sinus sampling (IPSS) - Inferior petrosal sinus ACTH sampling after CRH stimulation is the best method available for the intra-pituitary localization of microadenomas causing Cushing’s disease. This test also helps distinguish Cushing’s disease from ectopic ACTH secretion, provided that the appropriate technique of blood sampling is used meticulously (79). Bilateral IPSS and simultaneous peripheral ACTH sampling at baseline, 3 and 10 minutes after intravenous administration of ovine CRH (1 µg/kg) is performed, and baseline and/or stimulated IPS-to-peripheral ACTH ratios are calculated. A post-IPS-to-peripheral ACTH ratio >2 is sufficient for diagnosing Cushing’s disease (80). In about 70-80% of the cases, a ratio of greater than 1.4 between the right and left inferior petrosal sinuses confirms the presence of a microadenoma (79, 81, 82). Anomalous venous drainage, abnormal venous anatomy, and lack of expertise can lead to false-negative IPSS results and thereby disease misclassification. Prolactin measurement during IPSS can improve diagnostic accuracy and decrease false negative results (83).

 

Imaging modalities in Cushing syndrome:

Pituitary MRI pre- and post-gadolinium enhancement - MRI is the modality of choice in the evaluation of the pituitary gland and surrounding tissues. MRI provides excellent anatomical tissue discrimination without exposure to ionizing radiation. Sagittal and coronal planes are considered the most accurate in evaluating the anatomy of the pituitary gland and other CNS structures. When Cushing disease is suspected, contrast-enhanced magnetic resonance imaging (MRI) is recommended. T1-weighted (T1W) sequences and/or spoiled gradient recalled acquisition (SPGR) techniques provide the best images of the sella. 95% of microadenomas appear hypointense with no post-gadolinium enhancement in relation with normal surrounding tissues on T1W sequences (84-86). Only ~60-80% of pituitary adenomas are detected and ~10% of healthy individuals have abnormal findings (incidentalomas) on MRI (87). Diffuse hyperplasia of ACTH-producing cells and small microadenomas may not be seen on conventional or enhanced MRIs. Other techniques (IPSS or integrated 18F-FDG PET/CT) may be employed to increase the odds of disease detection. Dynamic MRI may further increase the detection rate of pituitary microadenomas at the expense of specificity.

 

High-resolution chest, neck, and/or abdominal CT - This technique may detect tumors in ectopic or adrenocortical areas. Small lesions (<1 cm) could be missed (bronchial carcinoids, pancreatic neuroendocrine tumors). The sensitivity is lower than MRI (~50%) (88).

 

Nuclear imaging - These techniques include 111In-pentetreotide (OCT), 131I/123I-metaiodobenzylguanidine, 18F-fluoro-2-deoxyglucose-positron emission tomography (FDG-PET), 18F-fluorodopa-PET (F-DOPA-PET), 68Ga-DOTATATE-PET/CT or 68Ga-DOTATOC-PET/CT scan (68Gallium-SSTR-PET/CT), which may be used in select cases, primarily for the detection of ectopic ACTH tumors, which express surface receptors for somatostatin. These scans improve the sensitivity of conventional radiology when tumor site identification is difficult (89). 68Gallium-SSTR-PET/CT likely offers the highest sensitivity (89). One study found that cortisol-producing adenomas had a higher average FDG-PET SUVmax of 5.9 compared to nonfunctioning masses (average SUVmax 4.2) and aldosterone-producing adenoma (SUVmax 3.2), and an SUVmax cut-off of 5.33 had 50.0% sensitivity and 81.8% specificity in localizing a cortisol-producing adenoma (90). Thus, FDG-PET may aid in the characterization and prioritization of adrenocortical nodules for surgery, particularly in the setting of bilateral adrenocortical masses.

Figure 2. General diagnostic approach of Cushing syndrome (62, 65, 91)

Figure 2. General diagnostic approach of Cushing syndrome (62, 65, 91)

Primary Aldosteronism (PA)

Primary aldosteronism (PA) refers to a group of disorders that produce aldosterone in an unregulated fashion, leading to HTN, sodium retention, suppression of plasma renin, and increased potassium excretion (± hypokalemia) (24, 92). The causes of PA are bilateral nodular or non-nodular adrenocortical hyperplasia (BAH; ~60%), aldosterone-producing adenoma (APA; ~30%), familial hyperaldosteronism (FH; ~2-6%), primary bilateral macronodular adrenocortical hyperplasia (PBMAH; <1%), adrenocortical carcinoma (ACC; < 1%), and ectopic aldosterone production (extremely rare). PA due to an aldosterone-producing adenoma is also known as Conn’s syndrome, in recognition of Dr Jerome Conn (93).

The 2016 Endocrine Society Clinical Practice Guidelines (24) recognizes PA as a public health issue (94), advocating for universal screening. This is particularly important, as the incidence of PA is likely higher than traditional assumptions, estimated to affect >10% of hypertensives, both in general and in specialty settings (24, 95-98). The current guidelines have modified their screening approach, to recommend screening ~50% (99) of patients with HTN (see Table 7).

 

Table 7. Patients at Increased Risk of Developing Primary Aldosteronism

 

 

·       Sustained BP >150/100 mm Hg on each of three measurements obtained on different days

·       Hypertension (>140/90 mm Hg) resistant to three conventional antihypertensive drugs (including a diuretic)

·       Controlled BP (<140/90 mm Hg) on four or more antihypertensive drugs

·       Hypertension and spontaneous or diuretic-induced hypokalemia

·       Hypertension and adrenal incidentaloma*

·       Hypertension and obstructive sleep apnea syndrome

·       Hypertension and a family history of early onset hypertension

·       Hypertension and cerebrovascular accident at a young age (<40 years)

·       All hypertensive first-degree relatives of patients with primary aldosteronism

 

*An asymptomatic adrenal mass detected on imaging not performed for suspected adrenal disease

Screening tests for Primary Aldosteronism:

The following factors should be considered before screening patients for PA. Medications, advanced age, dietary sodium/potassium, and hypokalemia may affect the screening tests. Presence of kidney disease, RVH, malignant HTN or pregnancy should be ruled out to avoid false-positives or negatives. In pregnancy, increases in plasma progesterone or other steroids competitively inhibit the effects of excess aldosterone on its receptor, and may cause remission of PA (100). Lack of uniformity in screening criteria exist due to variability in screening protocols, assay methods, and individual factors, including medication use, age, sex, and presence of kidney disease (24). Although HTN and hypokalemia are suggestive of PA, the majority of patients are normokalemic (24, 98). However, an uncorrected hypokalemia may lead to a false negative screen for PA.

 

PA is screened with the plasma aldosterone concentration (PAC) to plasma renin activity (PRA) ratio (ARR). ARR is an easy, inexpensive, rapid and accurate means of screening for PA (96). After the patient has been up for at least 2 hours and seated for 5 (up to 15) minutes, a midmorning sample for PAC and PRA is collected and maintained at room temperature. The best correlation between PRA and PAC is achieved with low sodium intake, while the patient is in upright position. Although substantial variability in cutoff values for ARR exists, a minimum PAC of 15 ng/dL (410 pmol/L) and PRA of <1 ng/mL/h are used as screening criteria, with the most commonly adopted cutoff values for ARR is 30 (in conventional units; 90% sensitivity and 91% specificity). In the 2016 Guidelines (24), the need for further confirmatory testing in the setting of spontaneous hypokalemia, PRA below detection levels plus PAC >20 ng/dL (550 pmol/L), is not needed. However, ~ 36% of patients with PA have a PAC <15 ng/dL (101).

 

PAC and PRA are influenced by salt intake. Disproportionate elevation of PAC in relation to 24-hour urinary sodium excretion is usually seen in patients with PA. Suppressed PRA, for the level of the preceding day 24-hour urinary sodium excretion, is suggestive of PA. Individuals with PA tend to have a higher BP and lower serum potassium levels while on a high-sodium diet. A sodium restriction neutralized these differences (102). Thus, optimal screening for PA should occur under conditions of high sodium, as sodium restriction can significantly raise PRA, normalize ARR, and result in false interpretation of PA screening, particularly in the milder phenotypes of PA, where PRA is not as suppressed (103). It is important to note that a suppressed PRA does not differentiate between PA and low renin essential HTN or other secondary forms of HTN. In addition, other diseases such as excess mineralocorticoids other than aldosterone, apparent mineralocorticoid excess, and Liddle syndrome are associated with suppression of PRA.

There are situations where ARR may produce false negative or positive results. A quarter of patients with essential HTN have low or suppressed PRA, which may affect the interpretation of ARR. Drugs that interfere with renin or aldosterone measurements should be stopped at least two weeks prior to screening. The most common drugs that should be avoided during the screening process are: adrenergic blockers, central α-2 agonists (e.g.: clonidine and α-methyldopa), NSAIDs, K-wasting diuretics, K-sparing diuretics, ACEi, ARBs, dihydropyridine calcium channel blockers, and renin inhibitors. Drugs that have no/or limited influence on ARR are α-adrenergic blocker (prazosin, doxazosin, terazosin), hydralazine, verapamil, fosinopril, and atenolol. Mineralocorticoid antagonists (e.g. eplerenone and spironolactone) raise PAC and PRA, and should be withdrawn for at least 4-6 weeks, or longer, prior to testing (104). β-blockers lower PRA (105, 106) and produce false positive results (104). Calcium channel blockers could lower PAC, increase PRA (107), and produce false negative results (104), masking the diagnosis of PA (108). Similarly, ACEi and ARBs could produce false negative results through raising PRA (104, 109). Alpha-blockers and α-methyldopa when used for a short time during the work up of PA may not affect screening. The use of statin therapy among hypertensive and diabetic subjects was associated with lower aldosterone secretion in response to angiotensin II and a low-sodium diet (110). The 2016 Endocrine Society Clinical Practice Guidelines on PA (24) advocate for testing with interfering medications particularly in severe cases (111), to avoid delay in diagnosis, with the caveat that testing should be repeated if the results are inconclusive or difficult to interpret.

The lower limit of detection varies among different PRA assays and can have a dramatic effect on ARR. Since ARR is highly dependent on plasma renin, false positive PRA could be expected when PAC is low. Although PRA is convenient for estimating the biological activity of the renin system, it does not necessarily reflect its actual concentration. A direct renin concentration (DRC; mU/L, conversion factor from PRA to DRC is 12) is an alternative test that can confirm a low renin state. This could be particularly useful in low-renin HTN, as observed in African Americans (112). DRC and PRA are poorly correlated in the range where PRA is <1 ng/mL/h.

The post captopril ARR enhances the accuracy for diagnosing PA (98). A ratio greater than 35 has sensitivity and specificity of 100% and 67-91%, respectively, compared with 95.4% and 28.3%, respectively, at baseline in patients with PA (98). This test appears to be as sensitive as salt loading in confirming a diagnosis of PA (113). This test is performed by administering 25 mg of captopril orally, taken 2 hours before sampling (114).

Recently, a new overnight diagnostic test was developed to simplify screening for PA. The test consisted of the fludrocortisone-dexamethasone suppression test (FDST) and the new overnight diagnostic test (DCVT) using a combination of dexamethasone, captopril and valsartan (115). The estimated sensitivity and specificity were 91 and 100%, respectively, for the post-FDST ARR, whereas 98% and 89% and 100% and 82% for the post-DCVT ARR and post-DCVT autonomous aldosterone secretion, respectively, with selected cutoffs of 0.32 and 3 ng/dL, respectively (115). The diagnosis of PA was confirmed in 44/45 (98%) using this non-laborious approach.

Prior evidence from observational and intervention studies suggested a modifiable relationship between the renin-angiotensin-aldosterone system (RAAS) and parathyroid hormone levels in humans (116). Recent evidence suggests that higher serum aldosterone concentration is associated with higher serum parathyroid hormone concentration, which is decreased with the use of RAAS inhibitors (117). This relationship does not alter the current screening recommendations for PA.

Confirmatory Tests for Primary Aldosteronism:

The 2016 Endocrine Society Clinical Practice Guidelines on PA (24) recommends that patients with a positive ARR undergo one or more confirmatory tests to definitively confirm or exclude the diagnosis of PA. However, in the setting of spontaneous hypokalemia, PRA below detection levels, and PAC >20 ng/dL (550 pmol/L), proceeding directly to subtype classification with confirmatory testing is warranted. The following tests are used to confirm PA:

Oral sodium loading test - 24h urinary excretion of aldosterone is measured after 3 days of high salt intake (>200 meq/day, ~ 6 g/day). 24 hour urinary sodium and creatinine should be measured simultaneously to ensure high sodium intake and adequacy of urine collection. Failure of high salt to suppress urinary aldosterone excretion to <11 µg/24 hours is diagnostic for PA. This test has a sensitivity of 96% and specificity of 93% for PA (113, 118).

 

Fludrocortisone suppression test - This test is performed by administering fludrocortisone 0.1 mg orally every 6 hours or 0.2 mg orally every 12 hours and sodium chloride >200 mmol orally/day for 4 days. Failure to suppress upright PAC to <5 ng/dL by day 4 confirms the diagnosis of PA. Upright PRA should be suppressed to <1 ng/ml/h on day 4 of the test. Since hypokalemia inhibits aldosterone secretion, potassium chloride supplement should be given to keep plasma potassium levels close to or in the normal range. This test is considered the most sensitive test to diagnose PA.

 

Saline suppression test - This test is performed by measuring PAC in the supine position after intravenous administration of 500 mL/hour of 0.9% sodium chloride for 4 hours. Failure to suppress PAC <10 ng/dL at the end of this test confirms the diagnosis of PA (119, 120). This test is easy to perform on an outpatient basis.
Both the fludrocortisone and saline suppression tests are contraindicated in patients with severe HTN, congestive heart failure, advanced kidney disease, cardiac arrhythmia, or severe hypokalemia.

 

Captopril challenge test - This test is performed by administering 25-50 mg captopril orally in the seated or standing position for at least 1 hour. PAC and PRA are measured before and 1 hour after administration of captopril. If PAC >12 ng/dL or ARR >26, the test is considered positive (121).

 

Differentiating Between Aldosterone-Producing Adenoma (APA) and Bilateral Adrenal Hyperplasia (BAH)

Changes in PAC on upright posture - Patients with APA show no change or reduction in PAC on upright posture, unlike patients with BAH. This test is performed by measuring PAC in the supine position and after 4 hours of upright posture. ~ 70% of patients with BAH respond by increasing PAC by at least 50%.

Bilateral adrenal venous sampling (AVS) - When surgical treatment is feasible and desired by the patient, an experienced radiologist should use AVS to differentiate between APA and BAH (24). Younger patients (<35 years old) with spontaneous hypokalemia, marked aldosterone excess, and unilateral adrenocortical lesions may not need AVS before proceeding to unilateral adrenalectomy (24).

PAC and cortisol levels are measured in the inferior vena cava (IVC) and right and adrenal veins before and serially after intravenous injection of synthetic ACTH, cosyntropin 0.25 mg. ACTH stimulation improves cortisol gradients and aldosterone secretion, resulting in a reduction in the proportion of nondiagnostic studies (122). Moreover, ACTH stimulation significantly reduces bilateral aldosterone suppression (aldosterone/cortisol (A/C) ratios in the adrenal veins are bilaterally lower than that in the inferior vena cava) with a single AVS procedure (123). The purpose of measuring plasma cortisol is to confirm the site of the sampling catheter, by correcting for differences in dilution of adrenal with non-adrenal venous blood when assessing for lateralization. Plasma cortisol levels are much higher in adrenal veins than IVC. However, simultaneous autonomous overproduction of cortisol and aldosterone is increasingly recognized, particularly in BAH, and unilateral cortisol overproduction with contralateral suppression could confound the interpretation of AVS results. Thus, measuring plasma free metanephrine during AVS to calculate lateralization ratios may circumvent this problem (124, 125). Basal combined ratio during AVS carries the best sensitivity for the detection of AVS selectivity at all cutoff values (126).ACTH stimulation acutely stimulates aldosterone secretion and will help magnify the differences in PAC levels between the two adrenal glands. The A/C ratio of the involved to contralateral side provides the best diagnostic accuracy for determining if one adrenal is responsible for increased aldosterone production. With determination of bilateral selective samples, ratios (A/C on involved side)/(A/C of IVC) ≥1.1, or of (A/C involved side)/(A/C opposite side) ≥2 provide the best compromise of sensitivity and false positive rates for lateralization of the etiology of PA (127). Contralateral suppression, defined as A/C (adrenal)  ≤ A/C (peripheral) on the unaffected side, combined with a ratio ≥2 times peripheral on the affected side, correlates with good BP and biochemical outcomes from surgery, and could be used as a factor in deciding whether to offer surgery for treatment of PA (128). Moreover, in patients with PA, where the lateralization index is <4 on AVS, contralateral suppression of aldosterone is an accurate predictor of a unilateral source of aldosterone excess (129). Basal aldosterone contralateral suppression could predict residual hyperplasia and post-operative outcomes (130).In patients without the right AVS due to issues related to canulation or nonselective, a multinomial regression modeling can detect lateralization of aldosterone secretion in most patients and could eliminate the need for repeat AVS (131).

Current evidence supports the use of LC-MS/MS-based steroid profiling during AVS to achieve higher aldosterone lateralization ratios in patients with APAs than immunoassay (132). Moreover, LC-MS/MS enables multiple measures for discriminating unilateral from bilateral aldosterone excess, with potential use of peripheral plasma for subtype classification (132).

Unilateral adrenalectomy is beneficial in patients with a unilateral source of hyperaldosteronism and/or in some patients with apparent bilateral PA (133). Patients with PA that undergo unilateral adrenalectomy enjoy a higher quality of life scores than their medically treated counterparts (134). To identify the likelihood of complete resolution of HTN without further need of lifelong antihypertensive therapy following unilateral adrenalectomy, the Aldosteronoma Resolution Score could be calculated (135). This score accurately identifies individuals at low (≤1) or high (≥4) likelihood of complete resolution of HTN, based on four readily available predictors (2 or fewer antihypertensive medications, BMI ≤25 kg/m2, duration of HTN ≤6 years, and female sex) (135), and can help clinicians objectively inform patients of likely clinical outcomes before surgical intervention.

In 2014, a consensus was reached on several key issues in relation to AVS, including the selection and preparation of the patients, the procedure for its optimal performance, and the interpretation of its results for diagnostic purposes even in the most challenging cases (136). A recent study demonstrated that treatment of PA based on CT or AVS subtype classification did not show significant differences in intensity of antihypertensive medication or clinical benefits for patients after 1 year of follow-up, which challenges the current recommendation to perform AVS in all patients with PA (137).

Imaging Modalities Useful in the Evaluation of Adrenocortical Masses:

Ultrasonography - Although simple and economic, this imaging modality has a lower sensitivity in detecting adrenocortical masses than CT or MRI (138, 139). The sensitivity varies with the extent of the adrenocortical mass (65% for mass <3 cm, and up to 100% if >3 cm) (140). The role of ultrasonography in differentiating benign from malignant adrenocortical masses is limited (141).

 

Adrenocortical scintigraphy - This modality uses cholesterol based radioactive tracers and include 131iodine 6-β-iodomethylnorcholesterol (NP-59) and 75selenium-Se-6-selenomethyl-19-Norcholesterol (142). Concordant and discordant patterns of uptake may not be differentiated in lesions <2.0 cm in diameter (143, 144). Sensitivity (71%-100%) and specificity (50%-100%) range widely for differentiating benign from malignant tumors (143, 145).

 

CT and MRI - These conventional imaging modalities assist in the subtyping of the etiology of PA. High resolution CT and MRI of the adrenal glands have poor sensitivity in localizing small APAs (<5mm in diameter) (127, 146). CT of the adrenal glands analyzes contiguous 2–5 mm-thick CT slices on multiple sections using multidetector row protocols (147). CT and MRI can help determine whether an adrenocortical mass is an adrenocortical carcinoma and can also assess for local tumor invasion and metastatic disease (148, 149). A CT cut-off at 4.0 cm has a sensitivity of 93% (150), while an unenhanced CT density of ≤10 HU has a sensitivity of 96–100% and a specificity of 50–100% in differentiating benign from malignant tumors (151-155).

 

A systematic review of diagnostic procedures to differentiate unilateral from bilateral adrenocortical lesions in PA has found that CT/MRI misdiagnosed 37.8% of patients when diagnostic accuracy of AVS was used as a main criterion for diagnosing laterality of aldosterone secretion, suggesting that these imaging modalities may not be sufficient for a definitive diagnosis of PA (156). Enhanced CT assists in distinguishing between lesions that are lipid-rich (aldosterone-producing adenoma, cortisol-producing adenoma) and lipid-poor (eg: pheochromocytoma, adrenocortical carcinoma). Lipid-rich adenomas “washout’’ contrast faster. They can be differentiated by attenuation values or the percentage or relative percentage of washout as early as 5-15 min after enhancement if the unenhanced CT density is >10 HU (153). Lipid-rich and lipid-poor lesions have a relative percentage washout on delayed scans of >50% and <50%, respectively (157). One study demonstrated a washout value of 51% at 5 min and 70% at 15 min in benign lesions, with a sensitivity and specificity for the diagnosis of adrenocortical adenoma of ~ 96% at a threshold attenuation value of 37 HU on the 15-min delayed enhanced scan (153). MRI is as accurate in distinguishing lesions that are lipid-rich from lipid-poor. Chemical shift imaging MRI can sort out lipid-rich lesions with a sensitivity of 84–100% and a specificity of 92–100% (149, 158-160). Adenomas appear as hypo- or iso-intense on T1-weighted images, and hyper- or iso-intense on T2-weighted images (161). Combining adrenal imaging and AVS, the effective surgical cure rate for PA is 95.5%, with a poor (58.6%) accuracy of CT and MRI in detecting unilateral adrenal disease, although the performance was well in patients <35 years old (162).

 

18F-FDG PET - This modality has a sensitivity of 93-100%, and specificity of 80-100% in identifying malignant masses in the adrenal glands or elsewhere (163-167). However, some primary malignant tumors (necrotic, hemorrhagic) or those that are metastatic (>1 cm) may show a lower FDG uptake than the liver, leading to false-negatives (163, 164, 167). One study found that APAs had a SUVmax 3.2 (67), which may aid in the characterization and prioritization of adrenocortical nodules for surgery, particularly in the setting of bilateral adrenocortical masses.

 

PET-CT - This modality has a sensitivity of 98.5%-100%, and specificity of 92%-93.8% in detecting and differentiating between the various types of adrenocortical masses. When enhanced CT is added, the specificity is reached to 100% (168).

 

11C-metomidate PET- Metomidate-based tracers (bound to adrenal CYP11B enzymes) have been introduced in clinical practice recently. These techniques provide good visualization of adrenocortical lesions. This new investigation has been considered promising in differentiating between lesions of adrenocortical and non-adrenocortical origins (169, 170). 11C-metomidate PET-CT demonstrates a good sensitivity and specificity in the detection of APA (171, 172). Based on SUVmax, the specificity was as much as 100% (172). Therefore, 11C-metomidate PET-CT could be a useful noninvasive and rapid investigation to AVS in patients with adrenocortical tumors (153, (172), although this technique has low selectivity for CYP11B2 over CYP11B1.

 

18F-CDP2230 - This recently described modality combines nuclear imaging with a new agent that has a high selectivity for CYP11B2 over CYP11B1 with a favorable biodistribution for imaging CYP11B2 (173). 18F-CDP2230 could be a promising imaging agent for detecting unilateral subtypes of PA.

 

 

Familial Hyperaldosteronism (FH)

Familial aldosteronism (FH) represent a group of autosomal dominant (AD) disorders that is estimated to affect ~ 2-6% of all patients with PA. FH is classified into three major subtypes:

FH-I, also known as Glucocorticoid-Remediable Aldosteronism (GRA) is an AD disorder characterized by a chimeric fusion of CYP11B2 and CYP11B1 (8q24.3), rendering the aldosterone synthase hybrid gene to be under the regulation of ACTH rather than the renin-angiotensin system (174, 175). This rare monogenic form of HTN in humans with no gender predilection accounts for ~1% of PA. Increased production of aldosterone and hybrid steroids, such as 18-oxocortisol and 18-hydroxycortisol, which is suppressible to dexamethasone, is seen in GRA. Significant phenotypic and biochemical heterogeneity exist (176); males tend to have more severe HTN, and likely related to the degree of hybrid gene-induced aldosterone overproduction (177), while others may never develop HTN. Some patients may develop benign adrenocortical tumors (178). GRA   should   be   suspected   in   patients   with   early-onset HTN (<20 years) in the setting of a suppressed PRA, a family history of PA, or early cerebral hemorrhage (<40 years) from intracranial aneurysms or hemorrhagic strokes (179).

 

FH-II (7p22) represents the most common form of FH that typically affects adults. FH-II is characterized by PA due to BAH, APA, or both, which is not glucocorticoid remediable (180). FH-II is clinically indistinguishable from sporadic PA. The mutations that cause FH-II are unknown, but linkage analysis has mapped them to chromosome 7p22 (181-185).

 

FH-III is due to a gain-of-function heterozygous germline mutation in KCNJ5 (11q2) that increase constitutive and angiotensin II-induced aldosterone synthesis. FH-III presents earlier,   in childhood, with severe HTN and metabolic derangements. In FH-III, KCNJ5 is aberrantly co-expressed with CYP11B2 and in some cells with CYP11B1, which likely explains the abnormally high secretion rate of the hybrid steroid, 18-oxocortisol (186).

 

Recently, germline mutations in ARMC5 (16p11.2) have been implicated in PA. One study identified germline mutations across the entire ARMC5 gene in 39.3% of patients with APA (187). In addition to the germline mutations, a second somatic variant was required in AMRC5 to mediate tumorigenesis leading to polyclonal adrenocortical nodularity (60, 61, 188). Interestingly, all mutant APA’ s affected patients of African Americans decent (187), which may explain their increased predisposition to PA and/or HTN. These findings suggest that ARMC5 plays an important role in the development of APA or other adrenocortical tumors, and may represent a new subtype of FH.

 

Recently, germline mutations in CACNA1D, which codes for an L-type calcium channel, have been found in two cases with a syndrome of PA, seizures, and neurologic abnormalities (189, 190). A recent exome sequencing study identified a recurrent gain of function germline mutation in CACNA1H (a T-type calcium channel), in 5 unrelated families with early-onset PA and HTN (189, 191). These findings suggest that mutations in calcium channels play an important role in the development of PA, and may represent a new subtype of FH.

 

The 2016 Endocrine Society Clinical Practice Guidelines (24) recommends screening for GRA in patients diagnosed with PA and:

  • Onset of HTN <20 years of age
  • A family history of PA
  • Strokes or other early cerebrovascular complications at <40 years of age

 

Screening could begin at puberty and then at every 5 years interval but its utility needs further confirmation. Most patients with a clear diagnosis of GRA are severely hypertensive (177, 192). The biochemical profile of individuals with GRA is represented in Table 8.

 

Table 8. Laboratory confirmation of Glucocorticoid-Remediable Aldosteronism

 

Disease Laboratory profile
GRA

·       Hypokalemia (not always present), high urinary potassium

·       Suppressed PRA

·       PAC or urinary aldosterone is normal or mildly elevated

·       ARR >30. Plasma aldosterone fails to rise or falls during 2 hour of upright posture following overnight recumbency (193, 194)

·       Elevated 24 hour urinary and plasma levels of 18-oxocortisol and   18-hydroxycortisol (195). Increased urinary 18-hydroxycortisol/total cortisol metabolites ratio. 18-oxocortisol is 20-30 higher in GRA than APA (196, 197)

·       PAC <4 ng/dL after suppression with dexamethasone 0.5 mg PO every 6h for 2-4 days (LDDST) is diagnostic for GRA (197). However, some patients may fail to suppress. Aldosterone is markedly elevated in response to ACTH administration (198-200). Other studies have found some patients with PA but without the chimeric gene and suppression of PAC with dexamethasone treatment so that this test may be interpreted with caution in patients with possible GRA (201)

·       Genetic testing using long PCR-based methods for detecting the hybrid GRA gene (11β-hydroxylase gene/aldosterone synthase gene) is the gold standard test for diagnosis (100% sensitivity and specificity) (174, 202)

 

Some considerations have to be made regarding the biochemical profile of GRA:
PRA level is non-specific since ~20% of patients with essential HTN have a low or suppressed renin

The degree of HTN, hypokalemia, urinary 18-oxocortisol and 18-hydroxycortisol, suppressed PRA or elevated PAC cannot be used to identify patients with GRA as they lack specificity (occurring in the other subtypes of PA) (132)

The 1 mg ODST may result in false positives while the LDDST test could produce false negative results

Dexamethasone may also suppress aldosterone in patients with APA. However, since aldosterone secretion is autonomous, DST fails to suppress to very low levels

Although LDDST is highly sensitive and specific (>90%) for GRA, some patients may show initial suppression only to rise again by day 4 of treatment or fail to suppress PAC to <4 ng/dL (197)

The major drawback of LDDST is the need for multiple blood tests, requiring either hospitalization or repeated outpatient visits. Also, LDDST is difficult to perform in children

Genetic testing could be used as a screening test for newborns of affected parents. Placental tissue or cord blood could be used. A negative test eliminates the possibility of GRA diagnosis

The LDDST and elevated 24-hour urinary and plasma levels of 18-oxocortisol and 18-hydroxycortisol could be used to diagnose GRA with the caveat that APA and BAH may result in elevated values.

Modern diagnosis of GRA relies on identification of the CYP11B1/CYP11B2 chimeric gene

 

Pheochromocytoma And Paraganglioma (PPGL)

PPGLs are neuroendocrine tumors that arise from adrenal (~85%, pheochromocytoma) or extra-adrenal (~15%, paraganglioma) chromaffin cells (203). These cells continuously produce and release, in an unregulated fashion, metanephrine   and/or   normetanephrine   (“metanephrines”) from epinephrine and norepinephrine (“catecholamines”), respectively (204). The most frequent location of PPGLs is in the adrenal glands (150). The prevalence of PPGL in patients with HTN is ~0.3% (26). Testing for PPGLs should be performed in the following circumstances (205, 206):

 

  • Signs and symptoms of PPGL, in particular if paroxysmal and/or provoked by certain medications such as glucocorticoids
  • Adrenal incidentaloma, with or without HTN
  • Hereditary predisposition or syndromic features suggesting hereditary PPGL (MEN1, MEN2, VHL, NF1, SDHx (203, 207-209))
  • Previous history of PPGL

 

Sympathetic paragangliomas that arise from the sympathetic paravertebral ganglia of thorax, abdomen, and pelvis produce catecholamines and are rarely silent. Parasympathetic paragangliomas that arise from the glossopharyngeal and vagal nerves in the neck and at the base of the skull head and neck do not produce catecholamines, but rather dopamine (206, 210, 211). The classic triad of headaches 71%, sweating 65%, palpitations 65% and HTN has a 91% sensitivity, and 94% specificity, in diagnosing PPGL, although only seen in <30% of cases (212). Other symptoms that may point to the diagnosis of PPGLs are orthostatic hypotension (head and neck tumors that secrete dopamine) (10-50%), hyperglycemia (40%), weight loss (20-40%), flushing (10-20%), constipation, and pallor. Symptoms may be exacerbated by activity depending on the location of the PPGL, including urination (bladder PPGL), sexual intercourse or exercise (206, 212). Asymptomatic individuals with PPGL have been reported by various studies (206, 213). Normotensive PPGLs exist, and are often times identified as adrenal incidentalomas (incidence ~5%; 43% of patients had HTN) (150, 213). Patients with such tumors may not have elevated plasma or urinary fractionated metanephrines and may not need preoperative alpha blockade (214).

Biochemical screening for PPGL:

Figure 3 represents a general screening algorithm for PPGL. Measurement of plasma free or urinary fractionated   metanephrines   (“fractionated  metanephrines”) by LC-MS/MS is the most reliable and specific screening test for the diagnosis PPGLs (215-217). This assumes that the reference intervals have been appropriately established and measurement methods are accurate and precise (206, 215). Since the concentrations of normetanephrine sulfate and metanephrine-sulfate in plasma are about 20-30 fold higher than the levels of their free metabolites, the measurements of their deconjugated metabolites in plasma provides major advantages over their traditional measurements (218). Plasma free metanephrines reflect direct production by the tumor tissue and are considered the best test for excluding or confirming pheochromocytoma (204). Free metanephrine production is continuous and independent of catecholamine release (166). Thus, measurement of plasma free metanephrines is more reflective of the tumor production than catecholamines and normetanephrine levels.

Figure 3. Flow chart for diagnostic evaluation for Pheochromocytoma. Adapted after: Waguespack SG et al. (219). The criteria of malignancy is adapted from de Wailly et al. (211).

Figure 3. Flow chart for diagnostic evaluation for Pheochromocytoma. Adapted after: Waguespack SG et al. (219). The criteria of malignancy is adapted from de Wailly et al. (211).

A single collection of plasma free metanephrines (sensitivity 99%, specificity 89%) or urinary fractionated metanephrines (sensitivity 97%, specificity 69%) is equally recommended for screening, and combination of tests offers no advantage (206, 215, 217, 220). If urinary fractionated metanephrines are favored, then measurement of urinary creatinine   for verification of collection should be performed. Fractionated   metanephrines are superior to plasma catecholamines (sensitivity 84%, specificity 81%), urinary catecholamines (sensitivity 86%, specificity 88%) or urinary vanillylmandelic acid (sensitivity 64%, specificity 95%) (206, 217) (see Table 9). A spot urine sample should not be used for screening (206). Metanephrines can be detected in saliva with LC-MS/MS with sufficient sensitivity and precision but are not currently validated for screening (221). To avoid false positives, caffeine, smoking and alcohol intake should be withheld for 24 hours prior to testing, and   the blood sample should be drawn in lavender or green-top tube, transferred on ice, and then stored at -80°C until analyzed (206, 215). During testing, the patient should not be ill or under significant stress or strenuous physical activity.

 

Table 9. Sensitivity and specificity of biochemical tests for diagnosis of pheochromocytoma and paraganglioma (217)

 

Test Sensitivity (%)  Specificity (%)
Plasma free metanephrines 99 89
Plasma catecholamines 84 81
Urine fractionated metanephrines 97 69
Urine catecholamines 86 88
Urine total metanephrines 77 93
Urine vanillylmandelic acid 64 95

 

 

The highest diagnostic sensitivity for plasma free metanephrines is reached if the collection is performed in the supine position after an overnight fast and while the patient is recumbent in a quiet room for at least 20-30 minutes, and interpreted using the upper limit of the age-adjusted reference interval (206, 215, 222). If supine fasting sampling is performed, then the majority of patients with PPGLs can be recognized (215, 223). In centers where the supine position is not possible, then urinary fractionated metanephrines should be used for screening. A negative test result for plasma free metanephrines while seated is as effective for ruling out PPGL as negative results while supine (222). Measurement of seated serum metanephrines above the upper limit of the seated reference range has a diagnostic sensitivity and specificity of 93% and 90%, respectively (224). However, this approach can lead to a 5.7-fold increase in false-positives (223), necessitating repeat sampling.

 

Levels of fractionated metanephrines within the reference range usually exclude PPGLs (225), while equivocal results, which is seen in ~25% of all patients with PPGLs, require additional tests. Functional PPGL in patients with MEN2 and NF1 are characterized by increases in plasma free metanephrine, indicating epinephrine production, while VHL patients have increases in normetanephrine (indicating norepinephrine production) (226). Elevations in methoxytyramine (indicating dopamine production) are seen in ~70% of patients with SDHB and SDHD mutations (226). Exceptions for false-negative fractionated metanephrines exist, including PPGLs that are <1 cm in size, those that produce only dopamine (e.g. head and neck PPGLs), or those that are biochemically silent.

True- from False-positive Elevations of Fractionated Metanephrines

Levels of fractionated metanephrines >3 fold above the upper limit of the age-adjusted reference interval are rarely false-positives (227). The clonidine-suppression test is useful in distinguishing between true- from false-positive elevations in screening tests (228): 0.3 mg/70 kg body weight of clonidine hydrochloride is given orally, and plasma normetanephrine is measured at baseline and 3-hours after administration. A decrease in levels by >40%, or below the assay’s upper reference limit, suggests sympathetic activation, with a diagnostic specificity of 100% and a sensitivity of 87% in ruling out PPGLs with borderline elevations (227). Sensitivity increases to ~95% for elevations above borderline values (227). It is important to note that this test can only be used for PPGLs that secrete either norepinephrine or normetanephrine (229). The glucagon stimulation test should not be used in clinical practice given its insufficient diagnostic sensitivity for PPGL and potential to induce crisis (206, 230). There is insufficient evidence to use urine fractionated metanephrines and serum/plasma chromogranin A in conjunction for evaluation of borderline elevations in screening tests (231, 232).

 

Plasma norepinephrine and dopamine can be increased up to 3-fold in patients on hemodialysis without PPGL (233). Plasma free metanephrines and catecholamines are 2-3 fold higher in patients with renal failure compared with other groups (healthy normotensives, hypertensives and patients with VHL). However, plasma free metanephrines are relatively independent of renal function and are the test of choice in the diagnosis of PPGL among patients with renal failure (233) or those in the intensive-care unit. Urinary and plasma fractionated metanephrines have the highest sensitivity to diagnose PPGL in pregnancy (234).

 

In preparation for screening, discontinuation of all medications and substances that could interfere with the results should be performed (235) (see Figure 4.). The following medications raise   fractionated   metanephrines   and catecholamines: tricyclic antidepressants, selective norepinephrine reuptake inhibitors > selective serotonin reuptake inhibitors, monoamine oxidase inhibitors, cocaine, and α or β-blockers. The following medications have less or little influence on biochemical screening and can be continued during screening: selective α1-adrenoceptor blockers, diuretics, ACEi, and ARBs. The following medications may cause a direct analytical interference with the assays (not observed with LC-MS/MS), and should be withdrawn 5 days before testing: acetaminophen, labetalol, sotalol, buspirone, α-methyldopa,   and   5-aminosalicylic   acid   (mesalamine, and its prodrug, sulfasalazine). Withdrawal from the following can markedly elevate plasma or urinary fractionated metanephrines or catecholamines: sedatives, opioids, benzodiazepines, alcohol and smoking.

Figure 4. Mechanisms of Pharmacologic Interference with Catecholamines and Metanephrines (235). Monoamine oxidase- (MAO), dihydroxyphenylglycol- (DHPG), DOPA- dihydroxyphenylalanine. Adapted from Neary et al. (235)

Figure 4. Mechanisms of Pharmacologic Interference with Catecholamines and Metanephrines (235). Monoamine oxidase- (MAO), dihydroxyphenylglycol- (DHPG), DOPA- dihydroxyphenylalanine. Adapted from Neary et al. (235)

 

 

Imaging Modalities useful in the Evaluation of PPGL

CT and MRI - CT with contrast is very sensitive (88-100%) for localization of PPGLs that are >5 mm in diameter (236, 237). As with MRI, CT provides excellent topographical resolution but lacks specificity. PPGLs are homogeneous or heterogeneous, necrotic with some calcifications, solid, or cystic on CT. Most (>85%) PPGLs have an unenhanced attenuation of >10 HU on CT (154), although occasionally could have >60% washout on delayed imaging (238). A high signal intensity on T2-weighted MRI, referred to as a bright signal, is characteristic of PPGLs (239). The Endocrine Society Guidelines (206) recommend a CT scan of the abdomen and pelvis as the first radiographic test in evaluating PPGL, and is the preferred initial test in detecting lung metastasis. MRI is preferred in patients with metastatic PPGLs, for detection of skull base and neck paragangliomas, in patients with surgical clips causing artifacts when using CT, in patients with an allergy to CT contrast, and in patients in whom radiation exposure should be limited (children, pregnant women, patients with known germline mutations, and those with recent excessive radiation exposure) (206).

 

Functional imaging - 123I-metaidobenzylguandine (123I-MIBG) has a sensitivity and specificity of ~ 85% for pheochromocytomas and ~ 70% for paragangliomas, and is recommended as a functional imaging modality in patients with metastatic PPGLs detected by other imaging modalities when radiotherapy using 131I-MIBG is planned (206). Extra-adrenal and small adrenal pheochromocytomas are more likely to be under detected by 123I-MIBG and produce false negative results (240). Preparation with sodium perchlorate or potassium iodide is necessary 2 days before and 1 week after 131I-MIBG therapy to protect the thyroid gland (241). Some drugs that interfere with MIBG uptake such as labetalol, reserpine, digoxin, ACEi, various antidepressants/antipsychotics, and some of the sympathomimetics should be withdrawn 3-10 days before treatment (242). ~50% of normal adrenal glands demonstrate physiological uptake of 123I-MIBG, which may lead to false positives (243).
18F-FDA PET/CT is the preferred technique for the evaluation of patients with metastatic PPGLs. However, 68Ga-DOTA(0)-Tyr(3)-octreotate (68Ga-DOTATATE), a PET radiopharmaceutical with both high and selective affinity for somatostatin receptors, showed the highest lesion-based detection rate of 97.6 % when compared to other modalities in the localication of sporadic metastatic PPGLs (244, 245). 18F-FDG PET/CT, 18F-FDOPA PET/CT, 18F-FDA PET/CT, and CT/MRI showed detection rates of 49.2 %, 74.8 %, 77.7 %, and 81.6 %, respectively (244). Moreover, 68Ga-DOTATATE PET/CT was found to be the most sensitive technique in the detection of head and neck PPGLs, especially those that harbor mutations in SDHD, which may be very small and fail to concentrate sufficient 18F-FDOPA (246). Gene-targeted radiotherapeutics and nanobodies-based theranostic approaches are the future of imaging in PPGLs (247).

 

 

 

RARE CAUSES OF ENDOCRINE HYPERTENSION

Congenital Adrenal Hyperplasia: 11ß-Hydroxylase Deficiency

11β-hydroxylase deficiency   is the second   commonest variant of CAH (after 21-hydroxylase deficiency) with an incidence of 1 in 100,000–200,000 live births (248). This condition is caused   by mutations in   the 11β-hydroxylase gene   (CYP11B1) and is inherited in an autosomal recessive (AR) manner. The highest prevalence is seen in   Moroccan Jews, with an approximate incidence of 1 in   5000–7000 live births (249). A defective CYP11B1   enzyme leads to HTN from   elevated deoxycorticosterone (DOC) and   possibly other steroid   precursors, and hyperandrogenism from shunted precursors   into     the   androgen   synthesis     pathway (250). The classic form presents as   virilization of the   external genitalia in   46,XX newborn females, and precocious pseudopuberty in both sexes. The nonclassic form presents with hyperandrogenism during childhood. 11ß-hydroxylase deficiency is caused by several mutations in the CYP11B1 gene. Patients     do   not   present     with   adrenal insufficiency due to the glucocorticoid effects of excess   corticosterone. Biochemical profile includes variable hypokalemia, variable hyporeninemia, ↓↓ aldosterone, ↓cortisol, ↑ACTH, ↑11-deoxycortisol, ↑DOC, and ↑ 19-nor-DOC. Other laboratory abnormalities include elevated serum level of 17-hydroxyprogesterone (17-OHP), androstenedione and urinary pregnanetriol. See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Congenital Adrenal Hyperplasia: 17α-Hydroxylase Deficiency

17α-hydroxylase deficiency is an AR condition that results from a   defective CYP17A1 (248, 251, 252). This enzyme is responsible for catalyzing both 17-hydroxylase and 17,20 lyase activity. Thus, this condition results in glucocorticoid and   sex steroid deficiency,   and impairs   both   adrenal     and   gonadal   function. Certain ethnicities are at higher risk of 17α-hydroxylase deficiency, including Canadian Mennonites and Dutch Frieslanders, suggesting a founder effect (253). Biochemical features are ↑DOC, ↓11-deoxycortisol, ↓↓ aldosterone, ↓renin , ↓K, ↓plasma 17-OHP, and ↓testosterone. The   accumulation   of     the   mineralocorticoid precursors   corticosterone   and     DOC   exert   glucocorticoid   and mineralocorticoid   activity respectively and lead   to HTN with   hypokalemia. Adrenal   insufficiency is not   a characteristic feature of   this condition. The classic presentation of the severe form is a phenotypic female (46,XX or 46,XY) with HTN   and absence of   secondary sexual characteristics (254).   In the partial   form, 46,XY patients may   present with undervirilization and   present as infants   with ambiguous genitalia. Diagnosis of this autosomal recessive condition is suggested by delayed puberty, absent secondary sexual characteristics or amenorrhea combined with the typical biochemical findings (254). Genetic testing for mutations in the CYP17A1 gene confirms the condition. See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Apparent Mineralocorticoid Excess

The enzyme 11β-hydroxysteroid dehydrogenase type 2 (11βHSD2) is highly expressed in the kidneys where it metabolizes cortisol to cortisone to prevent the mineralocorticoid receptor (MCR) from inappropriate activation by cortisol (248, 255). Apparent mineralocorticoid excess (AME) is an AR condition that leads to reduced enzymatic activity of 11βHSD2 due to loss-of-function mutations or epigenetic changes in the HSD11B2 gene (256-258). The deficient gene alters the inactivation of cortisol in the target renal cells leading to activation of MCR, which leads to renal Na retention, severe hypokalemia from kaliuresis and HTN. The severity of HTN correlates with the degree of loss of enzymatic activity (259, 260). Clinical presentation of AME may include growth retardation, short stature, HTN, and hypokalemia that can lead to diabetes insipidus (261). The typical presentation of AME includes childhood-onset HTN with hypokalemia, suppressed renin, very low to undetectable aldosterone levels (hyporeninemic hypoaldosteronism) and metabolic alkalosis. Heterozygotes usually develop HTN later in life, without the phenotypic characteristics of AME (262). The biochemical diagnosis can be made by profiling of urinary steroid metabolites, which shows decreased cortisol inactivation, with the urinary tetrahydrocortisol and tetrahydrocortisone ratio (THF + 5αTHF)/THE and nearly absent urinary free cortisone (259, 263). AME is responsive to low sodium diet and spironolactone therapy (259). Genetic testing for AME-associated loss-of-function mutations in HSD11B2 confirms the diagnosis. See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Liddle Syndrome (Pseudohyperaldosteronism)

Liddle syndrome is a rare AD form of early-onset monogenic HTN with a prevalence of 1.52% in young hypertensives (264). The condition is caused by gain-of-function mutations in the genes (16p13) encoding β (SCNN1B) and γ (SCNN1G) subunits of the epithelial sodium channel (ENaC) (265-267). ENac is rate limiting for Na absorption in the aldosterone-sensitive distal nephron comprising the late distal convoluted tubule, the connecting tubule, and the entire collecting duct (268, 269). As a consequence, increased renal Na reabsorption with subsequent volume expansion and kaliuresis leads to severe HTN, hypokalemia and metabolic alkalosis (267). The typical biochemical profile is ↓K, ↑urinary K, ↓PRA, and suppressed aldosterone levels (hyporeninemic hypoaldosteronism). HTN usually responds to a combination of salt restriction (<100 mmol/day) and amiloride or triamterene therapy. Hyporeninemic hypoaldosteronism may present in the elderly population and mimic the biochemical patterns of Liddle syndrome (270). However, true Liddle syndrome may be underappreciated and undiagnosed in adults (271). Genetic testing can identify the disease mutations (266, 267). Screening of Liddle syndrome should be encouraged in young hypertensives, particularly those with early penetrance, hypokalemia, and low renin levels after exclusion of common secondary causes (264). See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Pseudohypoaldosteronism Type 2

Pseudohypoaldosteronism type 2 (PHA-2), or Gordon syndrome, is an AD condition that is caused by loss-of-function mutations in WNK1 or WNK4, which are part of a family of serine-threonine protein kinases. Mutations in WNK1 or WNK4 lead to an increased activity of the NaCl cotransporter in the distal tubule and consequently Na and fluid retention. More recently, mutations in the KLHL3, CUL3, and SPAK genes have been linked to Gordon syndrome (272, 273). Clinical features of patients with this syndrome include short stature, hyperchloremic metabolic acidosis, normal aldosterone levels and severe HTN. Biochemical profile includes ↑↑K, hyperchloremic metabolic acidosis, normal or ↓aldosterone, ↓PRA, ↓serum HCO3 (variable in children), and hypercalciuria (occasionally). The condition is confirmed by sequencing of WNK1 or WNK4, or the other rare genes implicated. See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Pseudohypoaldosteronism Type 1

Pseudohypoaldosteronism type 1 (PHA-1) is a rare AR form of monogenic HTN that is characterized by resistance to aldosterone. In affected patients, aldosterone levels are normal or elevated, but the renal response to aldosterone is disrupted due to functional abnormalities in either MCR (autosomal dominant or sporadic PHA-1; NR3C2 mutations) or the amiloride-sensitive ENaC (autosomal recessive PHA-1; SCNN1B mutations) (274, 275). Clinically, PHA-1 is characterized by Na wasting, failure to thrive, hyperkalemia, hypovolemia and metabolic acidosis (276). The diagnosis of PHA-1 may be missed until adulthood (274). See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

Constitutive Activation Of The Mineralocorticoid Receptor (Geller Syndrome)

 

Geller syndrome is an AD condition caused by gain-of-function mutations (4q31) in the gene encoding the MCR. One report described early-onset HTN due to a mutation in the MCR that was markedly exacerbated in pregnancy (277). The striking feature of this disorder is a severe exacerbation of HTN and hypokalemia during pregnancy due to the agonistic activity of progesterone and other mineralocorticoid antagonists on the MCR (278). The presence of HTN in males and non-pregnant females suggests that other functional mineralocorticoids are present (278). Biochemical profile includes ↑K, ↓aldosterone, and ↓PRA. See Tables 9-11 for laboratory tests and ratios provided by questdiagnostics.com.

 

 

OTHER POTENTIAL CAUSES (OR “BIOMARKERS”) OF ENDOCRINE HYPERTENSION

  • Insulin resistance without obesity
  • Obesity with and without insulin resistance
  • Growth hormone deficiency
  • Growth hormone excess
  • Testosterone deficiency (279)
  • Testosterone excess including polycystic ovarian syndrome
  • Thyrotoxicosis
  • Hypothyroidism
  • Primary hyperparathyroidism
  • Vitamin D deficiency (280)

 

COLLECTION OF SPECIMENS

Collection of specimens, special instructions and method used for laboratory tests commonly used in the diagnosis of endocrine hypertension (see Table 9 and 10)

 

Table 9. Tests/Code* Method Specimen Adult reference range**

Aldosterone 24-hour U

19552X

 

LC-MS/MS

·       5 ml refrigerated U

·       min 0.8 ml

·       2.3-21 µg/24-h

Aldosterone, serum

17181X

LC-MS/MS

·       Red-top tube

·       1 mL refrigerated

·       min 0.25 mL

·       Standing 8-10 AM: ≤28 ng/dL

·       Supine 8-10 AM: 3-16 ng/dl

 

Aldosterone / Plasma Renin Activity Ratio (ARR)
CPT code:  82088 CPTcode:  84244
LC-MS/MS

·       Lavender-top tube

·       1.8 mL frozen EDTA plasma

·       min 0.8 ml

·       The most commonly adopted cutoff is >30

CAH panel 1: 11-β Hydroxylase deficiency)

15269X

LC-MS/MS

·       No additives red top tube

·       0.6ml refrigerated

·       min 0.3 ml

·       11-Deoxycortisol/ Cortisol ratio >100

·       Androstenedione ↑

·       Testosterone ↑

CAH panel 3: Aldosterone synthase deficiency

15273X

RIA

LC-MS/MS

·       No additives red top tube

·       1.8ml refrigerated

·       min 0.8 ml

·       18-OH Corticosterone/Aldosterone ratio   >40

CAH panel 4 (females): 17-α Hydroxylase deficiency/

15274X

LC-MS/MS

·       No additives red top tube

·       1.2ml refrigerated

·       min 0.6 ml

·       Progesterone/17-OH Progesterone ratio   >6

·       Aldosterone ↓

·       Corticosterone ↑

·       Cortisol ↓

·       Estradiol ↓

CAH panel 8 (males): 17-α Hydroxylase deficiency

15279X

LC-MS/MS

·       No additives red top tube

·       0.8ml refrigerated

·       min 0.4 ml

·       Progesterone/17-OH Progesterone ratio   >6

·       Aldosterone ↓

·       Corticosterone ↑

·       Cortisol ↓

·       Testosterone ↓

Catecholamines fractioned 24-hour, urine

39627X

HPLC

·       10 ml room temp aliquot

·       Collect 25 ml U with

·       6N HCl/min 4.5 ml

·       Epinephrine: 2-24 µg/24h

·       Norepinephrine: 15-100 µg/24h

·       Total N+E: 26-121 µg/24h

·       Dopamine: 58-480 µg/24h

Catecholamines fractioned, plasma

314X

HPLC

·       4 ml sodium heparin

·       min 2.5 ml

·       Epinephrine: upright <95 pg/ml, supine <50 pg/ml

·       Norepinephrine: upright 217-1109 pg/ml, supine   112-658 pg/ml

·       Dopamine: upright<20 pg/ml, supine<10 pg/ml

·       Total N+E: upright 242-1125 pg/ml, supine 123-671 pg/ml

Corticosterone

6547X

LC-MS/MS

·       No additives red top tube

·       1 ml refrigerated

·       min 0.25 mL

·       8-10 AM 59-1293 ng/dL

Cortisol free, 24-hour urine

11280X

LC-MS/MS

·       2 ml frozen aliquot of 24-hour urine

·       min 0.5 ml

·       4-50 µg/24-hour

Deoxycorticosterone (DOC)

6559X

RIA

·       No additives red top tube

3 ml refrigerated S/

min 1.1 ml

·       Men:       3.5-11.5 ng/dL

·       Women follicular phase 1.5-8.5 ng/dL

·       luteal phase 3.5-13 ng/dL

 

*Available from Quest Diagnostics (www.questdiagnostics.com)

** Reference range for adults; P-plasma; U-urine; S-serum, RIA- extraction chromatography, radioimmunoassay; LC-MS/MS-liquid chromatography, tandem mass spectrometry

 

Table 10. Tests Special instructions*

Aldosterone 24-hour urine

 

·       Collect urine in 10 g of boric acid. Refrigerate during collection.

·       Record 24-h volume on vial and request form

Aldosterone

 

·       Draw upright blood half an hour after patient sits up; results vary with sodium excretion, electrolytic balance and posture (standing or recumbent)

 

Aldosterone/Plasma Renin Activity Ratio (ARR)

·       Do not refrigerate the specimen; refrigeration causes falsely-high PRA results.

·       Samples are collected in the morning after the patient has been out of bed for ≥ 2 hours and after sitting 5 -15 minutes.

·       Dietary salt intake should not be restricted, and potassium should be normalized if possible.

·       The patient can continue therapy with verapamil, hydralazine, prazosin hydrochloride, doxazosin mesylate, and terazosin hydrochloride during testing

CAH panel 1: 11-β Hydroxylase deficiency

CAH panel 3: Aldosterone synthase deficiency

CAH panel 4 (females): 17-α Hydroxylase deficiency

CAH panel 8 (males): 17-α Hydroxylase deficiency

 

·       Early morning specimen preferred (age, sex and time of specimen collection need to be specified)

 

Catecholamines fractioned, 24-hour urine

·       Urine must be collected with 25 mL 6 N Hydrochloric Acid.

·       It is advisable for the patients to be off medications 3 days prior to the test and avoid coffee, alcohol, tea, tobacco and strenuous exercise

Catecholamines fractioned, plasma

·       Collect in a pre-chilled vacutainer

·       Centrifuge in a refrigerated centrifuge within 30 minutes of collection

·       Separate plasma and freeze immediately

·       Avoid coffee, alcohol, tea, tobacco and strenuous exercise

·       Overnight fasting is necessary

Cortisol free, 24-hour urine ·       No preservatives preferred. However 25 mL 6 N hydrochloric acid or 10 g boric acid may be used
11-Deoxycortisol ·       Early morning specimen preferred (age, sex and time of specimen collection need to be specified)
Metanephrines, Fractionated plasma

·       Patient must refrain from using acetaminophen for 48 hours before testing.

·       Patient must refrain from using caffeine, medications, and tobacco, and from drinking coffee, tea or alcoholic beverages, for at least 4 hours before testing.

Plasma renin activity (PRA) ·       Collect and transport at room temperature. Centrifuge and freeze the plasma immediately.
Tetrahydroaldosterone 24-hour urine

·       Refrigerate during collection.

·       No use for preservatives

*Available from Quest Diagnostics ( www.questdiagnostics.com)

 

 

Table 11. The laboratory testing protocols for rare causes of endocrine hypertension

Disease Laboratory testing* Genetic testing

CAH: 17α-OH deficiency

 

·       ↑DOC, ↓11-deoxycortisol, ↓↓ aldosterone, ↓PRA, ↓K, ↓plasma 17-OHP, ↓testosterone,

·       ↑urinary100*THDOC/(THE+THF+5αTHF) and(THA+THB+5αTHB)/(THE+THF+5αTF)

 

CYP17A1

CAH: 11ß-OH deficiency

 

·       Hypokalemia (variable), ↓PRA, ↓↓ aldosterone, (the degree of hyporeninemia may vary widely), ↓cortisol, ↑ACTH, ↑11-deoxycortisol, ↑DOC, ↑ 19-nor-DOC

·       ↑↑urinary 100*THS/(THE+THF+5αTHF) and 100*THDOC/(THE+THF+5αTHF)

 

CYP11B1

 

Apparent mineralocorticoid excess

·       ↑ 24-hour urinary free cortisol/cortisone and ↑urinary (THF+5αTHF)/THE

·       check the level of tritiated water in plasma samples when 11-tritiated cortisol is injected

 

11ßHSD2

Liddle syndrome

 

·       ↓ K, ↑urinary K, ↓PRA, suppressed aldosterone secretion, metabolic acidosis

·       ↓urinary THALDO (<2 µg/24h), normal steroid profile (24-hour urine cortisone/cortisol and other ratios)

 

SCNN1B and   SCNN1G

Pseudohypo

aldosteronism

Type 2

·       ↑↑K, hyperchloremic metabolic acidosis, ↓aldosterone, ↓PRA, ↓serum HCO3 (variable in children), hypercalciuria (occasionally)

·       ↓urinary THALDO

 

WNK1 and WNK4

 

Rarely KLHL3, CUL3, and SPAK

 

Geller Syndrome

 

·       ↑K, ↓aldosterone, ↓PRA

·       ↓urinary THALDO

NR3C2

*Available from Quest Diagnostics ( www.questdiagnostics.com)

 

REFERENCES

  1. Nwankwo T, Yoon SS, Burt V, Gu Q. Hypertension among adults in the United States: National Health and Nutrition Examination Survey, 2011-2012. NCHS data brief. 2013(133):1-8.
  2. Fields LE, Burt VL, Cutler JA, Hughes J, Roccella EJ, Sorlie P. The burden of adult hypertension in the United States 1999 to 2000: a rising tide. Hypertension. 2004;44(4):398-404.
  3. Cutler JA, Sorlie PD, Wolz M, Thom T, Fields LE, Roccella EJ. Trends in hypertension prevalence, awareness, treatment, and control rates in United States adults between 1988-1994 and 1999-2004. Hypertension. 2008;52(5):818-27.
  4. Calhoun DA, Jones D, Textor S, Goff DC, Murphy TP, Toto RD, et al. Resistant hypertension: diagnosis, evaluation, and treatment. A scientific statement from the American Heart Association Professional Education Committee of the Council for High Blood Pressure Research. Hypertension. 2008;51(6):1403-19.
  5. Group SR, Wright JT, Jr., Williamson JD, Whelton PK, Snyder JK, Sink KM, et al. A Randomized Trial of Intensive versus Standard Blood-Pressure Control. N Engl J Med. 2015;373(22):2103-16.
  6. Yusuf S, Lonn E, Pais P, Bosch J, Lopez-Jaramillo P, Zhu J, et al. Blood-Pressure and Cholesterol Lowering in Persons without Cardiovascular Disease. N Engl J Med. 2016;374(21):2032-43.
  7. de Abreu-Silva EO, Beltrami-Moreira M. Sleep apnea: an underestimated cause of resistant hypertension. Current hypertension reviews. 2014;10(1):2-7.
  8. Goodfriend TL, Calhoun DA. Resistant hypertension, obesity, sleep apnea, and aldosterone: theory and therapy. Hypertension. 2004;43(3):518-24.
  9. James PA, Oparil S, Carter BL, Cushman WC, Dennison-Himmelfarb C, Handler J, et al. 2014 evidence-based guideline for the management of high blood pressure in adults: report from the panel members appointed to the Eighth Joint National Committee (JNC 8). JAMA. 2014;311(5):507-20.
  10. Reisin E, Harris RC, Rahman M. Commentary on the 2014 BP guidelines from the panel appointed to the Eighth Joint National Committee (JNC 8). Journal of the American Society of Nephrology : JASN. 2014;25(11):2419-24.
  11. Hypertension: The Clinical Management of Primary Hypertension in Adults: Update of Clinical Guidelines 18 and 34. National Institute for Health and Clinical Excellence: Guidance. London2011.
  12. Mancia G, Fagard R, Narkiewicz K, Redon J, Zanchetti A, Bohm M, et al. 2013 ESH/ESC Guidelines for the management of arterial hypertension: the Task Force for the management of arterial hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). J Hypertens. 2013;31(7):1281-357.
  13. Weber MA, Schiffrin EL, White WB, Mann S, Lindholm LH, Kenerson JG, et al. Clinical practice guidelines for the management of hypertension in the community: a statement by the American Society of Hypertension and the International Society of Hypertension. Journal of clinical hypertension. 2014;16(1):14-26.
  14. Daskalopoulou SS, Rabi DM, Zarnke KB, Dasgupta K, Nerenberg K, Cloutier L, et al. The 2015 Canadian Hypertension Education Program recommendations for blood pressure measurement, diagnosis, assessment of risk, prevention, and treatment of hypertension. The Canadian journal of cardiology. 2015;31(5):549-68.
  15. Stergiou GS, Ntineri A, Kollias A. Management of Masked Hypertension: Why Are We Still Sitting on the Fence? Hypertension. 2016.
  16. Burnier M. Resistant Hypertension: Is the Number of Drugs a Reliable Marker of Resistance? Hypertension. 2016.
  17. Calhoun DA, Jones D, Textor S, Goff DC, Murphy TP, Toto RD, et al. Resistant hypertension: diagnosis, evaluation, and treatment: a scientific statement from the American Heart Association Professional Education Committee of the Council for High Blood Pressure Research. Circulation. 2008;117(25):e510-26.
  18. Moser M, Setaro JF. Clinical practice. Resistant or difficult-to-control hypertension. N Engl J Med. 2006;355(4):385-92.
  19. Papadopoulos DP, Makris TK. Masked hypertension definition, impact, outcomes: a critical review. Journal of clinical hypertension. 2007;9(12):956-63.
  20. Bromfield SG, Shimbo D, Booth JN, 3rd, Correa A, Ogedegbe G, Carson AP, et al. Cardiovascular Risk Factors and Masked Hypertension: The Jackson Heart Study. Hypertension. 2016.
  21. Franklin SS, Thijs L, Hansen TW, O'Brien E, Staessen JA. White-coat hypertension: new insights from recent studies. Hypertension. 2013;62(6):982-7.
  22. Acelajado MC, Pisoni R, Dudenbostel T, Dell'Italia LJ, Cartmill F, Zhang B, et al. Refractory hypertension: definition, prevalence, and patient characteristics. Journal of clinical hypertension. 2012;14(1):7-12.
  23. Bavishi C, Goel S, Messerli FH. Isolated Systolic Hypertension: An Update After SPRINT. The American journal of medicine. 2016;129(12):1251-8.
  24. Funder JW, Carey RM, Mantero F, Murad MH, Reincke M, Shibata H, et al. The Management of Primary Aldosteronism: Case Detection, Diagnosis, and Treatment: An Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab. 2016;101(5):1889-916.
  25. Funder JW. Primary Aldosteronism: Seismic Shifts. J Clin Endocrinol Metab. 2015;100(8):2853-5.
  26. Anderson GH, Jr., Blakeman N, Streeten DH. The effect of age on prevalence of secondary forms of hypertension in 4429 consecutively referred patients. J Hypertens. 1994;12(5):609-15.
  27. Hannah-Shmouni F, Stratakis CA, Koch CA. Flushing in (neuro)endocrinology. Reviews in endocrine & metabolic disorders. 2016.
  28. Turnbull JM. The rational clinical examination. Is listening for abdominal bruits useful in the evaluation of hypertension? JAMA. 1995;274(16):1299-301.
  29. Berglund G, Andersson O, Wilhelmsen L. Prevalence of primary and secondary hypertension: studies in a random population sample. British medical journal. 1976;2(6035):554-6.
  30. Piecha G, Wiecek A, Januszewicz A. Epidemiology and optimal management in patients with renal artery stenosis. Journal of nephrology. 2012;25(6):872-8.
  31. Safian RD, Textor SC. Renal-artery stenosis. N Engl J Med. 2001;344(6):431-42.
  32. Senitko M, Fenves AZ. An update on renovascular hypertension. Current cardiology reports. 2005;7(6):405-11.
  33. Krumme B, Donauer J. Atherosclerotic renal artery stenosis and reconstruction. Kidney international. 2006;70(9):1543-7.
  34. Gandhi SK, Powers JC, Nomeir AM, Fowle K, Kitzman DW, Rankin KM, et al. The pathogenesis of acute pulmonary edema associated with hypertension. N Engl J Med. 2001;344(1):17-22.
  35. Hirsch AT, Haskal ZJ, Hertzer NR, Bakal CW, Creager MA, Halperin JL, et al. ACC/AHA 2005 guidelines for the management of patients with peripheral arterial disease (lower extremity, renal, mesenteric, and abdominal aortic): executive summary a collaborative report from the American Association for Vascular Surgery/Society for Vascular Surgery, Society for Cardiovascular Angiography and Interventions, Society for Vascular Medicine and Biology, Society of Interventional Radiology, and the ACC/AHA Task Force on Practice Guidelines (Writing Committee to Develop Guidelines for the Management of Patients With Peripheral Arterial Disease) endorsed by the American Association of Cardiovascular and Pulmonary Rehabilitation; National Heart, Lung, and Blood Institute; Society for Vascular Nursing; TransAtlantic Inter-Society Consensus; and Vascular Disease Foundation. J Am Coll Cardiol. 2006;47(6):1239-312.
  36. Derkx FH, Schalekamp MA. Renal artery stenosis and hypertension. Lancet. 1994;344(8917):237-9.
  37. Petruzzelli M, Taylor KP, Koo B, Brown MJ. Telling Tails: Very High Plasma Renin Levels Prompt the Diagnosis of Renal Artery Stenosis, Despite Initial Negative Imaging. Hypertension. 2016;68(1):11-6.
  38. Brown MJ. Clinical value of plasma renin estimation in the management of hypertension. American journal of hypertension. 2014;27(8):1013-6.
  39. Vasbinder GB, Nelemans PJ, Kessels AG, Kroon AA, de Leeuw PW, van Engelshoven JM. Diagnostic tests for renal artery stenosis in patients suspected of having renovascular hypertension: a meta-analysis. Annals of internal medicine. 2001;135(6):401-11.
  40. Schreier DZ, Weaver FA, Frankhouse J, Papanicolaou G, Shore E, Yellin AE, et al. A prospective study of carbon dioxide-digital subtraction vs standard contrast arteriography in the evaluation of the renal arteries. Archives of surgery. 1996;131(5):503-7; discussion 7-8.
  41. Amis ES, Jr., Bigongiari LR, Bluth EI, Bush WH, Jr., Choyke PL, Fritzsche P, et al. Radiologic investigation of patients with renovascular hypertension. American College of Radiology. ACR Appropriateness Criteria. Radiology. 2000;215 Suppl:663-70.
  42. Beregi JP, Elkohen M, Deklunder G, Artaud D, Coullet JM, Wattinne L. Helical CT angiography compared with arteriography in the detection of renal artery stenosis. AJR American journal of roentgenology. 1996;167(2):495-501.
  43. Postma CT, van Aalen J, de Boo T, Rosenbusch G, Thien T. Doppler ultrasound scanning in the detection of renal artery stenosis in hypertensive patients. The British journal of radiology. 1992;65(778):857-60.
  44. Radermacher J, Chavan A, Schaffer J, Stoess B, Vitzthum A, Kliem V, et al. Detection of significant renal artery stenosis with color Doppler sonography: combining extrarenal and intrarenal approaches to minimize technical failure. Clinical nephrology. 2000;53(5):333-43.
  45. De Cobelli F, Venturini M, Vanzulli A, Sironi S, Salvioni M, Angeli E, et al. Renal arterial stenosis: prospective comparison of color Doppler US and breath-hold, three-dimensional, dynamic, gadolinium-enhanced MR angiography. Radiology. 2000;214(2):373-80.
  46. Mann SJ, Pickering TG, Sos TA, Uzzo RG, Sarkar S, Friend K, et al. Captopril renography in the diagnosis of renal artery stenosis: accuracy and limitations. The American journal of medicine. 1991;90(1):30-40.
  47. Olin JW, Piedmonte MR, Young JR, DeAnna S, Grubb M, Childs MB. The utility of duplex ultrasound scanning of the renal arteries for diagnosing significant renal artery stenosis. Annals of internal medicine. 1995;122(11):833-8.
  48. Broome DR, Girguis MS, Baron PW, Cottrell AC, Kjellin I, Kirk GA. Gadodiamide-associated nephrogenic systemic fibrosis: why radiologists should be concerned. AJR American journal of roentgenology. 2007;188(2):586-92.
  49. Sadowski EA, Bennett LK, Chan MR, Wentland AL, Garrett AL, Garrett RW, et al. Nephrogenic systemic fibrosis: risk factors and incidence estimation. Radiology. 2007;243(1):148-57.
  50. Setaro JF, Chen CC, Hoffer PB, Black HR. Captopril renography in the diagnosis of renal artery stenosis and the prediction of improvement with revascularization. The Yale Vascular Center experience. American journal of hypertension. 1991;4(12 Pt 2):698S-705S.
  51. Pedersen EB. Angiotensin-converting enzyme inhibitor renography. Pathophysiological, diagnostic and therapeutic aspects in renal artery stenosis. Nephrology, dialysis, transplantation : official publication of the European Dialysis and Transplant Association - European Renal Association. 1994;9(5):482-92.
  52. Etxabe J, Vazquez JA. Morbidity and mortality in Cushing's disease: an epidemiological approach. Clinical endocrinology. 1994;40(4):479-84.
  53. Lindholm J, Juul S, Jorgensen JO, Astrup J, Bjerre P, Feldt-Rasmussen U, et al. Incidence and late prognosis of cushing's syndrome: a population-based study. J Clin Endocrinol Metab. 2001;86(1):117-23.
  54. Singer J, Werner F, Koch CA, Bartels M, Aigner T, Lincke T, et al. Ectopic Cushing's syndrome caused by a well differentiated ACTH-secreting neuroendocrine carcinoma of the ileum. Experimental and clinical endocrinology & diabetes : official journal, German Society of Endocrinology [and] German Diabetes Association. 2010;118(8):524-9.
  55. Stratakis CA, Boikos SA. Genetics of adrenal tumors associated with Cushing's syndrome: a new classification for bilateral adrenocortical hyperplasias. Nature clinical practice Endocrinology & metabolism. 2007;3(11):748-57.
  56. Lodish M, Stratakis CA. A genetic and molecular update on adrenocortical causes of Cushing syndrome. Nature reviews Endocrinology. 2016;12(5):255-62.
  57. Kirschner MA, Powell RD, Jr., Lipsett MB. Cushing's Syndrome: Nodular Cortical Hyperplasia of Adrenal Glands with Clinical and Pathological Features Suggesting Adrenocortical Tumor. J Clin Endocrinol Metab. 1964;24:947-55.
  58. Fragoso MC, Alencar GA, Lerario AM, Bourdeau I, Almeida MQ, Mendonca BB, et al. Genetics of primary macronodular adrenal hyperplasia. The Journal of endocrinology. 2015;224(1):R31-43.
  59. Alencar GA, Lerario AM, Nishi MY, Mariani BM, Almeida MQ, Tremblay J, et al. ARMC5 mutations are a frequent cause of primary macronodular adrenal Hyperplasia. J Clin Endocrinol Metab. 2014;99(8):E1501-9.
  60. Assie G, Libe R, Espiard S, Rizk-Rabin M, Guimier A, Luscap W, et al. ARMC5 mutations in macronodular adrenal hyperplasia with Cushing's syndrome. N Engl J Med. 2013;369(22):2105-14.
  61. Faucz FR, Zilbermint M, Lodish MB, Szarek E, Trivellin G, Sinaii N, et al. Macronodular adrenal hyperplasia due to mutations in an armadillo repeat containing 5 (ARMC5) gene: a clinical and genetic investigation. J Clin Endocrinol Metab. 2014;99(6):E1113-9.
  62. Stratakis CA. Diagnosis and Clinical Genetics of Cushing Syndrome in Pediatrics. Endocrinology and metabolism clinics of North America. 2016;45(2):311-28.
  63. Lacroix A, Feelders RA, Stratakis CA, Nieman LK. Cushing's syndrome. Lancet. 2015;386(9996):913-27.
  64. Nieman LK, Biller BM, Findling JW, Murad MH, Newell-Price J, Savage MO, et al. Treatment of Cushing's Syndrome: An Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab. 2015;100(8):2807-31.
  65. Nieman LK, Biller BM, Findling JW, Newell-Price J, Savage MO, Stewart PM, et al. The diagnosis of Cushing's syndrome: an Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab. 2008;93(5):1526-40.
  66. Afshari A, Ardeshirpour Y, Lodish MB, Gourgari E, Sinaii N, Keil M, et al. Facial Plethora: Modern Technology for Quantifying an Ancient Clinical Sign and Its Use in Cushing Syndrome. J Clin Endocrinol Metab. 2015;100(10):3928-33.
  67. El Ghorayeb N, Bourdeau I, Lacroix A. Multiple aberrant hormone receptors in Cushing's syndrome. Eur J Endocrinol. 2015;173(4):M45-60.
  68. Glass AR, Zavadil AP, 3rd, Halberg F, Cornelissen G, Schaaf M. Circadian rhythm of serum cortisol in Cushing's disease. J Clin Endocrinol Metab. 1984;59(1):161-5.
  69. Refetoff S, Van Cauter E, Fang VS, Laderman C, Graybeal ML, Landau RL. The effect of dexamethasone on the 24-hour profiles of adrenocorticotropin and cortisol in Cushing's syndrome. J Clin Endocrinol Metab. 1985;60(3):527-35.
  70. Liu H, Bravata DM, Cabaccan J, Raff H, Ryzen E. Elevated late-night salivary cortisol levels in elderly male type 2 diabetic veterans. Clinical endocrinology. 2005;63(6):642-9.
  71. Meikle AW. Dexamethasone suppression tests: usefulness of simultaneous measurement of plasma cortisol and dexamethasone. Clinical endocrinology. 1982;16(4):401-8.
  72. Qureshi AC, Bahri A, Breen LA, Barnes SC, Powrie JK, Thomas SM, et al. The influence of the route of oestrogen administration on serum levels of cortisol-binding globulin and total cortisol. Clinical endocrinology. 2007;66(5):632-5.
  73. Wood PJ, Barth JH, Freedman DB, Perry L, Sheridan B. Evidence for the low dose dexamethasone suppression test to screen for Cushing's syndrome--recommendations for a protocol for biochemistry laboratories. Annals of clinical biochemistry. 1997;34 ( Pt 3):222-9.
  74. Pecori Giraldi F, Ambrogio AG, De Martin M, Fatti LM, Scacchi M, Cavagnini F. Specificity of first-line tests for the diagnosis of Cushing's syndrome: assessment in a large series. J Clin Endocrinol Metab. 2007;92(11):4123-9.
  75. Fassnacht M, Arlt W, Bancos I, Dralle H, Newell-Price J, Sahdev A, et al. Management of adrenal incidentalomas: European Society of Endocrinology Clinical Practice Guideline in collaboration with the European Network for the Study of Adrenal Tumors. Eur J Endocrinol. 2016;175(2):G1-G34.
  76. Lin CL, Wu TJ, Machacek DA, Jiang NS, Kao PC. Urinary free cortisol and cortisone determined by high performance liquid chromatography in the diagnosis of Cushing's syndrome. J Clin Endocrinol Metab. 1997;82(1):151-5.
  77. Nieman LK, Oldfield EH, Wesley R, Chrousos GP, Loriaux DL, Cutler GB, Jr. A simplified morning ovine corticotropin-releasing hormone stimulation test for the differential diagnosis of adrenocorticotropin-dependent Cushing's syndrome. J Clin Endocrinol Metab. 1993;77(5):1308-12.
  78. Sheldon WR, Jr., DeBold CR, Evans WS, DeCherney GS, Jackson RV, Island DP, et al. Rapid sequential intravenous administration of four hypothalamic releasing hormones as a combined anterior pituitary function test in normal subjects. J Clin Endocrinol Metab. 1985;60(4):623-30.
  79. Landolt AM, Schubiger O, Maurer R, Girard J. The value of inferior petrosal sinus sampling in diagnosis and treatment of Cushing's disease. Clinical endocrinology. 1994;40(4):485-92.
  80. Findling JW, Kehoe ME, Shaker JL, Raff H. Routine inferior petrosal sinus sampling in the differential diagnosis of adrenocorticotropin (ACTH)-dependent Cushing's syndrome: early recognition of the occult ectopic ACTH syndrome. J Clin Endocrinol Metab. 1991;73(2):408-13.
  81. Oldfield EH, Doppman JL, Nieman LK, Chrousos GP, Miller DL, Katz DA, et al. Petrosal sinus sampling with and without corticotropin-releasing hormone for the differential diagnosis of Cushing's syndrome. N Engl J Med. 1991;325(13):897-905.
  82. Kaltsas GA, Giannulis MG, Newell-Price JD, Dacie JE, Thakkar C, Afshar F, et al. A critical analysis of the value of simultaneous inferior petrosal sinus sampling in Cushing's disease and the occult ectopic adrenocorticotropin syndrome. J Clin Endocrinol Metab. 1999;84(2):487-92.
  83. Sharma ST, Nieman LK. Is prolactin measurement of value during inferior petrosal sinus sampling in patients with adrenocorticotropic hormone-dependent Cushing's Syndrome? J Endocrinol Invest. 2013;36(11):1112-6.
  84. Invitti C, Pecori Giraldi F, de Martin M, Cavagnini F. Diagnosis and management of Cushing's syndrome: results of an Italian multicentre study. Study Group of the Italian Society of Endocrinology on the Pathophysiology of the Hypothalamic-Pituitary-Adrenal Axis. J Clin Endocrinol Metab. 1999;84(2):440-8.
  85. Doppman JL, Frank JA, Dwyer AJ, Oldfield EH, Miller DL, Nieman LK, et al. Gadolinium DTPA enhanced MR imaging of ACTH-secreting microadenomas of the pituitary gland. Journal of computer assisted tomography. 1988;12(5):728-35.
  86. Patronas N, Bulakbasi N, Stratakis CA, Lafferty A, Oldfield EH, Doppman J, et al. Spoiled gradient recalled acquisition in the steady state technique is superior to conventional postcontrast spin echo technique for magnetic resonance imaging detection of adrenocorticotropin-secreting pituitary tumors. J Clin Endocrinol Metab. 2003;88(4):1565-9.
  87. Hall WA, Luciano MG, Doppman JL, Patronas NJ, Oldfield EH. Pituitary magnetic resonance imaging in normal human volunteers: occult adenomas in the general population. Annals of internal medicine. 1994;120(10):817-20.
  88. Escourolle H, Abecassis JP, Bertagna X, Guilhaume B, Pariente D, Derome P, et al. Comparison of computerized tomography and magnetic resonance imaging for the examination of the pituitary gland in patients with Cushing's disease. Clinical endocrinology. 1993;39(3):307-13.
  89. Isidori AM, Sbardella E, Zatelli MC, Boschetti M, Vitale G, Colao A, et al. Conventional and Nuclear Medicine Imaging in Ectopic Cushing's Syndrome: A Systematic Review. J Clin Endocrinol Metab. 2015;100(9):3231-44.
  90. Patel D, Gara SK, Ellis RJ, Boufraqech M, Nilubol N, Millo C, et al. FDG PET/CT Scan and Functional Adrenal Tumors: A Pilot Study for Lateralization. World journal of surgery. 2016;40(3):683-9.
  91. Boscaro M, Arnaldi G. Approach to the patient with possible Cushing's syndrome. J Clin Endocrinol Metab. 2009;94(9):3121-31.
  92. McKenzie TJ, Lillegard JB, Young WF, Jr., Thompson GB. Aldosteronomas--state of the art. The Surgical clinics of North America. 2009;89(5):1241-53.
  93. Conn JW, Cohen EL, Rovner DR. Landmark article Oct 19, 1964: Suppression of plasma renin activity in primary aldosteronism. Distinguishing primary from secondary aldosteronism in hypertensive disease. By Jerome W. Conn, Edwin L. Cohen and David R. Rovner. JAMA. 1985;253(4):558-66.
  94. Funder JW. Primary aldosteronism as a public health issue. The lancet Diabetes & endocrinology. 2016;4(12):972-3.
  95. Hannemann A, Bidlingmaier M, Friedrich N, Manolopoulou J, Spyroglou A, Volzke H, et al. Screening for primary aldosteronism in hypertensive subjects: results from two German epidemiological studies. Eur J Endocrinol. 2012;167(1):7-15.
  96. Schwartz GL, Turner ST. Screening for primary aldosteronism in essential hypertension: diagnostic accuracy of the ratio of plasma aldosterone concentration to plasma renin activity. Clinical chemistry. 2005;51(2):386-94.
  97. Mulatero P, Stowasser M, Loh KC, Fardella CE, Gordon RD, Mosso L, et al. Increased diagnosis of primary aldosteronism, including surgically correctable forms, in centers from five continents. J Clin Endocrinol Metab. 2004;89(3):1045-50.
  98. Rossi E, Regolisti G, Negro A, Sani C, Davoli S, Perazzoli F. High prevalence of primary aldosteronism using postcaptopril plasma aldosterone to renin ratio as a screening test among Italian hypertensives. American journal of hypertension. 2002;15(10 Pt 1):896-902.
  99. Funder JW. Genetic disorders in primary aldosteronism-familial and somatic. The Journal of steroid biochemistry and molecular biology. 2017;165(Pt A):154-7.
  100. Gordon RD, Tunny TJ. Aldosterone-producing-adenoma (A-P-A): effect of pregnancy. Clinical and experimental hypertension Part A, Theory and practice. 1982;4(9-10):1685-93.
  101. Stowasser M, Gordon RD. Primary aldosteronism--careful investigation is essential and rewarding. Molecular and cellular endocrinology. 2004;217(1-2):33-9.
  102. Williams JS, Williams GH, Raji A, Jeunemaitre X, Brown NJ, Hopkins PN, et al. Prevalence of primary hyperaldosteronism in mild to moderate hypertension without hypokalaemia. Journal of human hypertension. 2006;20(2):129-36.
  103. Baudrand R, Guarda FJ, Torrey J, Williams G, Vaidya A. Dietary Sodium Restriction Increases the Risk of Misinterpreting Mild Cases of Primary Aldosteronism. J Clin Endocrinol Metab. 2016;101(11):3989-96.
  104. Mulatero P, Rabbia F, Milan A, Paglieri C, Morello F, Chiandussi L, et al. Drug effects on aldosterone/plasma renin activity ratio in primary aldosteronism. Hypertension. 2002;40(6):897-902.
  105. Buhler FR, Laragh JH, Baer L, Vaughan ED, Jr., Brunner HR. Propranolol inhibition of renin secretion. A specific approach to diagnosis and treatment of renin-dependent hypertensive diseases. N Engl J Med. 1972;287(24):1209-14.
  106. Gordon MS, Williams GH, Hollenberg NK. Renal and adrenal responsiveness to angiotensin II: influence of beta adrenergic blockade. Endocrine research. 1992;18(2):115-31.
  107. Cappuccio FP, Markandu ND, Sagnella GA, Singer DR, Buckley MG, Miller MA, et al. Effects of amlodipine on urinary sodium excretion, renin-angiotensin-aldosterone system, atrial natriuretic peptide and blood pressure in essential hypertension. Journal of human hypertension. 1991;5(2):115-9.
  108. Brown MJ, Hopper RV. Calcium-channel blockade can mask the diagnosis of Conn's syndrome. Postgraduate medical journal. 1999;75(882):235-6.
  109. Mantero F, Fallo F, Opocher G, Armanini D, Boscaro M, Scaroni C. Effect of angiotensin II and converting enzyme inhibitor (captopril) on blood pressure, plasma renin activity and aldosterone in primary aldosteronism. Clinical science. 1981;61 Suppl 7:289s-93s.
  110. Baudrand R, Pojoga LH, Vaidya A, Garza AE, Vohringer PA, Jeunemaitre X, et al. Statin Use and Adrenal Aldosterone Production in Hypertensive and Diabetic Subjects. Circulation. 2015;132(19):1825-33.
  111. Gallay BJ, Ahmad S, Xu L, Toivola B, Davidson RC. Screening for primary aldosteronism without discontinuing hypertensive medications: plasma aldosterone-renin ratio. American journal of kidney diseases : the official journal of the National Kidney Foundation. 2001;37(4):699-705.
  112. Campbell DJ, Nussberger J, Stowasser M, Danser AH, Morganti A, Frandsen E, et al. Activity assays and immunoassays for plasma Renin and prorenin: information provided and precautions necessary for accurate measurement. Clinical chemistry. 2009;55(5):867-77.
  113. Racine MC, Douville P, Lebel M. Functional tests for primary aldosteronism: value of captopril suppression. Curr Hypertens Rep. 2002;4(3):245-9.
  114. Castro OL, Yu X, Kem DC. Diagnostic value of the post-captopril test in primary aldosteronism. Hypertension. 2002;39(4):935-8.
  115. Tsiavos V, Markou A, Papanastasiou L, Kounadi T, Androulakis, II, Voulgaris N, et al. A new highly sensitive and specific overnight combined screening and diagnostic test for primary aldosteronism. Eur J Endocrinol. 2016;175(1):21-8.
  116. Brown JM, Williams JS, Luther JM, Garg R, Garza AE, Pojoga LH, et al. Human interventions to characterize novel relationships between the renin-angiotensin-aldosterone system and parathyroid hormone. Hypertension. 2014;63(2):273-80.
  117. Brown J, de Boer IH, Robinson-Cohen C, Siscovick DS, Kestenbaum B, Allison M, et al. Aldosterone, parathyroid hormone, and the use of renin-angiotensin-aldosterone system inhibitors: the multi-ethnic study of atherosclerosis. J Clin Endocrinol Metab. 2015;100(2):490-9.
  118. Agharazii M, Douville P, Grose JH, Lebel M. Captopril suppression versus salt loading in confirming primary aldosteronism. Hypertension. 2001;37(6):1440-3.
  119. Holland OB, Brown H, Kuhnert L, Fairchild C, Risk M, Gomez-Sanchez CE. Further evaluation of saline infusion for the diagnosis of primary aldosteronism. Hypertension. 1984;6(5):717-23.
  120. Litchfield WR, Dluhy RG. Primary aldosteronism. Endocrinology and metabolism clinics of North America. 1995;24(3):593-612.
  121. Rossi GP, Belfiore A, Bernini G, Desideri G, Fabris B, Ferri C, et al. Comparison of the captopril and the saline infusion test for excluding aldosterone-producing adenoma. Hypertension. 2007;50(2):424-31.
  122. Wolley MJ, Ahmed AH, Gordon RD, Stowasser M. Does ACTH improve the diagnostic performance of adrenal vein sampling for subtyping primary aldosteronism? Clinical endocrinology. 2016;85(5):703-9.
  123. Shibayama Y, Wada N, Umakoshi H, Ichijo T, Fujii Y, Kamemura K, et al. Bilateral aldosterone suppression and its resolution in adrenal vein sampling of patients with primary aldosteronism: analysis of data from the WAVES-J study. Clinical endocrinology. 2016;85(5):696-702.
  124. Goupil R, Wolley M, Ungerer J, McWhinney B, Mukai K, Naruse M, et al. Use of plasma metanephrine to aid adrenal venous sampling in combined aldosterone and cortisol over-secretion. Endocrinology, diabetes & metabolism case reports. 2015;2015:150075.
  125. Freel EM, Stanson AW, Thompson GB, Grant CS, Farley DR, Richards ML, et al. Adrenal venous sampling for catecholamines: a normal value study. J Clin Endocrinol Metab. 2010;95(3):1328-32.
  126. Mailhot JP, Traistaru M, Soulez G, Ladouceur M, Giroux MF, Gilbert P, et al. Adrenal Vein Sampling in Primary Aldosteronism: Sensitivity and Specificity of Basal Adrenal Vein to Peripheral Vein Cortisol and Aldosterone Ratios to Confirm Catheterization of the Adrenal Vein. Radiology. 2015;277(3):887-94.
  127. Rossi GP, Sacchetto A, Chiesura-Corona M, De Toni R, Gallina M, Feltrin GP, et al. Identification of the etiology of primary aldosteronism with adrenal vein sampling in patients with equivocal computed tomography and magnetic resonance findings: results in 104 consecutive cases. J Clin Endocrinol Metab. 2001;86(3):1083-90.
  128. Wolley M, Ahmed A, Gordon R, Stowasser M. 9b.04: Does Contralateral Suppression at Adrenal Venous Sampling Predict Outcome Following Unilateral Adrenalectomy for Primary Aldosteronism? A Retrospective Study. J Hypertens. 2015;33 Suppl 1:e121.
  129. Umakoshi H, Tanase-Nakao K, Wada N, Ichijo T, Sone M, Inagaki N, et al. Importance of contralateral aldosterone suppression during adrenal vein sampling in the subtype evaluation of primary aldosteronism. Clinical endocrinology. 2015;83(4):462-7.
  130. El Ghorayeb N, Mazzuco TL, Bourdeau I, Mailhot JP, Zhu PS, Therasse E, et al. Basal and Post-ACTH Aldosterone and Its Ratios Are Useful During Adrenal Vein Sampling in Primary Aldosteronism. J Clin Endocrinol Metab. 2016;101(4):1826-35.
  131. Durivage C, Blanchette R, Soulez G, Chagnon M, Gilbert P, Giroux MF, et al. Adrenal venous sampling in primary aldosteronism: multinomial regression modeling to detect aldosterone secretion lateralization when right adrenal sampling is missing. J Hypertens. 2016.
  132. Eisenhofer G, Dekkers T, Peitzsch M, Dietz AS, Bidlingmaier M, Treitl M, et al. Mass Spectrometry-Based Adrenal and Peripheral Venous Steroid Profiling for Subtyping Primary Aldosteronism. Clinical chemistry. 2016;62(3):514-24.
  133. Sukor N, Gordon RD, Ku YK, Jones M, Stowasser M. Role of unilateral adrenalectomy in bilateral primary aldosteronism: a 22-year single center experience. J Clin Endocrinol Metab. 2009;94(7):2437-45.
  134. Ahmed AH, Gordon RD, Sukor N, Pimenta E, Stowasser M. Quality of life in patients with bilateral primary aldosteronism before and during treatment with spironolactone and/or amiloride, including a comparison with our previously published results in those with unilateral disease treated surgically. J Clin Endocrinol Metab. 2011;96(9):2904-11.
  135. Zarnegar R, Young WF, Jr., Lee J, Sweet MP, Kebebew E, Farley DR, et al. The aldosteronoma resolution score: predicting complete resolution of hypertension after adrenalectomy for aldosteronoma. Annals of surgery. 2008;247(3):511-8.
  136. Rossi GP, Auchus RJ, Brown M, Lenders JW, Naruse M, Plouin PF, et al. An expert consensus statement on use of adrenal vein sampling for the subtyping of primary aldosteronism. Hypertension. 2014;63(1):151-60.
  137. Dekkers T, Prejbisz A, Kool LJ, Groenewoud HJ, Velema M, Spiering W, et al. Adrenal vein sampling versus CT scan to determine treatment in primary aldosteronism: an outcome-based randomised diagnostic trial. The lancet Diabetes & endocrinology. 2016;4(9):739-46.
  138. Suzuki K, Fujita K, Ushiyama T, Mugiya S, Kageyama S, Ishikawa A. Efficacy of an ultrasonic surgical system for laparoscopic adrenalectomy. The Journal of urology. 1995;154(2 Pt 1):484-6.
  139. Abrams HL, Siegelman SS, Adams DF, Sanders R, Finberg HJ, Hessel SJ, et al. Computed tomography versus ultrasound of the adrenal gland: a prospective study. Radiology. 1982;143(1):121-8.
  140. Suzuki Y, Sasagawa, Suzuki H, Izumi T, Kaneko H, Nakada T. The role of ultrasonography in the detection of adrenal masses: comparison with computed tomography and magnetic resonance imaging. International urology and nephrology. 2001;32(3):303-6.
  141. Fontana D, Porpiglia F, Destefanis P, Fiori C, Ali A, Terzolo M, et al. What is the role of ultrasonography in the follow-up of adrenal incidentalomas? The Gruppo Piemontese Incidentalomi Surrenalici. Urology. 1999;54(4):612-6.
  142. Rubello D, Bui C, Casara D, Gross MD, Fig LM, Shapiro B. Functional scintigraphy of the adrenal gland. Eur J Endocrinol. 2002;147(1):13-28.
  143. Falke TH, Sandler MP. Classification of silent adrenal masses: time to get practical. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 1994;35(7):1152-4.
  144. Gross MD, Shapiro B, Bouffard JA, Glazer GM, Francis IR, Wilton GP, et al. Distinguishing benign from malignant euadrenal masses. Annals of internal medicine. 1988;109(8):613-8.
  145. Gross MD, Shapiro B, Francis IR, Glazer GM, Bree RL, Arcomano MA, et al. Scintigraphic evaluation of clinically silent adrenal masses. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 1994;35(7):1145-52.
  146. Daunt N. Adrenal vein sampling: how to make it quick, easy, and successful. Radiographics : a review publication of the Radiological Society of North America, Inc. 2005;25 Suppl 1:S143-58.
  147. Blake MA, Kalra MK, Sweeney AT, Lucey BC, Maher MM, Sahani DV, et al. Distinguishing benign from malignant adrenal masses: multi-detector row CT protocol with 10-minute delay. Radiology. 2006;238(2):578-85.
  148. Young WF, Jr. Conventional imaging in adrenocortical carcinoma: update and perspectives. Hormones & cancer. 2011;2(6):341-7.
  149. McNicholas MM, Lee MJ, Mayo-Smith WW, Hahn PF, Boland GW, Mueller PR. An imaging algorithm for the differential diagnosis of adrenal adenomas and metastases. AJR American journal of roentgenology. 1995;165(6):1453-9.
  150. Mantero F, Terzolo M, Arnaldi G, Osella G, Masini AM, Ali A, et al. A survey on adrenal incidentaloma in Italy. Study Group on Adrenal Tumors of the Italian Society of Endocrinology. J Clin Endocrinol Metab. 2000;85(2):637-44.
  151. Hamrahian AH, Ioachimescu AG, Remer EM, Motta-Ramirez G, Bogabathina H, Levin HS, et al. Clinical utility of noncontrast computed tomography attenuation value (hounsfield units) to differentiate adrenal adenomas/hyperplasias from nonadenomas: Cleveland Clinic experience. J Clin Endocrinol Metab. 2005;90(2):871-7.
  152. Lee MJ, Hahn PF, Papanicolaou N, Egglin TK, Saini S, Mueller PR, et al. Benign and malignant adrenal masses: CT distinction with attenuation coefficients, size, and observer analysis. Radiology. 1991;179(2):415-8.
  153. Korobkin M, Brodeur FJ, Francis IR, Quint LE, Dunnick NR, Londy F. CT time-attenuation washout curves of adrenal adenomas and nonadenomas. AJR American journal of roentgenology. 1998;170(3):747-52.
  154. Caoili EM, Korobkin M, Francis IR, Cohan RH, Platt JF, Dunnick NR, et al. Adrenal masses: characterization with combined unenhanced and delayed enhanced CT. Radiology. 2002;222(3):629-33.
  155. Szolar DH, Kammerhuber FH. Adrenal adenomas and nonadenomas: assessment of washout at delayed contrast-enhanced CT. Radiology. 1998;207(2):369-75.
  156. Kempers MJ, Lenders JW, van Outheusden L, van der Wilt GJ, Schultze Kool LJ, Hermus AR, et al. Systematic review: diagnostic procedures to differentiate unilateral from bilateral adrenal abnormality in primary aldosteronism. Annals of internal medicine. 2009;151(5):329-37.
  157. Pena CS, Boland GW, Hahn PF, Lee MJ, Mueller PR. Characterization of indeterminate (lipid-poor) adrenal masses: use of washout characteristics at contrast-enhanced CT. Radiology. 2000;217(3):798-802.
  158. Korobkin M, Lombardi TJ, Aisen AM, Francis IR, Quint LE, Dunnick NR, et al. Characterization of adrenal masses with chemical shift and gadolinium-enhanced MR imaging. Radiology. 1995;197(2):411-8.
  159. Outwater EK, Siegelman ES, Radecki PD, Piccoli CW, Mitchell DG. Distinction between benign and malignant adrenal masses: value of T1-weighted chemical-shift MR imaging. AJR American journal of roentgenology. 1995;165(3):579-83.
  160. Bilbey JH, McLoughlin RF, Kurkjian PS, Wilkins GE, Chan NH, Schmidt N, et al. MR imaging of adrenal masses: value of chemical-shift imaging for distinguishing adenomas from other tumors. AJR American journal of roentgenology. 1995;164(3):637-42.
  161. Heinz-Peer G, Honigschnabl S, Schneider B, Niederle B, Kaserer K, Lechner G. Characterization of adrenal masses using MR imaging with histopathologic correlation. AJR American journal of roentgenology. 1999;173(1):15-22.
  162. Lim V, Guo Q, Grant CS, Thompson GB, Richards ML, Farley DR, et al. Accuracy of adrenal imaging and adrenal venous sampling in predicting surgical cure of primary aldosteronism. J Clin Endocrinol Metab. 2014;99(8):2712-9.
  163. Boland GW, Goldberg MA, Lee MJ, Mayo-Smith WW, Dixon J, McNicholas MM, et al. Indeterminate adrenal mass in patients with cancer: evaluation at PET with 2-[F-18]-fluoro-2-deoxy-D-glucose. Radiology. 1995;194(1):131-4.
  164. Erasmus JJ, Patz EF, Jr., McAdams HP, Murray JG, Herndon J, Coleman RE, et al. Evaluation of adrenal masses in patients with bronchogenic carcinoma using 18F-fluorodeoxyglucose positron emission tomography. AJR American journal of roentgenology. 1997;168(5):1357-60.
  165. Maurea S, Mainolfi C, Bazzicalupo L, Panico MR, Imparato C, Alfano B, et al. Imaging of adrenal tumors using FDG PET: comparison of benign and malignant lesions. AJR American journal of roentgenology. 1999;173(1):25-9.
  166. Yun M, Kim W, Alnafisi N, Lacorte L, Jang S, Alavi A. 18F-FDG PET in characterizing adrenal lesions detected on CT or MRI. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 2001;42(12):1795-9.
  167. Tenenbaum F, Groussin L, Foehrenbach H, Tissier F, Gouya H, Bertherat J, et al. 18F-fluorodeoxyglucose positron emission tomography as a diagnostic tool for malignancy of adrenocortical tumours? Preliminary results in 13 consecutive patients. Eur J Endocrinol. 2004;150(6):789-92.
  168. Blake MA, Slattery JM, Kalra MK, Halpern EF, Fischman AJ, Mueller PR, et al. Adrenal lesions: characterization with fused PET/CT image in patients with proved or suspected malignancy--initial experience. Radiology. 2006;238(3):970-7.
  169. Minn H, Salonen A, Friberg J, Roivainen A, Viljanen T, Langsjo J, et al. Imaging of adrenal incidentalomas with PET using (11)C-metomidate and (18)F-FDG. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 2004;45(6):972-9.
  170. Hennings J, Lindhe O, Bergstrom M, Langstrom B, Sundin A, Hellman P. [11C]metomidate positron emission tomography of adrenocortical tumors in correlation with histopathological findings. J Clin Endocrinol Metab. 2006;91(4):1410-4.
  171. Burton TJ, Mackenzie IS, Balan K, Koo B, Bird N, Soloviev DV, et al. Evaluation of the sensitivity and specificity of (11)C-metomidate positron emission tomography (PET)-CT for lateralizing aldosterone secretion by Conn's adenomas. J Clin Endocrinol Metab. 2012;97(1):100-9.
  172. Powlson AS, Gurnell M, Brown MJ. Nuclear imaging in the diagnosis of primary aldosteronism. Current opinion in endocrinology, diabetes, and obesity. 2015;22(3):150-6.
  173. Abe T, Naruse M, Young WF, Jr., Kobashi N, Doi Y, Izawa A, et al. A Novel CYP11B2-Specific Imaging Agent for Detection of Unilateral Subtypes of Primary Aldosteronism. J Clin Endocrinol Metab. 2016;101(3):1008-15.
  174. Lifton RP, Dluhy RG, Powers M, Rich GM, Cook S, Ulick S, et al. A chimaeric 11 beta-hydroxylase/aldosterone synthase gene causes glucocorticoid-remediable aldosteronism and human hypertension. Nature. 1992;355(6357):262-5.
  175. O'Mahony S, Burns A, Murnaghan DJ. Dexamethasone-suppressible hyperaldosteronism: a large new kindred. Journal of human hypertension. 1989;3(4):255-8.
  176. Dluhy RG, Anderson B, Harlin B, Ingelfinger J, Lifton R. Glucocorticoid-remediable aldosteronism is associated with severe hypertension in early childhood. The Journal of pediatrics. 2001;138(5):715-20.
  177. Stowasser M, Bachmann AW, Huggard PR, Rossetti TR, Gordon RD. Severity of hypertension in familial hyperaldosteronism type I: relationship to gender and degree of biochemical disturbance. J Clin Endocrinol Metab. 2000;85(6):2160-6.
  178. Jeunemaitre X, Charru A, Pascoe L, Guyene TT, Aupetit-Faisant B, Shackleton CH, et al. [Hyperaldosteronism sensitive to dexamethasone with adrenal adenoma. Clinical, biological and genetic study]. Presse medicale. 1995;24(27):1243-8.
  179. Litchfield WR, Anderson BF, Weiss RJ, Lifton RP, Dluhy RG. Intracranial aneurysm and hemorrhagic stroke in glucocorticoid-remediable aldosteronism. Hypertension. 1998;31(1 Pt 2):445-50.
  180. Torpy DJ, Gordon RD, Lin JP, Huggard PR, Taymans SE, Stowasser M, et al. Familial hyperaldosteronism type II: description of a large kindred and exclusion of the aldosterone synthase (CYP11B2) gene. J Clin Endocrinol Metab. 1998;83(9):3214-8.
  181. Carss KJ, Stowasser M, Gordon RD, O'Shaughnessy KM. Further study of chromosome 7p22 to identify the molecular basis of familial hyperaldosteronism type II. Journal of human hypertension. 2011;25(9):560-4.
  182. Sukor N, Mulatero P, Gordon RD, So A, Duffy D, Bertello C, et al. Further evidence for linkage of familial hyperaldosteronism type II at chromosome 7p22 in Italian as well as Australian and South American families. J Hypertens. 2008;26(8):1577-82.
  183. Jeske YW, So A, Kelemen L, Sukor N, Willys C, Bulmer B, et al. Examination of chromosome 7p22 candidate genes RBaK, PMS2 and GNA12 in familial hyperaldosteronism type II. Clinical and experimental pharmacology & physiology. 2008;35(4):380-5.
  184. So A, Duffy DL, Gordon RD, Jeske YW, Lin-Su K, New MI, et al. Familial hyperaldosteronism type II is linked to the chromosome 7p22 region but also shows predicted heterogeneity. J Hypertens. 2005;23(8):1477-84.
  185. Geller DS, Zhang J, Wisgerhof MV, Shackleton C, Kashgarian M, Lifton RP. A novel form of human mendelian hypertension featuring nonglucocorticoid-remediable aldosteronism. J Clin Endocrinol Metab. 2008;93(8):3117-23.
  186. Gomez-Sanchez CE, Qi X, Gomez-Sanchez EP, Sasano H, Bohlen MO, Wisgerhof M. Disordered zonal and cellular CYP11B2 enzyme expression in familial hyperaldosteronism type 3. Molecular and cellular endocrinology. 2017;439:74-80.
  187. Zilbermint M, Xekouki P, Faucz FR, Berthon A, Gkourogianni A, Schernthaner-Reiter MH, et al. Primary Aldosteronism and ARMC5 Variants. J Clin Endocrinol Metab. 2015;100(6):E900-9.
  188. Dutta RK, Soderkvist P, Gimm O. Genetics of primary hyperaldosteronism. Endocrine-related cancer. 2016;23(10):R437-54.
  189. Korah HE, Scholl UI. An Update on Familial Hyperaldosteronism. Hormone and metabolic research = Hormon- und Stoffwechselforschung = Hormones et metabolisme. 2015;47(13):941-6.
  190. Scholl UI, Goh G, Stolting G, de Oliveira RC, Choi M, Overton JD, et al. Somatic and germline CACNA1D calcium channel mutations in aldosterone-producing adenomas and primary aldosteronism. Nature genetics. 2013;45(9):1050-4.
  191. Scholl UI, Stolting G, Nelson-Williams C, Vichot AA, Choi M, Loring E, et al. Recurrent gain of function mutation in calcium channel CACNA1H causes early-onset hypertension with primary aldosteronism. eLife. 2015;4:e06315.
  192. Rich GM, Ulick S, Cook S, Wang JZ, Lifton RP, Dluhy RG. Glucocorticoid-remediable aldosteronism in a large kindred: clinical spectrum and diagnosis using a characteristic biochemical phenotype. Annals of internal medicine. 1992;116(10):813-20.
  193. Ganguly A, Grim CE, Weinberger MH. Anomalous postural aldosterone response in glucocorticoid-suppressible hyperaldosteronism. N Engl J Med. 1981;305(17):991-3.
  194. Stowasser M, Gordon RD. Familial hyperaldosteronism. The Journal of steroid biochemistry and molecular biology. 2001;78(3):215-29.
  195. Ulick S, Chu MD. Hypersecretion of a new corticosteroid, 18-hydroxycortisol in two types of adrenocortical hypertension. Clinical and experimental hypertension Part A, Theory and practice. 1982;4(9-10):1771-7.
  196. Ulick S, Chu MD, Land M. Biosynthesis of 18-oxocortisol by aldosterone-producing adrenal tissue. The Journal of biological chemistry. 1983;258(9):5498-502.
  197. Litchfield WR, New MI, Coolidge C, Lifton RP, Dluhy RG. Evaluation of the dexamethasone suppression test for the diagnosis of glucocorticoid-remediable aldosteronism. J Clin Endocrinol Metab. 1997;82(11):3570-3.
  198. New MI, Peterson RE, Saenger P, Levine LS. Evidence for an unidentified ACTH-induced steroid hormone causing hypertension. J Clin Endocrinol Metab. 1976;43(6):1283-93.
  199. Jamieson A, Inglis GC, Campbell M, Fraser R, Connell JM. Rapid diagnosis of glucocorticoid suppressible hyperaldosteronism in infants and adolescents. Archives of disease in childhood. 1994;71(1):40-3.
  200. Stowasser M, Gordon RD. Primary aldosteronism: learning from the study of familial varieties. J Hypertens. 2000;18(9):1165-76.
  201. Fardella CE, Pinto M, Mosso L, Gomez-Sanchez C, Jalil J, Montero J. Genetic study of patients with dexamethasone-suppressible aldosteronism without the chimeric CYP11B1/CYP11B2 gene. J Clin Endocrinol Metab. 2001;86(10):4805-7.
  202. Stowasser M, Gartside MG, Gordon RD. A PCR-based method of screening individuals of all ages, from neonates to the elderly, for familial hyperaldosteronism type I. Australian and New Zealand journal of medicine. 1997;27(6):685-90.
  203. Jochmanova I, Pacak K. Pheochromocytoma: The First Metabolic Endocrine Cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2016;22(20):5001-11.
  204. Eisenhofer G, Keiser H, Friberg P, Mezey E, Huynh TT, Hiremagalur B, et al. Plasma metanephrines are markers of pheochromocytoma produced by catechol-O-methyltransferase within tumors. J Clin Endocrinol Metab. 1998;83(6):2175-85.
  205. Plouin PF, Amar L, Dekkers OM, Fassnacht M, Gimenez-Roqueplo AP, Lenders JW, et al. European Society of Endocrinology Clinical Practice Guideline for long-term follow-up of patients operated on for a phaeochromocytoma or a paraganglioma. Eur J Endocrinol. 2016;174(5):G1-G10.
  206. Lenders JW, Duh QY, Eisenhofer G, Gimenez-Roqueplo AP, Grebe SK, Murad MH, et al. Pheochromocytoma and paraganglioma: an endocrine society clinical practice guideline. J Clin Endocrinol Metab. 2014;99(6):1915-42.
  207. Majumdar S, Friedrich CA, Koch CA, Megason GC, Fratkin JD, Moll GW. Compound heterozygous mutation with a novel splice donor region DNA sequence variant in the succinate dehydrogenase subunit B gene in malignant paraganglioma. Pediatric blood & cancer. 2010;54(3):473-5.
  208. Moramarco J, El Ghorayeb N, Dumas N, Nolet S, Boulanger L, Burnichon N, et al. Pheochromocytomas are diagnosed incidentally and at older age in neurofibromatosis type 1. Clinical endocrinology. 2016.
  209. Aufforth RD, Ramakant P, Sadowski SM, Mehta A, Trebska-McGowan K, Nilubol N, et al. Pheochromocytoma Screening Initiation and Frequency in von Hippel-Lindau Syndrome. J Clin Endocrinol Metab. 2015;100(12):4498-504.
  210. Varoquaux A, Kebebew E, Sebag F, Wolf K, Henry JF, Pacak K, et al. Endocrine tumors associated with the vagus nerve. Endocrine-related cancer. 2016;23(9):R371-9.
  211. de Wailly P, Oragano L, Rade F, Beaulieu A, Arnault V, Levillain P, et al. Malignant pheochromocytoma: new malignancy criteria. Langenbeck's archives of surgery. 2012;397(2):239-46.
  212. Lenders JW, Eisenhofer G, Mannelli M, Pacak K. Phaeochromocytoma. Lancet. 2005;366(9486):665-75.
  213. Motta-Ramirez GA, Remer EM, Herts BR, Gill IS, Hamrahian AH. Comparison of CT findings in symptomatic and incidentally discovered pheochromocytomas. AJR American journal of roentgenology. 2005;185(3):684-8.
  214. Shao Y, Chen R, Shen ZJ, Teng Y, Huang P, Rui WB, et al. Preoperative alpha blockade for normotensive pheochromocytoma: is it necessary? J Hypertens. 2011;29(12):2429-32.
  215. Eisenhofer G, Peitzsch M. Laboratory evaluation of pheochromocytoma and paraganglioma. Clinical chemistry. 2014;60(12):1486-99.
  216. Peaston RT, Graham KS, Chambers E, van der Molen JC, Ball S. Performance of plasma free metanephrines measured by liquid chromatography-tandem mass spectrometry in the diagnosis of pheochromocytoma. Clinica chimica acta; international journal of clinical chemistry. 2010;411(7-8):546-52.
  217. Lenders JW, Pacak K, Walther MM, Linehan WM, Mannelli M, Friberg P, et al. Biochemical diagnosis of pheochromocytoma: which test is best? JAMA. 2002;287(11):1427-34.
  218. Taylor RL, Singh RJ. Validation of liquid chromatography-tandem mass spectrometry method for analysis of urinary conjugated metanephrine and normetanephrine for screening of pheochromocytoma. Clinical chemistry. 2002;48(3):533-9.
  219. Waguespack SG, Rich T, Grubbs E, Ying AK, Perrier ND, Ayala-Ramirez M, et al. A current review of the etiology, diagnosis, and treatment of pediatric pheochromocytoma and paraganglioma. J Clin Endocrinol Metab. 2010;95(5):2023-37.
  220. Singer J, Koch CA, Kassahun W, Lamesch P, Eisenhofer G, Kluge R, et al. A patient with a large recurrent pheochromocytoma demonstrating the pitfalls of diagnosis. Nature reviews Endocrinology. 2011;7(12):749-55.
  221. Osinga TE, van der Horst-Schrivers AN, van Faassen M, Kerstens MN, Dullaart RP, Pacak K, et al. Mass spectrometric quantification of salivary metanephrines-A study in healthy subjects. Clinical biochemistry. 2016;49(13-14):983-8.
  222. Lenders JW, Willemsen JJ, Eisenhofer G, Ross HA, Pacak K, Timmers HJ, et al. Is supine rest necessary before blood sampling for plasma metanephrines? Clinical chemistry. 2007;53(2):352-4.
  223. Darr R, Pamporaki C, Peitzsch M, Miehle K, Prejbisz A, Peczkowska M, et al. Biochemical diagnosis of phaeochromocytoma using plasma-free normetanephrine, metanephrine and methoxytyramine: importance of supine sampling under fasting conditions. Clinical endocrinology. 2014;80(4):478-86.
  224. Boot C, Toole B, Johnson SJ, Ball S, Neely D. Single-centre study of the diagnostic performance of plasma metanephrines with seated sampling for the diagnosis of phaeochromocytoma/paraganglioma. Annals of clinical biochemistry. 2016.
  225. Sawka AM, Prebtani AP, Thabane L, Gafni A, Levine M, Young WF, Jr. A systematic review of the literature examining the diagnostic efficacy of measurement of fractionated plasma free metanephrines in the biochemical diagnosis of pheochromocytoma. BMC endocrine disorders. 2004;4(1):2.
  226. Eisenhofer G, Lenders JW, Timmers H, Mannelli M, Grebe SK, Hofbauer LC, et al. Measurements of plasma methoxytyramine, normetanephrine, and metanephrine as discriminators of different hereditary forms of pheochromocytoma. Clinical chemistry. 2011;57(3):411-20.
  227. Eisenhofer G, Goldstein DS, Walther MM, Friberg P, Lenders JW, Keiser HR, et al. Biochemical diagnosis of pheochromocytoma: how to distinguish true- from false-positive test results. J Clin Endocrinol Metab. 2003;88(6):2656-66.
  228. Bravo EL, Tarazi RC, Fouad FM, Vidt DG, Gifford RW, Jr. Clonidine-suppression test: a useful aid in the diagnosis of pheochromocytoma. N Engl J Med. 1981;305(11):623-6.
  229. Sjoberg RJ, Simcic KJ, Kidd GS. The clonidine suppression test for pheochromocytoma. A review of its utility and pitfalls. Arch Intern Med. 1992;152(6):1193-7.
  230. Sebel EF, Hull RD, Kleerekoper M, Stokes GS. Responses to glucagon in hypertensive patients with and without pheochromocytoma. The American journal of the medical sciences. 1974;267(6):337-43.
  231. Algeciras-Schimnich A, Preissner CM, Young WF, Jr., Singh RJ, Grebe SK. Plasma chromogranin A or urine fractionated metanephrines follow-up testing improves the diagnostic accuracy of plasma fractionated metanephrines for pheochromocytoma. J Clin Endocrinol Metab. 2008;93(1):91-5.
  232. Miehle K, Kratzsch J, Lenders JW, Kluge R, Paschke R, Koch CA. Adrenal incidentaloma diagnosed as pheochromocytoma by plasma chromogranin A and plasma metanephrines. J Endocrinol Invest. 2005;28(11):1040-2.
  233. Eisenhofer G, Huysmans F, Pacak K, Walther MM, Sweep FC, Lenders JW. Plasma metanephrines in renal failure. Kidney international. 2005;67(2):668-77.
  234. Sarathi V, Lila AR, Bandgar TR, Menon PS, Shah NS. Pheochromocytoma and pregnancy: a rare but dangerous combination. Endocrine practice : official journal of the American College of Endocrinology and the American Association of Clinical Endocrinologists. 2010;16(2):300-9.
  235. Neary NM, King KS, Pacak K. Drugs and pheochromocytoma--don't be fooled by every elevated metanephrine. N Engl J Med. 2011;364(23):2268-70.
  236. Welch TJ, Sheedy PF, 2nd, van Heerden JA, Sheps SG, Hattery RR, Stephens DH. Pheochromocytoma: value of computed tomography. Radiology. 1983;148(2):501-3.
  237. Lumachi F, Tregnaghi A, Zucchetta P, Cristina Marzola M, Cecchin D, Grassetto G, et al. Sensitivity and positive predictive value of CT, MRI and 123I-MIBG scintigraphy in localizing pheochromocytomas: a prospective study. Nuclear medicine communications. 2006;27(7):583-7.
  238. Blake MA, Krishnamoorthy SK, Boland GW, Sweeney AT, Pitman MB, Harisinghani M, et al. Low-density pheochromocytoma on CT: a mimicker of adrenal adenoma. AJR American journal of roentgenology. 2003;181(6):1663-8.
  239. Jacques AE, Sahdev A, Sandrasagara M, Goldstein R, Berney D, Rockall AG, et al. Adrenal phaeochromocytoma: correlation of MRI appearances with histology and function. European radiology. 2008;18(12):2885-92.
  240. Bhatia KS, Ismail MM, Sahdev A, Rockall AG, Hogarth K, Canizales A, et al. 123I-metaiodobenzylguanidine (MIBG) scintigraphy for the detection of adrenal and extra-adrenal phaeochromocytomas: CT and MRI correlation. Clinical endocrinology. 2008;69(2):181-8.
  241. Scholz T, Eisenhofer G, Pacak K, Dralle H, Lehnert H. Clinical review: Current treatment of malignant pheochromocytoma. J Clin Endocrinol Metab. 2007;92(4):1217-25.
  242. Solanki KK, Bomanji J, Moyes J, Mather SJ, Trainer PJ, Britton KE. A pharmacological guide to medicines which interfere with the biodistribution of radiolabelled meta-iodobenzylguanidine (MIBG). Nuclear medicine communications. 1992;13(7):513-21.
  243. Mozley PD, Kim CK, Mohsin J, Jatlow A, Gosfield E, 3rd, Alavi A. The efficacy of iodine-123-MIBG as a screening test for pheochromocytoma. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 1994;35(7):1138-44.
  244. Janssen I, Chen CC, Millo CM, Ling A, Taieb D, Lin FI, et al. PET/CT comparing (68)Ga-DOTATATE and other radiopharmaceuticals and in comparison with CT/MRI for the localization of sporadic metastatic pheochromocytoma and paraganglioma. European journal of nuclear medicine and molecular imaging. 2016;43(10):1784-91.
  245. Janssen I, Chen CC, Taieb D, Patronas NJ, Millo CM, Adams KT, et al. 68Ga-DOTATATE PET/CT in the Localization of Head and Neck Paragangliomas Compared with Other Functional Imaging Modalities and CT/MRI. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 2016;57(2):186-91.
  246. Archier A, Varoquaux A, Garrigue P, Montava M, Guerin C, Gabriel S, et al. Prospective comparison of (68)Ga-DOTATATE and (18)F-FDOPA PET/CT in patients with various pheochromocytomas and paragangliomas with emphasis on sporadic cases. European journal of nuclear medicine and molecular imaging. 2016;43(7):1248-57.
  247. Castinetti F, Kroiss A, Kumar R, Pacak K, Taieb D. 15 YEARS OF PARAGANGLIOMA: Imaging and imaging-based treatment of pheochromocytoma and paraganglioma. Endocrine-related cancer. 2015;22(4):T135-45.
  248. Melcescu E, Phillips J, Moll G, Subauste JS, Koch CA. 11Beta-hydroxylase deficiency and other syndromes of mineralocorticoid excess as a rare cause of endocrine hypertension. Hormone and metabolic research = Hormon- und Stoffwechselforschung = Hormones et metabolisme. 2012;44(12):867-78.
  249. Paperna T, Gershoni-Baruch R, Badarneh K, Kasinetz L, Hochberg Z. Mutations in CYP11B1 and congenital adrenal hyperplasia in Moroccan Jews. J Clin Endocrinol Metab. 2005;90(9):5463-5.
  250. Reisch N, Hogler W, Parajes S, Rose IT, Dhir V, Gotzinger J, et al. A diagnosis not to be missed: nonclassic steroid 11beta-hydroxylase deficiency presenting with premature adrenarche and hirsutism. J Clin Endocrinol Metab. 2013;98(10):E1620-5.
  251. Biglieri EG, Herron MA, Brust N. 17-hydroxylation deficiency in man. The Journal of clinical investigation. 1966;45(12):1946-54.
  252. Goldsmith O, Solomon DH, Horton R. Hypogonadism and mineralocorticoid excess. The 17-hydroxylase deficiency syndrome. N Engl J Med. 1967;277(13):673-7.
  253. Imai T, Yanase T, Waterman MR, Simpson ER, Pratt JJ. Canadian Mennonites and individuals residing in the Friesland region of The Netherlands share the same molecular basis of 17 alpha-hydroxylase deficiency. Human genetics. 1992;89(1):95-6.
  254. Costa-Santos M, Kater CE, Auchus RJ, Brazilian Congenital Adrenal Hyperplasia Multicenter Study G. Two prevalent CYP17 mutations and genotype-phenotype correlations in 24 Brazilian patients with 17-hydroxylase deficiency. J Clin Endocrinol Metab. 2004;89(1):49-60.
  255. Ulick S, Levine LS, Gunczler P, Zanconato G, Ramirez LC, Rauh W, et al. A syndrome of apparent mineralocorticoid excess associated with defects in the peripheral metabolism of cortisol. J Clin Endocrinol Metab. 1979;49(5):757-64.
  256. Wilson RC, Krozowski ZS, Li K, Obeyesekere VR, Razzaghy-Azar M, Harbison MD, et al. A mutation in the HSD11B2 gene in a family with apparent mineralocorticoid excess. J Clin Endocrinol Metab. 1995;80(7):2263-6.
  257. Mune T, Rogerson FM, Nikkila H, Agarwal AK, White PC. Human hypertension caused by mutations in the kidney isozyme of 11 beta-hydroxysteroid dehydrogenase. Nature genetics. 1995;10(4):394-9.
  258. Pizzolo F, Friso S, Morandini F, Antoniazzi F, Zaltron C, Udali S, et al. Apparent Mineralocorticoid Excess by a Novel Mutation and Epigenetic Modulation by HSD11B2 Promoter Methylation. J Clin Endocrinol Metab. 2015;100(9):E1234-41.
  259. Wilson RC, Nimkarn S, New MI. Apparent mineralocorticoid excess. Trends in endocrinology and metabolism: TEM. 2001;12(3):104-11.
  260. White PC, Agarwal AK, Nunez BS, Giacchetti G, Mantero F, Stewart PM. Genotype-phenotype correlations of mutations and polymorphisms in HSD11B2, the gene encoding the kidney isozyme of 11beta-hydroxysteroid dehydrogenase. Endocrine research. 2000;26(4):771-80.
  261. Knops NB, Monnens LA, Lenders JW, Levtchenko EN. Apparent mineralocorticoid excess: time of manifestation and complications despite treatment. Pediatrics. 2011;127(6):e1610-4.
  262. Lavery GG, Ronconi V, Draper N, Rabbitt EH, Lyons V, Chapman KE, et al. Late-onset apparent mineralocorticoid excess caused by novel compound heterozygous mutations in the HSD11B2 gene. Hypertension. 2003;42(2):123-9.
  263. Morineau G, Sulmont V, Salomon R, Fiquet-Kempf B, Jeunemaitre X, Nicod J, et al. Apparent mineralocorticoid excess: report of six new cases and extensive personal experience. Journal of the American Society of Nephrology : JASN. 2006;17(11):3176-84.
  264. Wang LP, Yang KQ, Jiang XJ, Wu HY, Zhang HM, Zou YB, et al. Prevalence of Liddle Syndrome Among Young Hypertension Patients of Undetermined Cause in a Chinese Population. Journal of clinical hypertension. 2015;17(11):902-7.
  265. Kellenberger S, Gautschi I, Rossier BC, Schild L. Mutations causing Liddle syndrome reduce sodium-dependent downregulation of the epithelial sodium channel in the Xenopus oocyte expression system. The Journal of clinical investigation. 1998;101(12):2741-50.
  266. Yang KQ, Lu CX, Xiao Y, Liu YX, Jiang XJ, Zhang X, et al. A novel frameshift mutation of epithelial sodium channel beta-subunit leads to Liddle syndrome in an isolated case. Clinical endocrinology. 2015;82(4):611-4.
  267. Cui Y, Tong A, Jiang J, Wang F, Li C. Liddle syndrome: clinical and genetic profiles. Journal of clinical hypertension. 2016.
  268. Nesterov V, Krueger B, Bertog M, Dahlmann A, Palmisano R, Korbmacher C. In Liddle Syndrome, Epithelial Sodium Channel Is Hyperactive Mainly in the Early Part of the Aldosterone-Sensitive Distal Nephron. Hypertension. 2016;67(6):1256-62.
  269. Canessa CM, Schild L, Buell G, Thorens B, Gautschi I, Horisberger JD, et al. Amiloride-sensitive epithelial Na+ channel is made of three homologous subunits. Nature. 1994;367(6462):463-7.
  270. Kashif Nadeem M, Ling C. Liddle's-like syndrome in the elderly. Journal of clinical hypertension. 2012;14(10):728.
  271. Tapolyai M, Uysal A, Dossabhoy NR, Zsom L, Szarvas T, Lengvarszky Z, et al. High prevalence of liddle syndrome phenotype among hypertensive US Veterans in Northwest Louisiana. Journal of clinical hypertension. 2010;12(11):856-60.
  272. Glover M, O'Shaughnessy KM. Molecular insights from dysregulation of the thiazide-sensitive WNK/SPAK/NCC pathway in the kidney: Gordon syndrome and thiazide-induced hyponatraemia. Clinical and experimental pharmacology & physiology. 2013;40(12):876-84.
  273. Ohta A, Schumacher FR, Mehellou Y, Johnson C, Knebel A, Macartney TJ, et al. The CUL3-KLHL3 E3 ligase complex mutated in Gordon's hypertension syndrome interacts with and ubiquitylates WNK isoforms: disease-causing mutations in KLHL3 and WNK4 disrupt interaction. The Biochemical journal. 2013;451(1):111-22.
  274. Nobel YR, Lodish MB, Raygada M, Rivero JD, Faucz FR, Abraham SB, et al. Pseudohypoaldosteronism type 1 due to novel variants of SCNN1B gene. Endocrinology, diabetes & metabolism case reports. 2016;2016:150104.
  275. Geller DS, Rodriguez-Soriano J, Vallo Boado A, Schifter S, Bayer M, Chang SS, et al. Mutations in the mineralocorticoid receptor gene cause autosomal dominant pseudohypoaldosteronism type I. Nature genetics. 1998;19(3):279-81.
  276. Geller DS, Zhang J, Zennaro MC, Vallo-Boado A, Rodriguez-Soriano J, Furu L, et al. Autosomal dominant pseudohypoaldosteronism type 1: mechanisms, evidence for neonatal lethality, and phenotypic expression in adults. Journal of the American Society of Nephrology : JASN. 2006;17(5):1429-36.
  277. Geller DS, Farhi A, Pinkerton N, Fradley M, Moritz M, Spitzer A, et al. Activating mineralocorticoid receptor mutation in hypertension exacerbated by pregnancy. Science. 2000;289(5476):119-23.
  278. New MI, Geller DS, Fallo F, Wilson RC. Monogenic low renin hypertension. Trends in endocrinology and metabolism: TEM. 2005;16(3):92-7.
  279. Ullah MI, Washington T, Kazi M, Tamanna S, Koch CA. Testosterone deficiency as a risk factor for cardiovascular disease. Hormone and metabolic research = Hormon- und Stoffwechselforschung = Hormones et metabolisme. 2011;43(3):153-64.
  280. Ullah MI, Uwaifo GI, Nicholas WC, Koch CA. Does vitamin d deficiency cause hypertension? Current evidence from clinical studies and potential mechanisms. International journal of endocrinology. 2010;2010:579640.

Diagnosis and Treatment of Graves’ Disease

 

 ABSTRACT


Diagnosis of the classic form of Graves’ disease is easy and depends on the recognition of the cardinal features of the disease and confirmation by tests such as TSH and FTI. The differential diagnosis includes other types of thyrotoxicosis, such as that occurring in a nodular gland, accompanying certain tumors of the thyroid, or thyrotoxicosis factitia, and nontoxic goiter. Types of hypermetabolism that imitate symptoms of thyrotoxicosis must also enter the differential diagnosis. Examples are certain cases of pheochromocytoma, polycythemia, lymphoma, and the leukemias. Pulmonary disease, infection, parkinsonism, pregnancy, or nephritis may stimulate certain features of thyrotoxicosis.
Treatment of Graves’ disease cannot yet be aimed at the cause because it is still unknown. One seeks to control thyrotoxicosis when that seems to be the major indication, or the ophthalmopathy when that aspect of the disease appears to be more urgent. The available forms of treatment, including surgery, drugs, and 131-I therapy, are reviewed. There is a difference of opinion as to which of these modalities is best, but to a large degree guidelines governing choice of therapy can be drawn. Antithyroid drugs are widely used for treatment on a long- term basis. About one-third of the patients undergoing long-term antithyroid therapy achieve permanent euthyroidism. Drugs are the preferred initial therapy in children and young adults. Subtotal thyroidectomy is a satisfactory form of therapy, if an excellent surgeon is available, but is less used in 2016. The combined use of antithyroid drugs and iodine makes it possible to prepare patients adequately before surgery, and operative mortality is approaching the vanishing point. Many young adults, are treated by surgery if antithyroid drug treatment fails.
Currently, most endocrinologists consider RAI to be the best treatment for adults, and consider the associated hypothyroidism to be a minor problem. Evidence to date after well over five decades of experience indicates that the risk of late thyroid  carcinoma must be near zero. The authors advise this therapy in most patients over age 40, and believe that it is not contraindicated above the age of about 15. Dosage is calculated on the basis of 131-I uptake and gland size. Most patients are cured by one treatment. Hypothyroidism.occurs with a fairly constant frequency for many years after therapy and may be  unavoidable  if cure of the disease is to be achieved by 131-I.. Many therapists accept this as an anticipated outcome of treatment.
Thyrotoxicosis in children is best handled initially by antithyroid drug therapy. If this therapy does not result in a cure, surgery may be performed. Treatment with 131-I is accepted as an alternative form of treatment by some physicians, especially as age increase toward  15 years. Neonatal thyrotoxicosis is a rarity. Antithyroid drugs, propranolol and iodide may be required for several weeks until maternally-derived antibodies have been metabolized.
The physician applying any of these forms of therapy to the control of thyrotoxicosis should also pay heed to the patient’s emotional needs, as well as to his or her requirements for rest, nutrition, and specific antithyroid medication. Consult our FREE web-book WWW.ENDOTEXT.ORG  for complete coverage on this and related topics.

We note that there  are currently available 2 very extensive Guidelines on Diagnosis and Treatment of Graves’ Disease—The 2016 ATA guideline  --- http://online.liebertpub.com/doi/pdfplus/10.1089/thy.2016.0229 (270 pages), and the AACE 2011 version on Hyperthyroidism and other Causes of Thyrotoxicosis (65 pages)--https://www.aace.com/files/hyperguidelinesapril2013.pdf.
Both are well worth reviewing.

CLINICAL DIAGNOSIS

The diagnosis of Graves’ disease is usually easily made. The combination of eye signs, goiter, and any of the characteristic symptoms and signs of hyperthyroidism forms a picture that can hardly escape recognition (Fig -1). It is only in the atypical cases, or with coexisting disease, or in mild or early disease, that the diagnosis may be in doubt. The symptoms and signs have been described in detail in the section on manifestations of Graves’ disease. For convenience, the classic findings from the history and physical examination are grouped together in Table 1a and 1b.These occur with sufficient regularity that clinical diagnosis can be reasonably accurate. Scoring the presence or absence and severity of particular symptoms and signs can provide a clinical diagnostic index almost as reliable a diagnostic measure as laboratory tests(1).

Occasionally diagnosis is not at all obvious.In patients severely ill with other disease, in elderly patients with "apathetic hyperthyroidism", or when the presenting symptom is unusual, such as muscle weakness, or psychosis, the diagnosis depends on clinical alertness and laboratory tests.

The diagnosis of Graves’ Disease does not only depend on thyrotoxicosis. Ophthalmopathy, or pretibial myxedema may occasionally occur without goiter and thyrotoxicosis, or even with spontaneous hypothyroidism. While proper classification can be debated, these patients seem to represent one end of the spectrum of Graves’ Disease. Thus we are usually making two coincident diagnoses:1)- Is the patient hyperthyroid? and 2)- Is the cause of the problem Graves’ disease ?.

Table 1a---Symptoms of Graves’ disease

 

  • Preference for cool temperature
  • Weight loss with increased appetite
  • Prominence of eyes, puffiness of lids
  • Pain or irritation of eyes
  • Blurred or double vision, decreasing acuity, decreased motility
  • Goiter
  • Dyspnea
  • Palpitations or pounding of the heart
  • Ankle edema (without cardiac disease)
  • Less frequently, orthopnea, paroxysmal tachycardia, anginal pain, and CHF
  • Increased frequency of stools
  • Polyuria
  • Decrease in menstrual flow; menstrual irregularity or amenorrhea
  • Decreased fertility
  • Fatigue
  • Weakness, Tremor
  • Occasional bursitis
  • Rarely periodic paralysis
  • Nervousness, irritability
  • Emotional lability
  • Insomnia or decreased sleep requirement
  • Thinning of hair, Loss of curl in hair
  • Increased perspiration
  • Change in texture of skin and nails
  • Vitiligo
  • Swelling over out surface of shin

Family history of any thyroid  disease, especially Graves’ disease

 

TABLE 1B       PHYSICAL SIGNS

 

  • Weight loss
  • Hyperkinetic behavior, thought, and speech
  • Restlessness
  • Lymphadenopathy and occasional splenomegaly
  • Eyes
  • Prominence of eyes, lid lag, globe lag
  • Exophthalmos, lid edema, chemosis, extraocular muscle weakness
  • Decreased visual acuity, scotomata, papilledema, retinal hemorrhage, and edema
  • Goiter
  • Sometimes enlarged cervical nodes
  • Thyroid thrill and bruit
  • Tachypnea on exertion
  • Tachycardia, overactive heart, widened pulse pressure, and bounding pulse
  • Occasional cardiomegaly, signs of congestive heart failure, and paroxysmal tachycardia or atrial fibrillation
  • Tremor
  • Objective muscle wasting and weakness
  • Quickened and hypermetric reflexes
  • Emotional lability
  • Fine, warm, moist skin
  • Fine and often straight hair
  • Oncholysis (Plummer’s nails)
  • Pretibial myxedema, Acropachy
    Hyperpigmentation or vitiligo

LABORATORY DIAGNOSIS OF GRAVE’S DISEASE

Serum Hormone Measurements

TSH and FT4 assay-Once the question of thyrotoxicosis has been raised, laboratory data are required to verify the diagnosis, help estimate the severity of the condition, and assist in planning therapy. A single test such as the TSH or estimate of FT4 (free T4) may be enough, but in view of the sources of error in all determinations, most clinicians prefer to assess two more or less independent measures of thyroid function. For this purpose, an assessment of FT4 and sensitive TSH are suitable.
As an initial single test, a sensitive TSH assay may be most cost-effective and specific. TSH should be 0 - .1 µU/ml in significant thyrotoxicosis, although values of .1 - .3 are seen in patients with mild illness, especially with smoldering toxic multinodular goiter in older patients(1.1). TSH can be low in some elderly patients without evidence of thyroid disease. TSH can be normal -- or elevated -- only if there are spurious test results from heterophile antibodies or other cause, or the thyrotoxicosis is TSH-driven, as in a pituitary TSH-secreting adenoma or pituitary resistance to thyroid hormone.
Measurement of FT4 or FTI (Free thyroxine index)is also usually diagnostic.The degree of elevation of the FT4 above normal provides an estimate of the severity of the disease. During replacement therapy with thyroxine the range of both FTI and fT4 values tend to be about 20% above the normal range, possibly because only T4, rather than T4 and T3 from the thyroid, is providing the initial supply of hormone. Thus many patients will have an fT4 or FTI above normal when appropriately replaced and while TSH is in the normal range. Except for this, elevations of fT4 not due to thyrotoxicosis are unusual, and causes are given in Table 3.

Of course the Total T4 level may normally be as high as 16 or 20 µg/dl in pregnancy, and can be elevated without thyrotoxicosis in patients with familial hyperthyroxinemia due to abnormal albumin, the presence of hereditary excess TBG, the presence of antibodies binding T4 , the thyroid hormone resistance syndrome, and other conditions listed in Table 3. The T4 level may be normal in thyrotoxic patients who have depressed serum levels of T4 -binding protein or because of severe illness, even though they are toxic. Thus, thyrotoxicosis may exist when the total T4 level is in the normal range. However measurement of FT4, FT3 (Free T3), or FTI (Free Thyroxine Index) usually obviates this source of error and is the best test. In the presence of typical symptoms, one measurement of suppressed TSH or elevated fT4 is sufficient to make a definite diagnosis, although it does not identify a cause. If the fT4 is normal, repetition is in order to rule out error, along with a second test such as serum FT3. And it should be noted that in much of Europe FT3 is the  preferred test, rather than FT4, and serves very well.

A variety of methods for FT4 determination have become available, including commercial kits. Although these methods are usually reliable, assays using different kits do not always agree among themselves or with the determination of FT4 by dialysis. Usually T4 and T3 levels are both elevated in thyrotoxicosis, as is the FTI (Free Thyroxin Index), or an index constructed using the serum T3 and rT3U levels, and the newer measures of FT3.

Table 3. Conditions Associated with Transient Elevations of the FT4 or FTI

Condition Explanation
Estrogen withdrawal Rapid decrease in TBG level
Amphetamine abuse Possibly induced TSH secretion(2)
Acute psychosis Unknown
Hyperemesis gravidarum Associated high hCG can cause thyrotoxicosis
Iodide administration Thyroid autonomy
Beginning of T4 administration Delayed T4 metabolism(3)
Severe illness (rarely) Decreased T4 to T3 conversion (4)
Amiodarone treatment Decreased T4 to T3 conversion, iodine load
Gallbladder contrast agents Decreased T4 to T3 conversion, iodine load
Propranolol (large doses) Inhibition of T4 to T3 conversion
Prednisone (rarely) Inhibition of T4 to T3 conversion
High altitude exposure Possibly hypothalamic activation
Selenium deficiency Decreased T4 to T3 conversion

T3 and FT3 ASSAY-The serumT3 level determined by RIA is almost always elevated in thyrotoxicosis and is a useful but not commonly needed secondary test. Usually the serum T3 test is interpreted directly without use of a correction for protein binding, since alterations of TBG affect T3 to a lesser extent than T4. Any confusion caused by alterations in binding proteins can be avoided by use of a FT3 assay or T3 index calculated as for the FTI. Generally the FT3 assay is as diagnostically effective as the FT4. In patients with severe illness and thyrotoxicosis, especially those with liver disease or malnutrition or who are taking steroids or propranolol, the serum T3 level may not be elevated, since peripheral deiodination of T4 to T3 is suppressed ("T4 toxicosis"). A normal T3 level has also been observed in thyrotoxicosis combined with diabetic ketoacidosis. Whether or not these patients actually have tissue hypermetabolism at the time their serum T3 is normal is not entirely certain. In these patients the rT3 level may be elevated. If the complicating illness subsides, the normal pattern of elevated T4 , FTI, and T3 levels may return(5,6). Elevated T4 levels with normal serum T3 levels are also found in patients with thyrotoxicosis produced by iodine ingestion(7).

T3 Toxicosis Since 1957, when the first patient with T3 thyrotoxicosis was identified, a number of patients have been detected who had clinical thyrotoxicosis, normal serum levels of T4 and TBG, and elevated concentrations of T3 and FT3[8]. Hollander et al [9] found that approximately 4% of patients with thyrotoxicosis in the New York area fit this category. These patients often have mild disease but otherwise have been indistinguishable clinically from others with thyrotoxicosis. Some have had the diffuse thyroid hyperplasia of Graves’ disease, others toxic nodular goiter, and still others thyrotoxicosis with hyperfunctioning adenomas. Interestingly, in Chile, a country with generalized iodine deficiency, 12.5% of thyrotoxic subjects fulfilled the criteria for T3 thyrotoxicosis [10]. Asymptomatic hypertriiodothyronemia is an occasional finding several months before the development of thyrotoxicosis with elevated T4 levels [11]. Since T4 is normally metabolized to T3, and the latter hormone is predominantly the hormone bound to nuclear receptors, it makes sense that elevation of T3 alone is already indicative of thyrotoxicosis.

Thyroid Isotope uptake-In patients with thyrotoxicosis the RAIU (Radioactive Iodine Uptake) at 24 hours is characteristically above normal. In the United States, which has had an increasing iodine supply in recent years, the upper limit of normal is now about 25% of the administered dose. This value is higher in areas of iodine deficiency and endemic goiter. The uptake value at a shorter time interval, for example 6 hours, is as valid a test and may be more useful in the infrequent cases having such a rapid isotope turnover that "uptake" has fallen to normal by 24 hours. If there is reason to suspect that thyroid isotope turnover is rapid, it is wise to do both a 6- and a 24-hour RAIU determination during the initial laboratory study. As noted below, rapid turnover of 131-I can seriously reduce the effectiveness of 131-I therapy. Similar studies can be done with 123-I and also technetium. Because of convenience, and since serum assays of thyroid hormones and TSH are reliable and readily available, the RAIU is now infrequently determined unless 131-I therapy is planned.. It is however useful in patients who are mildly thyrotoxic for factitia thyrotoxicosis, subacute thyroiditis and painless thyroiditis in whom RAIU is low, thus confirming thyrotoxicosis in the absence of  elevated RAIU. This may include patients with brief symptom duration, small goiter, or lacking eye signs, absent family history, or negative antibody test result. Obviously other causes of a low RAIU test need to be considered and excluded. Tests measuring suppressibility of RAIU are of historical interest(13-15)

Thyroid IsotopeScanning-Isotope scanning of the thyroid has a limited role in the diagnosis of thyrotoxicosis. It is useful in patients in whom the thyroid is difficult to feel or in whom nodules (single or multiple) are present that require evaluation, or rarely to prove the function of ectopic thyroid tissue. Nodules may be incidental, or may be the source of thyrotoxicosis (toxic adenoma), or may contribute to the thyrotoxicosis that also arises from the rest of the gland. Scanning should usually be done with 123-I in this situation, in order to combine it with an RAIU measurement.

Thyroid  Ultrasound- US exam of the thyroid is sometimes of value in diagnosis. For example, if a possible nodule is detected on physical exam. It also may confirm hypoechogenicity or intense vascularity of Graves’ disease if a color Doppler flow exam is done.

Antithyroid Antibodies Determination of antibody titers provides supporting evidence for Graves’ disease. More than 95% of patients have positive assays for TPO (thyroperoxidase or microsomal antigen), and about 50% have positive anti-thyroglobulin antibody assays. In thyroiditis the prevalence of positive TG antibody assays is higher. Positive assays prove that autoimmunity is present, and  patients with causes of thyrotoxicosis other than Graves’ disease usually have negative assays. During therapy with antithyroid drugs the titers characteristically go down, and this change persists during remission. Titers tend to become more elevated after RAI treatment.

Antibodies to TSH-Receptor-Thyrotrophin receptor antibody (TRAb) assays have become readily available, and a positive result strongly supports the diagnosis of Graves’ disease(15.1). Determination of TRAb is not required for the diagnosis, but the implied specificity of a positive test provides security in diagnosis, and for this reason the assay is now widely used. The assay is valuable as another supporting fact in establishing the cause of exophthalmos, in the absence of thyrotoxicosis. High maternal levels suggest possible fetal or neonatal thyrotoxicosis. TRAb assays measure any antibody that binds to the TSH-R. Assays for Thyroid Stimulating Antibodies (TSAb,TSI) are less available, but are more specific for the diagnosis. Using current tests, both are positive in about 90% of patients with Graves disease who are thyrotoxic. "Second generation" assays becoming available use monoclonal anti-TSH-R antibodies and biosynthetic TSH-R in coated tube assays, are reported to reach 99% specificity and sensitivity(15.2,15.3,3). Although rarely required, serial assays are of interest in following a patient’s course during antithyroid drug therapy, and a decrease predicts probable remission(15.4).

Other Assays Rarely Used-General availability of assays that can reliably measure suppressed TSH has made this the gold standard to which other tests must be compared, and has effectively eliminated the need for most previously used ancillary tests. There are only rare causes of confusion in the TSH assay. Severe illness, dopamine and steroids, and hypopituitarism, can cause low TSH, but suppression below 0.1 µ/ml is uncommon and below 0.05 µ/ml is exceptional, except in thyrotoxicosis. Thyrotoxicosis is associated with normal or high TSH in patients with TSH producing pituitary tumors and selective pituitary resistance to thyroid hormone.
If TSH, FT4, TRAb, and other tests noted above do not establish the diagnosis, it may be wise to do nothing further except to observe the course of events. In patients with significant thyroid hyperfunction, the symptoms and signs will become clearer, and the laboratory measurements will fall into line. Measurement of BMR, T3 suppression of RAIU, TRH testing, and clinical response to KI are of historical interest.

DIFFERENTIAL DIAGNOSIS of THYROTOXICOSIS

Graves’ disease must be differentiated from other conditions causing thyrotoxicosis. (Table -4).

Thyrotoxicosis factitia-Thyrotoxicosis may be caused by taking T4 or its analogs, most commonly due to administration of excessive replacement hormone by the patient’s physician. Hormone may be taken surreptitiously by the patient for weight loss or psychologic reasons. The typical findings are a normal or small thyroid gland, a low131-I uptake, a low serum TG, and, of course, a striking lack of response to antithyroid drug therapy. The problem can easily be confused with "painless thyroiditis", but in thyrotoxicosis factitia, the gland is typically small.

Toxic nodular goiter is usually distinguished by careful physical examination and a history of goiter for many years before symptoms of hyperthyroidism developed. The thyrotoxicosis comes on insidiously, and often, in the older people usually afflicted, symptoms may be mild, or suggest another problem such as heart disease. The thyroid scan may be diagnostic, showing areas of increased and decreased isotope uptake. The results of assays for antithyroid antibodies, including TRAb, are usually negative. TMNG is typically produced by activating somatic mutations in TSH-R in one or more nodules, allowing them to enlarge and become functional even in the absence of TSH stimulation. (Interestingly, cats are well known to develop hyperthyroidism, with thyroid autonomy, often due to TSH-R gene mutations as seen in humans.(16))

Hyperfunctioning solitary adenoma is suggested on the physical finding of a palpable nodule in a otherwise normal gland, and is proved by a scintiscan demonstrating preferential radioisotope accumulation in the nodule. This type of adenoma must be differentiated from congenital absence of one of the lobes of the thyroid. Toxic nodules typically present in adults with gradually developing hyperthyroidism and a nodule > 3 cm in size. These nodules are usually caused by activating somatic mutations in the TSH-R, which endows them with mildly increased function, compared to normal tissue, even in the absence of TSH. These nodules are usually, but not always, monoclonal(17). In adults toxic nodules are very rarely malignant. Rarely, functioning thyroid carcinomas produce thyrotoxicosis. The diagnosis is made by the history, absence of the normal thyroid, and usually widespread functioning metastasis in lung or bones. Invasion of the gland by lymphoma has produced thyrotoxicosis, presumably due to thyroid destruction (18).

Thyrotoxicosis associated with subacute thyroiditis is usually mild and transient, and the patient lacks the physical findings of long-standing thyrotoxicosis. If thyrotoxicosis is found in conjunction with a painful goiter and low or absent 131-I uptake, this diagnosis is likely. Usually the erythrocyte sedimentation rate (ESR) and CRP are greatly elevated, and the leukocyte count may also be increased. Occasionally the goiter is non-tender. Antibody titers are low or negative. Many patients have the HLA-B35 antigen, indicating a genetic predisposition to the disease. The rare TSH secreting pituitary adenoma will be missed unless one measures the plasma TSH level, or until the enlargement is sufficient to produce deficiencies in other hormones, pressure symptoms, or expansion of the sella turcica(19). These patients have thyrotoxicosis with inappropriately elevated TSH levels and may/or may not secrete more TSH after TRH stimulation. The characteristic finding is a normal or elevated TSH, and an elevated TSH alpha subunit level in blood, measured by special RIA. TRAbs are not present. Exophthalmos, and antibodies of Graves’ disease are absent. Family history is sometimes positive for a similar condition. Demonstration of a suppressed TSH level excludes these rare cases.

The category of patients with thyrotoxicosis and inappropriately elevated TSH levels also includes the rare persons with pituitary "T3 resistance" as a part of the Resistance to Thyroid Hormone syndrome caused by TH Receptor mutations. The syndrome of Pituitary Thyroid Hormone Resistance is usually marked by mild thyrotoxicosis, mildly elevated TSH levels, absence of pituitary tumor, a generous response to TRH, no excess TSH alpha subunit secretion [19,20, 21],and by TSH suppression if large doses of T3 are administered. Final diagnosis depends on laboratory demonstration of a mutation in the TR gene, if possible. Hyperthyroidism caused by excess TRH secretion is a theoretical but unproven possibiity.

Administration of large amounts of iodide in medicines, for roentgenographic examinations, or in foods can occasionally precipitate thyrotoxicosis in patients with multinodular goiter or functioning adenomas. This history is important to consider since the illness may be self-limiting. Induction of thyrotoxicosis has also been observed in apparently normal individuals following prolonged exposure to organic iodide containing compounds such as antiseptic soaps and amiodarone. Amiodarone is of special importance since the clinical problem often is the presentation of thyrotoxicosis in a patient with serious cardiac disease including dysrythmia. Amodarone can induce thyrotoxicosis in patients without known prior thyroid disease, or with multinodular goiter. The illness appears to come in two forms. In one the RAIU may be low or normal. In the second variety , which appears to be more of a thyroiditis-like syndrome, the RAIU is very suppressed, and IL-6 may be elevated. In either case TSH is suppressed, FTI may be normal or elevated, but T3 is elevated if the patient is toxic. Antibodies are usually negative.

An increasingly recognized form of thyrotoxicosis is the syndrome described variously as painless thyroiditis, transient thyrotoxicosis, or "hyperthyroiditis". Its hallmarks are self-limited thyrotoxicosis, small painless goiter, and low or zero RAIU(22,23). The patients usually have no eye signs, a negative family history, and often positive antibody titers. This condition is due to autoimmune thyroid disease, and is considered a variant of Hashimoto’s Thyroiditis. It occurs sporadically, usually in young adults. It frequently occurs 3 - 12 weeks after delivery, sometimes representing the effects of immunologic rebound from the immunosuppressive effects of pregnancy in patients with Hashimoto’s thyroiditis or prior Graves’ Disease, and is called Post Partum Thyroiditis(22-25). The course typically includes development of a painless goiter, mild to moderate thyrotoxicosis, no eye signs, remission of symptoms in 3 -20 weeks, and often a period of hypothyroidism before return to euthyroid function. The cycle may be repeated several times. Histologic examination shows chronic thyroiditis, but it is not typical of Hashimoto’s disease or subacute thyroiditis and may revert to normal after the attack(26). In most patients, the thyrotoxic episode occurs in the absence of circulating TSAb. This finding suggests that the pathogenesis is quite distinct from that in Graves’ disease. The thyrotoxicosis is caused by an inflammation-induced discharge of preformed hormone due to the thyroiditis. The T4/T3 ratio is higher than in typical Graves’ disease,and thyroid iodine stores are depleted. Since the thyrotoxicosis is due to an inflammatory process, therapy with antithyroid drugs or potassium iodide is usually to no avail, and RAI treatment of course cannot be given when RAIU  is suppressed. Propranolol is usually helpful for symptoms. Glucocorticoids may be of help if the process -- often transient and mild -- requires some form of therapy. Propylthiouracil and/or ipodate can be used to decrease T4 to T3 conversion and will ameliorate the illness. Repeated episodes may be handled by surgery or by RAI therapy during a remission. Occasionally painless post-partum thyroiditis is followed by typical Graves’ Disease(27-29.1).

Hyperemisis gravidarum is frequently associated with elevated serum T4 , FTI, and variably elevated T3, and suppressed TSH. The abnormalities in thyroid function are caused by high levels of hCG. This molecule, or a closely related form, share enough homology with TSH so that it has about 1/1000 the thyroid stimulating activity of TSH, and can produce thyroid stimulation or thyrotoxicosis(29.12-29.14). It is typically self limited without specific treatment, disappears with termination of pregnancy, but may occasionally require anti-thyroid treatment temporarily or throughout pregnancy(29.3). Patients with minimal signs and symptoms, small or no goiter, and elevation of FTI up to 50 % above normal probably do not require treatment. Rarely those with goiter, moderate or severe clinical evidence of thyrotoxicosis, highly elevated T4 and T3 and suppressed TSH are best treated with antithyroid drugs. If antibodies are positive or eye signs are present, the picture is usually interpreted as a form of Graves’ disease. Familial severe hyperemesis gravidarum with fetal loss has been reported with an activating germline mutation in the TSH-R, which made it specifically more sensitive to activation by hCG(.29.2,29.3).  Hyperthyroidism can be induced by “hyperplacentosis”, which is characterized by increased placental weight and circulating hCG levels higher than those in normal pregnancy(29.4). After hysterotomy, hCG levels declined in the one case reported and hyperthyroidism was corrected.

Congenital hyperthyroidism caused by a germ-line activating mutation in the TSH-R has recently been recognized . The mutations are usually single aminoacid transitions in the extracellular loops or transmembrane segments of the receptor trans-membrane domain. The diagnosis may be difficult to recognize in the absence of a family history. However the patients lack eye signs, and have negative assays for antibodies(29.2, 29.3)

Hydatidiform moles, choriocarcinomas, and rarely seminomas secrete vast amounts of hCG. hCG, with an alpha subunit identical to TSH , and beta subunit related to TSH , that binds to and activates the thyroid TSH receptor with about 1/1,000th the efficiency of TSH itself (Fig.-3)(30-33). Current evidence indicates that very elevated levels of native hCG or perhaps desialated hCG, cause the thyroid stimulation. Many patients have goiter or elevated thyroid hormone levels or both, but little evidence of thyrotoxicosis, whereas others are clearly thyrotoxic. Diagnosis rests on recognizing the tumor (typically during or after pregnancy) and measurement of hCG. Therapy is directed at the tumor.

Hyperthyroidism also is seen as one manifestation of autoimmune thyroid disease induced by interferon-alpha treatment of chronic hepatitis C. It can be self limiting, or severe enough to require cessation of IFN, or in some cases continue on after INF is stopped(33.1).

Hyperthyroidism also occurs during immune reconstitution seen after effective anti-viral therapy of patients with HIV(33.2), has occurred during recovery of low lymphocyte levels induced by therapy with CAMPATH in patients with Multiple sclerosis, has occurred after cessation of immune-suppressive treatment in patients with T1DM.

Table 4. Causes of Thyrotoxicosis

Disease Course of disease Physical finding Diagnostic finding Treatment/Comment
Graves’ disease Familial, prolonged Goiter + Ab, + RAIU, eye signs Antithyroids, RAI, Surgery
Transient thyrotoxicosis Brief Small goiter Low Ab, no eye signs, RAIU=0 Time, beta blocker, steroids
Subacute thyroiditis Brief Tender goiter RAIU=0, elevated ESR, recent URI Nothing, NSAID, steroids
Toxic multinodular goiter Prolonged, mild Nodular goiter Typical scan Antithyroids, RAI, surgery
Iodide induced Recent, mild Nodular goiter, occ.normal Low RAIU, abnormal scan Antithyroids, KClO4, time, stop I source
Toxic adenoma Prolonged, mild One nodule "Hot" nodule on scan Surgery, RAI
Thyroid carcinoma Recent Variable, metastases Functioning metastases Surgery + RAI
Exogenous hormone Variable Small thyroid RAIU and TG low, psychiatric illness Withdrawal, counseling
Hydatiform mole Recent, mild Goiter Pregnancy, bleeding,HCG Surgery, chemotherapy
Choriocarcinoma Recent, mild Goiter Increased HCG Surgery, chemotherapy
TSH-oma Prolonged Goiter Excess alpha, TSH, adenoma Op, somatostatin, thyroid ablation
Pituitary T3 resistance Prolonged Goiter Elevated or normal TSH, no tumor, mod. thyrotox, no excess alpha Triac, somatostatin, thyroid ablation, beta blocker
Struma ovarii Variable + / - goiter Positive scan or US Surgery
Thyroid destruction Variable Variable Variable Steroids
Hamburger toxicosis Recent, self-limited Small gland, no eye signs Suppressed TSH and TG and RAIU Avoid neck meat trimmings
Hyperemesis Onset first trimester Pregnancy, variably toxic UP FTI, Low TSH, High HCG ATD if severe, pregnancy termination
TSH-R mutation Congenital Typical thyrotoxicosis + FH, germline mutation Thyroid ablation
Familial gestational hyperthyroidism Onset first trimester Severe hyperthyroidism + FH, TSH-R mutation sensitizing to hCG ATD, Surgery
Amiodarone Prolonged Thyroid usually enlarged. Often heart disease. Suppressed RAIU, nl or increased FTI, elevated T3 ATD + KClO4,Prednisone, Surgery,iopanoic acid
Interferon-alpha induced Induced by INF treatment of hepatitis C Clinically significant Often remits if IFN stopped.
Treatment of HIV During T cell recovery Clinically significant With or without prior thyroid autoimmunity May need treatment
Administration of CAMPATH During recovery of T cells Clinically significant With or without prior thyroid autoimmunity May need treatment
Sunitinib therapy During tyrosine kinase therapy for cancer Clinically significant Usually induces hypothyroidism, rarely hyper May need treatment

 

Subclinical hyperthyroidism


 It should be remembered that thyrotoxicosis is today not only a clinical but also a laboratory diagnosis. Consistent elevation of the fT4 , and the T3 level, and suppressed TSH, or only suppression of TSH, can indicate that thyrotoxicosis is present even in the absence of clear-cut signs or symptoms These elevations themselves are a sufficient indication for therapy, especially in elderly patients with coincident cardiac disease(33a,b). Antithyroid drug treatment of patients with subclinical hyperthyroidism was found to result in a decrease in heart rate, decrease in number of atrial and ventricular premature beats, a reduction of the left ventricular mass index, and left ventricular posterior wall thickness, as well as a reduction in diastolic peak flow velocity. These changes are considered an argument for early treatment of subclinical hyperthyroidism. Subclinical hyperthyroidism may disappear or evolve into Graves hyperthyroidism, or when caused by MNG, persist for long periods unchanged.
Individuals of any age with consistent suppression of TSH should be fully evaluated to determine if evidence of hyperthyroidism is present, or there is coincident disease that might be aggrevated by hyperthyroidism. SCH with TSH of 0.2-0.3.5 may not need treatment. Individuals with TSH at or below 0.1uU/ml most likely will require treatment by one of the methods described below.

Apathetic hyperthyroidism designates a thyrotoxic condition characterized by fatigue, apathy, listlessness, dull eyes, extreme weakness, often congestive heart failure, and low-grade fever.[ 34, 35] Often such patients have small goiters, modest tachycardia, occasionally cool and even dry skin, and few eye signs. The syndrome may, in some patients, represent an extreme degree of fatigue induced by long-standing thyrotoxicosis. Once the diagnosis is considered, standard laboratory tests should confirm or deny the presence of thyrotoxicosis even in the absence of classical symptoms and signs.

Other diagnostic problems  Two common diagnostic problems involve (1) the question of hyperthyroidism in patients with goiter of another cause, and (2) mild neuroses such as anxiety, fatigue states, and neurasthenia. Most patients with goiter receive a battery of examinations to survey their thyroid function at some time. Usually these tests are done more for routine assessment than because there is serious concern over the possibility of thyrotoxicosis. In the absence of significant symptoms or signs of hyperthyroidism and ophthalmologic problems, a normal FTI or TSH determination is sufficiently reassuring to the physician and the patient. Of course, the most satisfactory conclusion of such a study is the identification of an alternate cause for enlargement of the thyroid.
Some patients complain of fatigue and palpitations, weight loss, nervousness, irritability, and insomnia. These patients may demonstrate brisk reflex activity, tachycardia (especially during examinations), perspiration, and tremulousness. In the abscence of thyrotoxicosis, the hands are more often cool and damp rather than warm and erythematous. Serum TSH assay should be diagnostic.

Mild and temporary elevation of the FTI may occur if there is a transient depression of TBG production -- for example, when estrogen administration is omitted. This problem is occasionally seen in hospital practice, usually involving a middle-aged woman receiving estrogen medication that is discontinued when the patient is hospitalized. Estrogen withdrawal leads to decreased TBG levels and a transiently elevated FTI. After two to three weeks, both the T4 level and the FTI return to normal ( Table -3).
In the differential diagnosis of heart disease, the possibility of thyrotoxicosis must always be considered. Some cases of thyrotoxicosis are missed because the symptoms are so conspicuously cardiac that the thyroid background is not perceived. This is especially true in patients with atrial fibrillation.
Many disorders may on occasion show some of the features of hyperthyroidism or Graves’ disease. In malignant disease, especially lymphoma, weight loss, low grade fever, and weakness are often present. Parkinsonism in its milder forms may initially suggest thyroid disease. So also do the flushed countenance, bounding pulse, thyroid hypertrophy, and dyspnea of pregnancy. Patients with chronic pulmonary disease may have prominent eyes, tremor, tachycardia, weakness, and even goiter from therapeutic use of iodine. One should remember the weakness, fatigue, and jaundice of hepatitis and the puffy eyes of trichinosis and nephritis. Cirrhotic patients frequently have prominent eyes and lid lag, and the alcoholic patient with tremor, prominent eyes, and flushed face may be initially suspected of having thyrotoxicosis. Distinguishing between Graves’ disease with extreme myopathy and myopathies of other origin can be clinically difficult. The term chronic thyrotoxic myopathy is used to designate a condition characterized by weakness, fatigability, muscular atrophy, and weight loss usually associated with severe thyrotoxicosis. Occasionally fasciculations are seen. The electromyogram result may be abnormal. If the condition is truly of hyperthyroid origin, the thyroid function tests are abnormal and the muscular disorder is reversed when the thyrotoxicosis is relieved. Usually a consideration of the total clinical picture and assessment of TSH and FTI are sufficient to distinguish thyrotoxicosis from polymyositis, myasthenia gravis, or progressive muscular atrophy. True myasthenia gravis may coexist with Graves’ disease, in which case the myasthenia responds to neostigmine therapy. (The muscle weakness of hyperthyroidism may be slightly improved by neostigmine, but never relieved.) Occasionally electromyograms, muscle biopsy, neostigmine tests, and ACH-receptor antibody assays must be used to settle the problem.

TREATMENT OF THYROTOXICOSIS–
SELECTION OF PRIMARY THERAPY

No treatment is ideal and thus indicated in all patients ( 35.1).Three forms of primary therapy for Graves’ disease are in common use today: (1) destruction of the thyroid by 131-I; (2) blocking of hormone synthesis by antithyroid drugs; and (3) partial or total surgical ablation of the thyroid. Iodine alone as a form of treatment was widely used in the past, but is not used today because its benefits may be transient or incomplete and because more dependable methods became available. Iodine is primarily used now in conjunction with antithyroid drugs to prepare patients for surgical thyroidectomy when that plan of therapy has been chosen. There is, however, some revival of interest in use of iodine treatment as described subsequently. Roentgen irradiation was also used in the past, but is not currently [36]. Suppression of the autoimmune response is being attempted, and currently new treatments blocking the action of Thyroid Stimulating Immunoglobulins are being investigated.

Selection of therapy depends on a multiplicity of considerations [36.1]. Availability of a competent surgeon, for example, undue emotional concern about the hazards of 131-I irradiation, or the probability of adherence to a strict medical regimen might govern one’s decision regarding one program of treatment as opposed to another. More than 90% of patients are satistactorly treated cumulating the effects of these treatment.(36.2) Fig. 2

Antithyroid drug therapy offers the opportunity to avoid induced damage to the thyroid (and parathyroids or recurrent nerves), as well as exposure to radiation and operation. In recent studies patients with thyroids under 40 gm weight, with low TRAb levels, and age over 40, were most likely to enter remission (in up to 80%) (36.3, 36.31). The difficulties are the requirement of adhering to a medical schedule for many months or years, frequent visits to the physician, occasional adverse reactions, and, most importantly, a disappointingly low permanent remission rate. Therapy with antithyroid drugs is used as the initial modality in most patients under age 18, in many adults through age 40, and in most pregnant women(36.31). Remission is most likely in young patients, with small thyroids, and mild disease. ATDs may be preferred in  elderly patients, those with serious co-morbidities and who have been previously operated upon.

Iodine-131 therapy is quick, easy, moderatly expensive, avoids surgery, and is without significant risk in adults and probably teenagers. The larger doses required to give prompt and certain control generally induce hypothyroidism, and low doses are associated with a frequent requirement for retreatment or ancillary medical management over one to two years. 131-I is used as the primary therapy in most persons over age 40 and in most adults above age 21 if antithyroid drugs fail to control the disease. Treatment of children with 131-I is less common, as discussed later. It can be used in the elderly and those with co-morbidities with precautions.

Surgery, which was the main therapy until 1950, has been to a large extent replaced by 131-I treatment. As the high frequency of 131-I induced hypothyroidism became apparent, some revival of interest in thyroidectomy occurred. The major advantage of surgery is that definitive management is often obtained over an 8- to 12-week period, including preoperative medical control, and many patients are euthyroid after operation. Its well-known disadvantages include expense, surgery itself, and the risks of recurrent nerve and parathyroid damage, hypothyroidism, and recurrence. Nevertheless, if a skillful surgeon is available, surgical management may be used as the primary or secondary therapy in many young adults, as the secondary therapy in children poorly controlled on antithyroid drugs, in pregnant women requiring excessive doses of antithyroid drugs, in patients with significant exophthalmos, and in patients with coincident suspicious thyroid nodules. Early total thyroidectomy has been recommended for treating older, chronically ill patients with thyrotoxic storm if high-dose thionamide treatment, iopanoic acid, and glucocorticoids fail to improve the patient’s condition within 12 – 24 hours (36.4).

Two recent surveys reporting trends in therapeutic choices made by thyroidologists have been published [37]. In Europe, most physicians tended to treat children and adults first with antithyroid drugs, and adults secondarily with 131-I or less frequently surgery. Surgery was selected as primary therapy for patients with large goiters. 131-I was selected as the primary treatment in older patients. Most therapists attempted to restore euthyroidism by use of 131-I or surgery. In the United States, 131-I  is the initial modality of therapy selected by members of the American Thyroid Association for management of uncomplicated Graves’ disease in an adult woman [38]. Two-thirds of these clinicians attempt to give 131-I in a dosage calculated to produce euthyroidism, and one-third plan for thyroid ablation.

131-I THERAPY FOR THYROTOXICOSIS OF GRAVES’ DISEASE

Introduction-In many thyroid clinics 131-I therapy is now used for most patients with Graves’ disease who are beyond the adolescent years. It is used in most patients who have had prior thyroid surgery, because the incidence of complications, such as hypoparathyroidism and recurrent nerve palsy, is especially high in this group if a second thyroidectomy is performed. Likewise, it is the therapy of choice for any patient who is a poor risk for surgery because of complicating disease. Surgery may be preferred in patients with significant ophthalmopathy, often combined with prednisone prophylaxis.

Treatment of children-The question of an age limit below which RAI should not be used frequently arises. With lengthening experience these limits have been lowered. Several studies with average follow-up periods of 12 - 15 years attest to the safety of 131-I therapy in adults [ 39- 41]. In two excellent studies treated persons showed no tendency to develop thyroid cancer, leukemia, or reproductive abnormalities, and their children had no increase in congenital defects or evidence of thyroid damage [ 42- 44]. Franklyn and co workers recently reported on a population based study of 7417 patients treated with 131-I for thyrotoxicosis in England [44.1]. They found an overall decrease in incidence of cancer mortality, but a specific increase in mortality from cancer of the small bowel (7 fold) and of the thyroid (3.25) fold. The absolute risk remains very low, and it is not possible to determine whether the association is related to the basic disease, or to radioiodine treatment. Although there is much less data on long term results in children, there is a increased use of this treatment in teenagers age 15-18, as discussed below. The epidemic of thyroid cancer apparently induced by radioactive iodine isotopes in infants and children living around Chernobyl suggests caution in use of 131-I in younger children.
Since the possibility of a general induction of cancer by 131-I is of central concern, it is interesting to calculate the risk in children using the data presented by Rivkees et al (44.2) who are proponents of use of RAI for therapy in young children..The risk of death from any cancer due specifically to radiation exposure is noted by these authors to be 0.16%/rem for children, and the whole body radiation exposure from RAI treatment at age 10 to be 1.45 rem/mCi administered. Rivkees et al advise treatment with doses of RAI greater then 160 uCi/gram thyroid, to achieve a thyroidal radiation dose of at least 150Gy (about 15000 rads). Assuming a reasonable RAIU of 50% and gland size of 40 gm, the administered dose would thus be 40(gm) x 160uCi/gm x 2 (to account for 50% uptake) =12.8 mCi. Thus the long term cancer death risk would be 12.8 (mCi) x 1.45 rem (per mCi) x 0.16% (per rem) = 3%. For a dose of 15mCi the theoretical incremental risk of a later radiation-induced cancer mortality would be 4% at age 5, 2% at age 10, and 1% at age 15.
Whether or not accepting a specific  2-4% risk of death from any cancer because of  this treatment is of course a matter of judgment by the physician and family. However, this would seem to many persons to constitute a significant risk that might be avoided. We note that this is a thoretical risk, based on known effects of ioniing radiation to induce malignancies, but not so far proven in this setting.

Low 131-I uptake-Certain other findings may dictate the choice of therapy. Occasionally, the 131-I uptake is significantly blocked by prior iodine administration. The effect of iodide dissipates in a few days after stopping exposure, but it may take 3-12 weeks for the effect of amiodarone or IV contrast dyes to be lost. One may either wait for a few days to weeks until another 131-I tracer indicates that the uptake is in a treatable  range or use an alternative therapeutic approach such as antithyroid drugs.

Coincident nodule(s)-
Sometimes a patient with thyrotoxicosis harbors a thyroid gland with a configuration suggesting the presence of a malignant neoplasm. These patients probably should have surgical exploration. While FNA may exclude malignancy, the safety of leaving a highly irradiated nodule in place for many years is not established. Currently few patients who will have RAI therapy are subjected to ultrasonagraphy or scintiscaning. However Stocker et al. found that 12% of Graves’ patients had cold defects on scan, and among these half were referred for surgery. Six of 22, representing 2% of all Graves’ patients, 15% of patients with cold nodules, 25% of patients with palpable nodules, and 27% of those going to surgery, had papillary cancer in the location corresponding to the cold defect. Of these patients, one had metastasis to bone and two required multiple treatments with radioiodine. They argue for evaluating patients with a thyroid scintigram and further diagnostic evaluation of cold defects(44.3). Certainly any patient with GD in whom a thyroid nodule is detected, deserves consideration for surgical treatment

Ophthalmopathy-131-I therapy causes an increase in titers of TSH-R Abs, and anti-TG or TPO antibodies, which reflects an activation of autoimmunity. It probably is due to release of thyroid antigens by cell damage, and possibly destruction of intrathyroidal T cells. Many thyroidologists are convinced that 131-I therapy can lead to exacerbation of infiltrative ophthalmopathy, perhaps because of this immunologic response. Tallstedt and associates published data indicating that 131-I therapy causes exacerbation of ophthalmopathy in nearly 25% of patients, while surgery is followed by this response in about half as many.The same group conducted a second randomized trial (44.3) with a follow-up of 4 yr. Patients with a recent diagnosis of Graves’ hyperthyroidism were randomized to treatment with iodine-131 (163 patients) or 18 months of medical treatment (150 patients). Early substitution with L-T4 was given in both groups.: Worsening or development of eye problems was significantly more common in the iodine-131 treatment group (63 patients; 38.7%) compared with the medical treatment group (32 patients; 21.3%) (P < 0.001). This adverse effect of RAI therapy has since been confirmed in multiple meta-analyses of randomized studies (44.4-44.7) Thus, as described below, patients with significant ophthalmopathy may receive corticosteroids along with131-I, or may be selected for surgical management. The indications and contraindications for 131-I  therapy are given in Table 5.

Table 5-Indications and Contraindication for RAI Therapy

Indications
  • Any patients above a preselected age limit (possibly 15-18 yrs)
  • Patients who fail to respond to antithyroid drugs
  • Prior thyroid or other neck surgery
    Contraindications to surgery, such as severe heart, lung,or renal disease
    Women  intending to become pregnant (more than 6 months later)
General Contraindications
  • Pregnancy or lactation
  • Insufficient 131-I uptake due to prior medication or disease
  • Question of malignant thyroid tumor
  • Age below a preselected age limit, such as (possibly) age 15-18
    Patient concerns regarding radiation exposure
Other Possible Contraindications
  • Unusually large glands
  • Active exophthalmos

SELECTION OF 131-I Dosage

There are two basically different goals in 131-I dose selection. The traditional approach has been to attempt to give the thyroid 1) sufficient radiation to return the patient to euthyroidism, but not induce hypothyroidism. An alternative approach is to intend to
2) induce hypothyroidism, or euthyroidism and avoid any possible return of hyperthyroidism.

Background-The dosage initially was worked out by a trial-and-error method and by successive approximations. By 1950, the standard dose was 160 uCi 131-I per gram of estimated thyroid weight. Of course, estimating the weight of the thyroid gland by examination of the neck is an inexact procedure, but can now be made more accurate by use of ultrasound. Also, marked variation in radiation sensitivity no doubt exists and cannot be estimated at all. It was gratifying that in practice this dosage scheme worked well enough. In the early 1960s, it was recognized that a complication of RAI therapy was a high incidence of hypothyroidism. This reached 20 - 40% in the first year after therapy and increased about 2.5% per year, so that by 10 years 50 - 80% of patients had low function [45,46]. In an effort to reduce the incidence of late hypothyroidism, Hagen and colleagues reduced the quantity of 131-I to 0.08 mCi per gram of estimated gland weight [48]. No increase was reported in the number of patients requiring retreatment, and there was a substantial reduction in the incidence of hypothyroidism. Most of these patients were maintained on potassium iodide for several months after therapy, in order to ameliorate the thyrotoxicosis while the radioiodine had its effect [ 49, 50]. Patients previously treated with 131-I are sensitive to and generally easily controlled by KI. However KI often precipitates hypothyroidism in these patients, which may revert to hyperthyroidism when the KI is discontinued.

Over the years some effort was made to refine the calculation. Account was taken of uptake, half-life of the radioisotope in the thyroid, concentration per gram, and so on, but it is evident that the result in a given instance depends on factors that cannot be estimated precisely [47,]. One factor must be the tendency of the thyroid to return to normal if a dose of radiation is given that is large enough to make the gland approach, for a time, a normal functional state. In many patients, "cure" is associated with partial or total thyroid ablation. Although we, and many endocrinologists, attempt to scale the dose to the particular patient, some therapists believe it is futile, advocate giving up this attempt, and provide a standard dose giving up to 10000 rads to the thyroid(47.1). Leslie et al reported a comparison of fixed dose treatment and treatment adjusted for 24 hour RAIU, using low or high doses, and found no difference in outcome in either rate of control or induction of hypothyroidism on comparison of the methods. They favor the use of a fixed dose treatment with a single high or low dose (47.2).

Many attempts have been made to improve the therapeutic program by giving the RAI in smaller doses. Reinwein et al [51]. studied 334 patients several years after they had been treated with serial doses of less than 50 uCi 131-I per gram of estimated thyroid weight. One-third of these patients had increased levels of TSH, although they were clinically euthyroid. Only 3% were reported to be clinically hypothyroid.

Dosage adjustmentsmade to induce euthyroidism usually include a factor inc reassing with gland size, a standard dose in microCuries per gram, and a correction to account for 131-I uptake [52]. A"Low Dose Protocol" was designed to compensate for the apparent radiosensitivity of small glands and resistance of larger glands [53]. Using this approach, after one year, 10% of patients were hypothyroid, 60% are euthyroid, and 30% remained intrinsically toxic [53], although euthyroid by virtue of antithyroid drug treatment. At ten year follow-up, 40% were euthyroid and 60% hypothyroid. A problem with low-dose therapy is that about 25% of patients require a second treatment and 5% require a third. Although this approach reduces early hypothyroidism, it does so at a cost in time, money and patient convenience (Fig. 2). To answer these problems, patients can be re-treated, if need be, within six months, and propranolol and antithyroid drugs can be given between 131-I doses if needed. Unfortunately, experience shows that even low-dose 131-Itherapy is followed by a progressive development of hypothyroidism in up to 40 - 50% of patients ten years after therapy[ 54- 57].

Table 6. LOW Dosage Schedule for 131-I Therapy

Thyroid wt. in gms. uCi retained/gm
thyroid at 24h

Thyroid rads, avg.

 

10-20 40 3310
21-30 45 3720
31-40 50 4135
41-50 60 4960
51-60 70 5790
61-70 75 6200
71-80 80 6620
81-90 85 7030
91-100 90 7440
100 + 100 8270

Impressed by the need to retreat nearly a third of patients, a "Moderate Dose Protocol" was developed Table -6). This is a fairly conventional program with a mean dose of about 9 mCi. The 131-I dosage is related to gland weight and RAIU, and is increased as gland weight increases. The calculation used is as follows:

uCi given = (estimated thyroid weight in grams) X (uCi/g for appropriate weight from Table 6) / (fractional RAIU at 24 hours) (For readers who may find difficult the conversion of older units in Curies, rads, and rems to newer units of measurement, see Table -7.)

Table 6. MODERATE Dosage Schedule for 131-I Therapy

Thyroid wt. in gms. Planned uCi retained/gm
thyroid at 24h

Thyroid rads, avg.

 

10-20 80 6620
21-30 90 7440
31-40 100 8270
41-50 120 9920
51-60 140 11580
61-70 150 12400
71-80 160 13240
81-90 170 14060
91-100 180 14880
100 + 200 16540

 

Table 7. Conversion of International Units of Measurement

 

International Units Conversion Factors
Becquerel (Bq) 2.7 x 10 -11Curies (1mCi=37MBq, 100mCi= 3.7GBq)
Gray (Gy) 100 rads ( 1 rad= 0.01Gy)
Sievert (Sv) 100 rems (1 rem = 0.01 Sv)

Probably it is wise to do uptakes and treatment using either capsules or liquid isotope for both events. Rini et al have reported that RAIU done with isotope in a capsule appears to give significantly lower values (25 – 30% lower) than when the isotope is administered in liquid form, and this can significantly influence the determination of the dosage given for therapy(57.1). Berg et al report using a relatively similar protocol (absorbed doses of 100-120 Gy) and that 93% of their patients required replacement therapy after 1-5 years [57.2]. Many studies have presented methods for more accurately delivering a specific radiation dose to the thyroid, and report curing up to 90% of patients, with low incidence of recurrence or hypothyroidism(57.3, 57.4). Franklyn and co-workers analyzed their data on treatment of 813 hyperthyroid patients with radioactive iodide and corroborate many of the previously recognized factors involved in response. Lower dose (in this case 5 mCi), male gender, goiters of medium or large size and severe hyperthyroidism were factors that were associated with failure to cure after one treatment. They suggest using higher fixed initial doses of radioiodine for treating such patients (58.2), as do Leslie et al(58.4). Santos et al (58.4) compared fixed doses of 10 and 15mCi and found no difference in outcome at 12 months post treatment.  These authors suggest a standard 10mCi dose, with the larger dose reserved for larger glands.

Planned thyroid partial or complete ablation-All attempts to induce euthyroidism by a calculated moderate dose protocol end up with some patients hypothyroid, and others with persistent hyperthyroidism requiring further treatment. At this time many physicians giving 131-I therapy make no attempt to achieve euthyroidism, and instead use  a dose sufficient to largely destroy the thyroid, followed by L- T4 replacement therapy [58]. For example, a dose is given that will result in 7-20 mCi retained at 24 hrs, which is intended to induce hypothyroidism, accepting that in some (or many) patients this will ablate the thyroid completely. A dose of 30 Mci was found to  offer a slightly higher cure rate, not surprisingly, at one year than 15 Mci (95 vs 74% (58.1), They argue that this is realistic and preferable since it offers 1) near certainty of prompt control, 2) avoids any chance of persistent or recurrent disease, 3)there is no benefit in having residual thyroid  tissue, and 4) hypothyroidism is inevitable in most patients given RAI. Probably many patients given this treatment do in fact have some residual thyroid tissue that is either heavily damaged or reduced in amount so that it can not produce normal amounts of hormone. So far there is no evidence, in adults, that this residual radiated tissue will develop malignant change. There is no certainty at this time that one approach is better than the other. It may be worth remembering that over 50% of patients given calculated moderate dose therapy remain euthyroid after ten years and can easily be surveyed at yearly intervals for hypothyroidism.
When giving large doses of 131-I it is prudent to calculate the rads delivered to the gland (as above), which can reach 40-50,000rads. Such large doses of radiation can cause clinically significant radiation thyroiditis, and occasionally damage surrounding structures.
And lastly, a speculation. Practitioners comment that the incidence of serious ophthalmopathy seems to be less that in former decades. Prompt diagnosis and therapy might contribute to such a change. Another factor could be the  more common ablation of the thyroid during therapy for Graves disease, since this should over time reduce exposure  of patient’s immune system to thyroid antigens.

Lithium with RAI therapy- Although rarely used, RAI combined with lithium is safe and more effective than RAI alone in the cure of hyperthyroidism due to Graves’ disease, probably because it it causes greater retention of RAI within the thyroid gland.. Bogazzi et al (58.5)reported a study combining lithium with RAI therapy. MMI treatment was withdrawn 5 days prior to treatment, Two hundred ninety-eight patients were treated with RAI plus lithium (900 mg/d for 12 d starting 5 days prior to 131-I treatment) and 353 with RAI alone. RAI dosage was 260mCi/g estimated thyroid weight, corrected for RAIU (done on lithium).. All patients receive prednisone 0.5mg/kg/day, beginning on day 7 after RAI, tapering over 2 months. Patients treated with RAI plus lithium had a higher cure rate (91.0%) than those treated with RAI alone (85.0%, P = 0.030). In addition, patients treated with RAI plus lithium were cured more rapidly (median 60 d) than those treated with RAI alone (median 90 d, P = 0.000). Treatment with lithium inhibited the serum FT4 increase seen after methimazole withdrawal and RAI therapy.

Pretreatment with antithyroid drugs--Patients are often treated directly after diagnosis, without prior therapy with antithyroid drugs. This is safe and common in patients with mild hyperthyroidsm and especially those without eye problems. However often physicians give antithyroid drugs before 131-I treatment in order to deplete the gland of stored hormone and to restore the FTI to normal before 131-I therapy. This offers several benefits. The possibility of 131-I induced exacerbation of thyrotoxicosis is reduced, the patient recovers toward normal health, and there is time to reflect on the desired therapy and review any concerns about the use of radioisotope for therapy. If the patient has been on antithyroid drug, it is discontinued two days before RAIU and therapy. Patients can be treated while on antithyroid drug, but this reduces the dose retained, reduces the post-therapy increment in hormone levels, and reduces the cure rate, so seems illogical(58.6) . When antithyroid drugs are discontinued the patient’s disease may exacerbate, and this must be carefully followed. Beta blockers can be given in this interim, but there is no reason for a prolonged interval between stopping antithyroid drug, and 131-I therapy, unless there is uncertainty about the need for the treatment. Pretreatment with antithyroid drug does not appear in most studies to reduce the efficacy of 131-I treatment. [59] but the debate about the effect of antithyroid drug pretreatment on the efficacy of radioactive iodine therapy continues. In recent studies in which patients were on or off antithyroid therapy, which was discontinued four days, or 1-2 days before treatment, there was no effect on the efficacy of treatment at a one year endpoint (59.1,59.2, 59,3). In another study Bonnema et al found that PTU pretreatment , stopped 4 days prior to 131-I, reduced the efficacy of 131-I(59.6).

Pretreatment is usually optional but is logical in patients with large glands and severe hyperthyroidism. Antithyroid drug therapy does reduce the pretreatment levels of hormone and reduces the rise in thyroid hormone level that may occur after radioactive iodide treatment. This certainly could have a protective effect in individuals who have coincident serious illness such as coronary artery disease, or perhaps individuals who have very large thyroid glands (59.3). It is indicated in two circumstances. In patients with severe heart disease, an 131-I- induced exacerbation of thyrotoxicosis could be serious or fatal. Pretreatment may reduce exacerbation of eye disease (see below), and it does reduce the post-RAI increase in antibody titers(59.1,59.31). The treatment dose of 131-I is best given as soon as possible after the diagnostic RAIU in order to reduce the period in which thyrotoxicosis may exacerbate without treatment, and since any intake of iodine (from diet or medicines or tests) would alter uptake of the treatment dose (59.4), and 2 days seems sufficient.

Post 131-I treatment management--Many patients remain on beta-blockers but require no other treatment after 131-I therapy. Antithyroid drugs can be reinstituted after 5 ( or preferably 7 ) days, with minimal effect on retention of the treatment dose of 131-I.

Alternatively, one may prescribe antithyroid drug (typically 10 mg methimazole q8h) beginning one day after administration of 131-I and add KI (2 drops q8h) after the second dose of methimazole. KI is continued for two weeks, and antithyroid drug as needed. This promotes a rapid return to euthyroidism, but by preventing recirculation of 131-I it can lower the effectiveness of the treatment. This method has been employed in a large number of patients, and is especially useful in patients requiring rapid control- for example, with CHF. A typical response is shown in Fig -3. It also has provided the largest proportion of patients remaining euthyroid at 10 years after therapy, in comparison to other treatment protocols. Glinoer and Verelst also report successful use of this strategy [59.1]. As noted, antithyroid drugs may be given starting 7-10 days after RAI without significantly lowering the radiation dose delivered to the gland.

Treatment using 125-I was tried as an alternative to 131-I, because it might offer certain advantages [60]. 125-I is primarily a gamma ray emitter, but secondary low-energy electrons are produced that penetrate only a few microns, in contrast to the high-energy beta rays of 131-I. Thus, it might theoretically be possible to treat the cytoplasm of the thyroid cell with relatively little damage to the nucleus. Appropriate calculations indicated that the radiation dose to the nucleus could be perhaps one-third that to the cytoplasm, whereas this difference would not exist for 131-I. Extensive therapeutic trials have nonetheless failed to disclose any advantage thus far for 125I. Larger doses -- 10-20 mCi -- are required, increasing whole body radiation considerably [ 61, 62].


SAFETY PRECAUTIONS AFTER 131-I THERAPY


Doses of 131-I up to 33 mCi can be given to an outpatient basis, and this level is rarely exceeded in treatment of Graves’ disease. However patients must be given advice (written if possible) on precautions to be followed to prevent unneccessary or excessive exposure of other individuals by radiaactivity administered to the patient. For maximum safety, patients who have received 20 mCi should avoid extended time in public places for 1 day, maximize distance (6 feet) from children and pregnant women for 2 days, may return to work after 1 day, sleep in a separate (6-feet separation) bed from adults for 8 days, sleep in a separate bed from pregnant partners, infant, or child for 20 days, and avoid contact with body fluids (saliva, urine) for at least one week. Lower therapeutic doses require proportionally more moderate precautions. The basic NRC rule is that patients may be released from hospital when (1) the 131I measured dose rate is ≤7 mrem/hr at 1 m, or (2) when the expected total dose another person would receive is unlikely to exceed 500 mrem (5 mSv). Written precaution instructions are required If 100 mrem (1 mSv) may be exceeded in any person. This topic is well covered in articles by Sisson et al
(http://www.ncbi.nlm.nih.gov/pubmed/21417738) andLiu et al (62.1).

 

Course After Treatment-

If adequate treatment has been given, the T4 level falls progressively, beginning in one to three weeks.. Labeled thyroid hormones, iodotyrosines, and iodoproteins appear in the circulation [63,63.1]. TG is released, starting immediately after therapy. Another iodoprotein, which seems to be an iodinated albumin, is also found in plasma. This compound is similar or identical to a quantitatively insignificant secretion product of the normal gland. It comprises up to 15% or more of the circulating serum 131-I in thyrotoxic patients [64]. It is heavily labeled after 131-I therapy, and its proportional secretion is probably increased by the radiation. Iodotyrosine present in the serum may represent leakage from the thyroid gland, or may be derived from peripheral metabolism of TG or iodoalbumin released from the thyroid.

The return to the euthyroid state usually requires at least two months, and often the declining function of the gland proceeds gradually over six months to a year. For this reason, it is logical to avoid retreating a patient before six months have elapsed unless there is no evidence of control of the disease. While awaiting the response to131-I  the symptoms may be controlled by propranolol, antithyroid drugs, or iodide. Hypothyroidism develops transiently in 10 - 20% of patients, but thyroid function returns to normal in most of these patients in a period ranging from three to six months. These patients rarely become toxic again. Others develop permanent hypothyroidism and require replacement therapy. It is advantageous to give the thyroid adequate time to recover function spontaneously before starting permanent replacement therapy. This can be difficult for the patient unless partial T4 replacement is given. Unfortunately, one of the common side effects of treating hyperthyroidism is weight gain, averaging about 20 lbs through four years after treatment (64.1).

Patients may develop transient increases in FTI and T3 at 2-4 months after treatment [63.1], sometimes associated with enlargement of the thyroid. This may represent an inflammatory or immune response to the irradiationinduced thyroid damage, and the course may change rapidly with a dramatic drop to hypothyroidism in the 4-5th month.

Hypothyroidism may ultimately be inescapable after any amount of radiation that is sufficient to reduce the function of the hyperplastic thyroid to normal [65]. Many apparently euthyroid patients (as many as half) have elevated serum levels of TSH long after 131-I therapy, with "normal" plasma hormone levels [66]. An elevated TSH level with a low normal T4level is an indicator of changes progressing toward hypothyroidism [67]. The hypothyroidism is doubtless also related to the continued autoimmune attack on thyroid cells. Hypofunction is a common end stage of Graves’ disease independent of 131-I use; it occurs spontaneously as first noted in 1895(!) [68] and in patients treated only with antithyroid drugs [69]. Just as after surgery, the development of hypothyroidism is correlated positively with the presence of antithyroid antibodies.

During the rapid development of postradiation hypothyroidism, the typical symptoms of depressed metabolism are evident, but two rather unusual features also occur. The patients may have marked aching and stiffness of joints and muscles. They may also develop severe centrally located and persistent headache. The headache responds rapidly to thyroid hormone therapy. Hair loss can also be dramatic at this time.

In patients developing hypothyroidism rapidly, the plasma T4 level and FTI accurately reflect the metabolic state. However, it should be noted that the TSH response may be suppressed for weeks or months by prior thyrotoxicosis; thus, the TSH level may not accurately reflect hypothyroidism in these persons and should not be used in preference to the FTI or FT4.

If permanent hypothyroidism develops, the patient is given replacement hormone therapy and is impressed with the necessity of taking the medication for the remainder of his or her life. Thyroid hormone replacement is not obligatory for those who develop only temporary hypothyroidism, although it is possible that patients in this group should receive replacement hormone, for their glands have been severely damaged and they are likely to develop hypothyroidism at a later date. Perhaps these thyroids, under prolonged TSH stimulation, may tend to develop adenomatous or malignant changes, but this has not been observed. Many middle-aged women gain weight excessively after radioactive iodide treatment of hyperthyroidism. Usually such patients are on what is presumed to be appropriate T4 replacement therapy. Tigas et al note that such weight gain is less common after ablative therapy for thyroid cancer, in which case larger doses of thyroxine are generally prescribed. Thus they question whether the excessive weight gain after radioactive iodide treatment of Graves’ disease is due to the fact that insufficient thyroid hormone is being provided, even though TSH is within the “normal” range. They suggest that restoration of serum TSH to the reference range by T4 alone may not constitute adequate hormone replacement [ 69a}. We noted above that the correct reference range for TT4 and FT4, when the patient is on replacement T4, should  be 20% higher than normal.

Permanent replacement therapy (regardless of the degree of thyroid destruction) for children who receive 131-I has a better theoretical basis. In these cases, it is advisable to prevent TSH stimulation of the thyroid and so mitigate any possible tendency toward carcinoma formation.

Exacerbation of thyrotoxicosis-During the period immediately after therapy, there may be a transient elevation of the T4 or T3 level [70], but usually the T4 level falls progressively toward normal. Among  treated hyperthyroid patients with Graves’ disease, only rare exacerbations of the disease are seen. These patients may have cardiac problems such as worsening angina pectoris, congestive heart failure, or disturbances of rhythm such as atrial fibrillation or even ventricular tachycardia. Radiation-induced thyroid storm and even death have unfortunately been reported [71- 73]. These untoward events argue for pretreatment of selected patients who have other serious illness, especially cardiac disease, with antithyroid drugs prior to 131-I therapy.

 

Other Problems Associate With 131-I Therapy

The immediate side effects of 131-I therapy are typically minimal. As noted above, transient exacerbation of thyrotoxicosis can occur, and apparent thyroid storm has been induced within a day (or days) after 131-I therapy. A few patients develop mild pain and tenderness over the thyroid and, rarely, dysphagia. Some patients develop temporary hair loss, but this condition occurs two to three months after therapy rather than at two to three weeks, as occurs after ordinary radiation epilation. Hair loss also occurs after surgical therapy, so that it is a metabolic rather than a radiation effect. If the loss of hair is due to the change in metabolic status, it generally recovers in a few weeks or months. However hair thinning, patchy alopecia, and total alopecia, are all associated with Graves’ Disease, probably as another auto-immune processes. In this situation the prognosis for recovery is less certain, and occasionally some other therapy for the hair loss (such as steroids) is indicated. Permanent hypoparathyroidism has been reported very rarely as a complication of RAI therapy for heart disease and thyrotoxicosis[ 74- 76]. Patients treated for hyperthyroidism with 131-I received approximately 39 microGy/MBq administered (about 0.144rad/mCi) of combined beta and gamma radiation to the testes. This is reported to cause no significant changes in FSH. Nevertheless, testosterone declines transiently for several months, but there is no variation in sperm motility or % abnormal forms (76.1). Long term studies of patients after RAI treatment by Franklyn et al (76.2) show a slight increase in mortality which appears to be related to cardiovascular disease, possibly related to periods of hypothyroidism.

 

Worsening of ophthalmopathy after RAI---In contrast to the experience with antithyroid drugs or surgery, antithyroid antibodies including TSAb levels increase after RAI [ 77, 78]. (Fig. 11-4, above). Coincident with this condition, exophthalmos may be worsened [79].(Fig. 11-5, below). This change is most likely an immunologic reaction to discharged thyroid antigens.The relationship of radiation therapy to exacerbation of exophthalmos has beem questioned], but much recent data indicates that there is a definite correlation[ 79, 80, 80.1, 80.2, 80.3]. Many therapists consider "bad eyes" to be a relative contraindication to RAI. Induction of hypothyroidism, with elevation of TSH, may contribute to worsening of ophthalmopathy. This offers support for early induction of T4 replacement (80.3).
Pretreatment with antithyroid drugs has been used empirically in an attempt to prevent this complication. Its benefit, if any, may be related to an immunosuppressive effect of PTU, described below. Treatment with methimazole before and for three months after I-131 therapy has been shown to help prevent the treatment-induced rise in TSH-R antibodies which is otherwise seen[81].

Prophylaxis with prednisone after 131-I helps prevent exacerbation of exophthalmos, and this approach is now the standard approach in patients who have significant exophthalmos at the time of treatment [ 82, 82.1]. (Fig. 6, below) The recommended dose is 30 mg/day for one month, tapering then over 2-3 months. Of course prednisone or other measures can be instituted at the time of any worsening of ophthalmopathy. In this instance doses of 30-60 mg/day are employed, and usually are required over several months. While treatment with prednisone helps prevent eye problems, it does not appear to reduce the effectiveness of RAI in controlling the hyperthyroidism(82.2).

Thyroidectomy
, with total removal of the gland, should be considered for patients with serious active eye disease. Operative removal of the thyroid is followed by gradual diminution is TSH-R antibodies.(82.3 ), and as shown by Tallstedt is associated with a lower incidence of worsening eye problems than is initial RAI treatment. Several studies document better outcomes of ophthalmopathy in patients with GD who have total thyroidectomy vs those treated by other means(82.4, 82.5, 82.6).

 

 

Failure of 131-I to cure thyrotoxicosis occurs occasionally even after 2 or 3 treatments, and rarely 4 or 5 therapies are given. The reason for this failure is usually not clear. The radiation effect may occur slowly. A large store of hormone in a large gland may be one cause of a slow response. Occasional glands having an extremely rapid turnover of 131-I  requiring such high doses of the isotope that surgery is preferable to continued 131-I therapy and its attendant whole body radiation. If a patient fails to respond to one or two doses of 131-I, it is important to consider that rapid turnover may reduce the effective radiation dose. Turnover can easily be estimated by measuring RAIU at 4, 12, 24, and 48 hours, or longer. The usual combined physical and biological half-time of 131-I retention is about 6 days. This may be reduced to 1 or 2 days in some cases, especially in patients who have had prior  therapy or subtotal thyroidectomy. If this rapid release of 131-I is found, and 131-I therapy is desired, the total dose given must be increased to compensate for rapid release. A rough guide to this increment is as follows:

Increased dose = usual dose X ( (usual half time of 6 days) / (observed half time of "X" days) )

Most successfully treated glands return to a normal or cosmetically satisfactory size. Some large glands remain large, and in that sense may constitute a treatment failure. In such a situation secondary thyroidectomy could be done, but it is rarely required in practice.

Long term care- Patients who have been treated with RAI should continue under the care of a physician who is interested in their thyroid problem for the remainder of their lives. The first follow-up visit should be made six to eight weeks after therapy. By this time, it will often be found that the patient has already experienced considerable improvement and has begun to gain weight. The frequency of subsequent visits will depend on the progress of the patient. Symptoms of hypothyroidism, if they develop, are usually not encountered until after two to four months, but one of the unfortunate facts of RAI therapy is that hypothyroidism may occur almost any time after the initial response.

 

HAZARDS OF 131-I TREATMENT

In the early days of RAI treatment for Graves’ disease, only patients over 45 years of age were selected for treatment because of the fear of ill effects of radiation. This age limit was gradually lowered, and some clinics, after experience extending over nearly 40 years, have now abandoned most age limitation. The major fear has been concern for induction of neoplasia, as well as the possibility that 131-I might induce undesirable mutations in the germ cells that would appear in later generations.

Table 8. Gonadal Radiation Dose (in Rads) From Diagnostic Procedures and 131-I Therapy

 

Proceedure Males- median Females- median
Barium meal 0.03 0.34
IV pyelogram 0.43 0.59
Retrograde pyelogram 0.58 0.52
Barium enema 0.3 0.87
Femur xray 0.92 0.24
131-I-therapy, 5mCi usually <1.6 usually <1.6
Adapted from Robertson and Gorman [95]

 

Carcinogenesis

Radiation is known to induce tumor formation in many kinds of tissues and to potentiate the carcinogenic properties of many chemical substances. Radiation therapy to the thymus or nasopharyngeal structures plays an etiologic role in thyroid carcinoma both in children and in adults[ 83- 85]. 131-I radiation to the animal thyroid can produce tumors, especially if followed by PTU therapy [86]. Cancer of the thyroid has appeared more frequently in survivors of the atomic explosions at Hiroshima and Nagasaki than in control populations [87]. Thyroid nodules, some malignant, have appeared in the natives of Rongelap Island as the result of fallout after a nuclear test explosion in which the radiation cloud unexpectedly passed over the island [88].

 

Thyroid cancer following 131-I treatment?


The experience at 26 medical centers with thyroid carcinoma after 131-I therapy was collected in a comprehensive study of the problem. A total of 34,684 patients treated in various ways were included. Beginning more than one year after 131-I therapy, 19 malignant neoplasms were found; this result did not differ significantly from the frequency after subtotal thyroidectomy. Thyroid adenomas occurred with increased frequency in the 131-I treated group, and the frequency was greatest when the patients were treated in the first two decades of life [39]. Holm et al [41] have thoroughly examined the history of a large cohort of 131-I-treated patients in Sweden and similarly found no evidence for an increased incidence of thyroid carcinoma or other tumors. For reasons that are not clear, the injury caused by 131-I therapy for Graves’ disease seems to induce malignant changes infrequently.. This may be because the treatment has largely been given to adults with glands less sensitive to radiation, because damage from 131-Itherapy is so severe that the irradiated cells are unable to undergo malignant transformations, or because all cells are destroyed, or possibly because of the slow rate at which the dose is delivered [89]. In up to one-half of patients followed for 5-10 years, there may be no viable thyroid cells remaining. We note that two studies reported above extend through an average follow-up period of 15 years. As described above [44.1], a recent report by Franklyn and coworkers indicated that there is an increased (3.25 fold) risk of mortality from cancer of the thyroid (and also bowel) after RAI, detected in along term follow up of a very large patient cohort. However it remains uncertain that this is related to hyperthyroidism per se, or radioiodine therapy.

While these data are reassuring in regard to 131-I use in adults, Chernoby made it clear that its use in children can not be considered safe. Children in the area surrounding Chernobyl have developed a hugely increased incidence of thyroid carcinoma predominately due to ingestion of iodine-131 [89.]. The latency has been about 5 years, and younger children are most affected. Risk is probably linearly related to dose. It is apparent that low doses, possibly down to 20 rads, produce malignant change in children(89.2).Risk of carcinogenesis decreases with increasing age at exposure, and is much less common after age 12. However some data indicates that an increased incidence of thyroid carcinoma is seen even among adults exposed at Chernobyl.

 

Leukemia

The incidence of leukemia among patients treated with RAI for Graves’ disease has not exceeded that calculated from a control group [90]. This problem was also studied by the consortium of 26 hospitals [91]. The incidence of leukemia in this group was slightly lower than in a control group treated surgically, but slightly higher in the latter surgical group than in the general population.

 

Genetic Damage


In the group of RAI-treated patients, there has been no evidence of genetic damage, although, as will shortly be seen, this problem cannot be disregarded. In the United States, about 100 x 106 children will be born to a population of over 200 x 106 persons. Approximately 4% of these children will have some recognizable defect at birth. Of these, about one-half will be genetically determined or ultimately mutational, and represent the effects of the baseline mutation rate in the human species. These mutations are attributed in part to naturally occurring radiation.

All penetrating radiation, from whatever sources, produces mutations. The effects may vary with rate of application, age of the subject, and no doubt many other factors, and are partially cumulative. Nearly all of these mutations behave as recessive genetic factors; perhaps 1% are dominant. Almost all are minor changes, and those produced by experimental radiation are the same as those produced by natural radiation.

Whether or not mutations are bad is in essence a philosophic question. Most of us would agree that the cumulative effect of mutations over past eras brought the human race to its present stage of development. However, most mutations, at least those that are observable, are detrimental to individual human adaptation to the present environment. In terms of the human population as a whole, detrimental mutant genes must be eliminated by the death of the carrier. We can agree that an increase in mutation rate is not desirable. It is hardly worth considering the pros and cons of the already considerable spontaneous mutation rate.

In mice, the occurrence of visible genetic mutations in any population group is probably doubled by acute exposure of each member of the group over many generations to about 30 - 40 rads, or by chronic exposure to 100 - 200 rads [92]. This radiation dosage is referred to as the doubling dose. Ten percent of this increase in mutations might be expressed in the first-generation offspring of radiated parents, the remainder gradually appearing over succeeding generations. The change in mutation rate in Drosophila is in proportion to the dosage in the range above 5 rads. Data from studies of mice indicate that at low exposures (from 0.8 down to 0.0007 rads/min), the dose causing a doubling in the spontaneous rate of identifiable mutations is 110 rads [92,93]. Linearity, although surmised, has not been demonstrated at lower doses.

At present, residents of the United States receive about 300 mrad/year, or 9 rad before age 30, the median parental age. Roughly half of this dose is from natural sources and half from medical and, to a lesser extent, industrial exposure. The National Research Council has recommended a maximum exposure rate for the general population of less than 10 rad above background before age 30. (The present level may therefore approach this limit.)

The radiation received by the thyroid and gonads during 131-I therapy of thyrotoxicosis can be estimated from the following formula:

Total beta radiation dose = 73.8 x concentration of 131-I in the tissue (µCi/g) x average beta ray energy (0.19 meV) x effective isotope half-life

For illustration, we can assume a gland weight of 50 g, an uptake of 50% at 24 hours, a peak level of circulating protein-bound iodide (PB 131-I) of 1% dose/liter, an administered dose of 10 mCi, a thyroidal iodide biologic half-life of 6 days, and a gamma dose of about 10% of that from beta rays. On this basis, the thyroid receives almost 8200 rads, or roughly 1,600 rads/mCi retained. The gonadal dose, being about one-half the body dose, would approximate 4 rads, or roughly 0.4 rads/mCi administered.

If the radiation data derived from Drosophila and lower vertebrates are applied to human radiation exposure (a tenuous but not illogical assumption), the increased risk of visible mutational defects in the progeny can be calculated. On the basis of administration to the entire population of sufficient 131-I to deliver to the gonads 2 rads or 2% of the doubling dose (assumed to be the same as in the mouse), the increase in the rate of mutational defects would ultimately be about 0.04%, although only one-tenth would be seen in the first generation. Obviously only a minute fraction of the population will ever receive therapeutic 131-I. The incidence of thyrotoxicosis is perhaps 0.03% per year, or 1.4% for the normal life span. At least one-half of these persons will have their disease after the childbearing age has passed. Although most of them will be women, this fact does not affect the calculations after a lapse of a few generations. Assuming that the entire exposed population receives 131-I therapy in an average amount of 5 mCi, the increase in congenital genetic damage would be on the order of 0.02 (present congenital defect rate) x 0.04 ( 131-I radiation to the gonads as a fraction of the doubling dose) x 0.014 (the fraction of the population ever at risk) x 0.5 (the fraction of patients of childbearing age) = 0.0000056.

This crude estimate, developed from several sources, also implies that, if all patients with thyrotoxicosis were treated with 131-I, the number of birth defects might ultimately increase from 4 to 4.0006%. This increase may seem startlingly small or large, depending on one’s point of view, but it is a change that would be essentially impossible to confirm from clinical experience.

Unfortunately, it is more difficult to provide a reliable estimate of the increased risk of genetic damage in the offspring of any given treated patient. Calculations such as the above simply state the problem for the whole population. Since most of the mutations are recessive, they appear in the children only when paired with another recessive gene derived from the normal complement carried by all persons. Assuming that only one parent received radiation from 131-I therapy amounting to 2% of the doubling dose, the risk of apparent birth defects in the patient’s children might increase from the present 4.0% to 4.008%.

0.02 (present genetic defect rate) x 0.04 (fraction of the doubling dose) x 0.1 (fraction of defects appearing in the first generation) = 0.00008, or an increase from 4.0% to 4.008%.

Similar estimates can be derived by considering the number of visible mutations derived from experimental radiation in lower species.[ 92, 93]

6 x 10-8 (mutations produced per genetic locus per rad of exposure) x 104 (an estimate of the number of genetic loci in humans) x 2 (gonadal radiation in rads as estimated above) x 0.1 (fraction of mutations appearing in the first generation) = 0.00012 or 0.012%

On this basis, the increase in the birth defect rate would be from 4.0% to 4.012%. One important observation stemming from these calculations is that large numbers of children born to irradiated parents must be surveyed if evidence of genetic damage is ever to be found. Reports of "no problems" among 30 to 100 such children are essentially irrelevant when one is seeking an increase in the defect rate of about 4.0% to about 4.008%.

These statistics are presented in an attempt to give some quantitation to the genetic risk involved in 131-I therapy, and should not be interpreted as in any sense exact or final. The point we wish to stress is that radiation delivered to future parents probably will result in an increased incidence of genetic damage, but an increase so slight that it is difficult to measure. Nonetheless, the use of 131-I for large numbers of women who subsequently become pregnant will inevitably introduce change in the gene pool.

In considering the significance of these risks, one must remember that the radiation exposure to the gonads from the usual therapeutic dose of 131-I may be only one or two times that produced during a procedure such as a barium enema [ 94, 95] and similar to the 10 rads received from a CAT scan. These examinations are ordered by most physicians without fear of radiation effect ( Table 11-8).

When assessing the risks of 131-I therapy, one must, of course, consider the risks of any alternative choice of procedure. Surgery carries a small but finite mortality, as well as a risk of permanent hypoparathyroidism, hypothyroidism, and vocal cord paralysis. Some of these risks are especially high in children, the group in which radiation damage is most feared. Some physicians have held that 131-I therapy should not be given to patients who intend subsequently to have children. In fact, there is at present no evidence to support this contention, as discussed above. Chapman [44] studied 110 women treated with 131-I who subsequently became pregnant and were delivered of 150 children. There was no evidence of any increase in congenital defects or of accidents of pregnancy. Sarkar et al [96] also found no evidence of excess abnormalities among children who received 131-I therapy for cancer. Other studies have confirmed the apparent lack of risk[ 42, 43]. It should be noted that no increase in congenital abnormalities has been detected among the offspring of persons who received much larger radiation doses during atomic bomb explosions [97].

Often the patient wishes to know about the possibility of carcinogenesis or genetic damage. These questions must be fully but delicately handled. It is not logical to treat a patient of childbearing age with 131-I and have the patient subsequently live in great fear of bearing children. These problems and considerations must be faced each time a patient is considered for RAI therapy.

Pregnancy and 131-I Pregnancy is an absolute contraindication to 131-I therapy. The fetus is exposed to considerable radiation from transplacental migration of 131-I, as well as from the isotope in the maternal circulatory and excretory systems. In addition, the fetal thyroid collects 131-I after the 12th week of gestation and may be destroyed. The increased sensitivity of fetal structures to radiation damage has already been described. Physicians treating women of childbearing age with 131-I should be certain that the patients are not pregnant when given the isotope. Therapy during or immediately after a normal menstrual period or performance of a pregnancy test are appropriate precautions if pregnancy is possible. Women should be advised to avoid pregnancy for at least six months after treatment with RAI, since it usually takes this long to be certain that retreatment will not be needed.

TREATMENT OF THYROTOXICOSIS WITH DRUGS

Drug therapy for thyrotoxicosis was introduced by Plummer when he observed that the administration of iodide ameliorated the symptoms of this disease [98]. (Fig 7) Administration of iodide has since been used occasionally as the complete therapeutic program for thyrotoxicosis, and widely as an adjunct in preparing patients for subtotal thyroidectomy. In 1941 the pioneering observations of MacKenzie and MacKenzie [99] and Astwood [100] led to the development of the thiocarbamide drugs, which reliably block the formation of thyroid hormones. It soon became apparent that, in a certain proportion of patients with Graves’ disease, use of these drugs could induce a prolonged or permanent remission of the disease even after the medication was discontinued. It is not yet understood why a temporary reduction in the formation of thyroid hormone should result in reduction of TSHR antibodies, and permanent amelioration of the disease.

The antithyroid drug initially introduced for treatment of Graves’ disease was thiourea, but this drug proved to have a large number of undesirable toxic effects. Subsequently a number of derivatives and related compounds were introduced that have potent antithyroid activity without the same degree of toxicity. Among these substances are propyl- and methylthiouracil, methimazole, and carbimazole. In addition to this class of compounds, potassium perchlorate has been used in the treatment of thyrotoxicosis, but is infrequently employed for this purpose because of occasional bone marrow depression. This drug prevents the concentration of iodide by the thyroid. Beta adrenergic blockers such as propranolol have a place in the treatment of thyrotoxicosis. These drugs alleviate some of the signs and symptoms of the disease but have little or no direct effect on the metabolic abnormality itself. They do not uniformly induce a remission of the disease and can be regarded as adjuncts, not as a substitute for more definitive therapy.

Mechanism of Action- Antithyroid drugs inhibit thyroid peroxidase, and PTU (not methimazole) has the further beneficial action of inhibiting T4 to T3 conversion in peripheral tissues. Antithyroid therapy is associated with a reduction in circulating antithyroid antibody titers [101], and anti-receptor antibodies [77, 78, 102]. Studies by MacGregor and colleagues [103] indicate that antibody reduction also occurs during antithyroid therapy in patients with thyroiditis maintained in a euthyroid state, thus indicating that the effect is not due only to lowering of the FT4 in Graves’ disease. These authors also found a direct inhibitory effect of PTU and carbimazole on antithyroid antibody synthesis in vitro and postulate that this is the mechanism for diminished antibody levels [104]. Other data argue against this hypothesis [105, 105.1].

Antithyroid drug therapy is also associated with a prompt reduction in the abnormally high levels of activated T lymphocytes in the circulation [106], although Totterman and co-workers found that this therapy caused a prompt and transient elevation of activated T suppressor lymphocytes in blood [107]. During antithyroid drug treatment the reduced numbers of T suppressor cells reported to be present in thyrotoxic patients return to normal [106, 108]. Antithyroid drugs do not directly inhibit T cell function [109]. All of these data argue that antithyroid drugs exert a powerful beneficial immunosuppressive effect on patients with Graves’ disease. While much has been learned about this process, the exact mechanism remains uncertain. Evidence that antithyroid drugs exert their immunosuppressive effect by a direct inhibition of thyroid cell production of hormones has been reviewed by Volpe [109].

Long-Term Antithyroid Drug Therapy with Thiocarbamides

Propylthiouracil warning-Propylthiouracil and methimazole have for years been considered effectively interchangeable, and liver damage was considered a very rare problem. Recently a commission appointed by the FDA reevaluated this problem, and concluded that the rare but severe complications of liver failure needing transplantation, and death, were sufficient to contraindicate the use of PTU as the normal first-line drug (109.1). The Endocrine  Society and other advisory groups have suggested that methimazole be used for treatment except in circumstances of inavailability of the drug, patient allergy, or pregnancy. Because of the association of scalp defects and probably a severe choanal syndrome with administration of methimazole during the first 12 weeks of pregnancy, current advice is to avoid use of methimazole during the first trimester, for instance giving PTU during the first trimester, and then switching to methimazole.

Selection of patients-Many patients with Graves’ disease under age 40 - 45 are given a trial of therapy with one of the thiocarbamide drugs. Younger patients, and those with recent onset of disease, small goiters [110], and mild disease, are especially favorable candidates, since they tend to enter remission most frequently (110.1). It is generally found that one-fourth to one-third of these patients who satisfactorily complete a one year course have a long term or permanent remission. The remainder need repeated courses of drug therapy, must be maintained on the drug for years or indefinitely [111, 112], or must be given some other treatment. It appears that the percentage of patients responding has progressively fallen over the past years from about 50% to at present 25 - 30%[113, 114]. This change was thought to reflect an alteration in iodide in our diet [115], which increased from about 150 µg/day in 1955 to 300 - 600 µg/day. However other factors including greater precision in diagnosis and more complete data probably play major roles in establishing the response rate recognized at present. Some physicians do not consider antithyroid drug therapy to be the most efficacious means of treating thyrotoxic patients because of the high recurrence rate.

Therapeutic program-Patients are initially given 100 - 150 mg PTU (if used) every 8 hours or 10 - 15 mg methimazole (Tapazole) every 12 hours. The initial dosage is varied depending on the severity of the disease, size of the gland, and medical urgency. Antithyroid drugs must usually be given frequently and taken with regularity since the half-time in blood is brief -- 1.65 hours or less for PTU [116]. Frequent dosage is especially needed when instituting therapy in a severely ill patient. Methimazole has the advantage of a longer therapeutic half-life, and appears to produce fewer reactions when given in low dosage. Propylthiouracil is preferred in patients with very severe hyperthyroidism since it inhibits T4>T3 conversion, and in early pregnancy[117, 118]

In most thyrotoxic patients, the euthyroid state, as assessed by clinical parameters, and FT4, can be reached within 4 - 6 weeks. If the patient fails to respond, the dosage may be increased. Iodine-131 studies may be performed to determine whether a sufficiently large dose of medication is being employed [119], but these studies are rarely needed. In general, it is assumed that iodide uptake should be nearly completely blocked, but the 24-hour 131-I thyroid uptake in the patient under therapy may range from O% to 40%. This iodide is partly unbound and is usually released rapidly from the gland by administration of 1 g potassium thiocyanate or 400 mg potassium 131-I perchlorate. If perchlorate or thiocyanate does not discharge the iodide, it is obvious that iodide organification is occurring despite the thiocarbamide therapy. The quantity of drug administered may then be increased. In experimental animals, the thiocarbamides block synthesis of iodothyronines more readily than they block formation of MIT and DIT. This observation suggests that a complete block in organification of iodide may not be necessary to produce euthyroidism. The patient’s thyroid might accumulate and organify iodide and form iodotyrosines, but be unable to synthesize the iodothyronines. Clinical observations to prove this point are not available.

An RIA for PTU has been developed but has not proven useful in monitoring therapy [120]. Doses of 300 mg PTU produced serum levels of about 7.1 µg/ml, and serum levels of PTU correlated directly with decreases in serum T 3 levels.

It is theoretically possible to give therapeutic doses of methimazole by rectal administration in a saline enema or by suppository if the oral route is unavailable [121]. Propylthiouracil has also been administered in suppositories or in enemas and found to be effective in treating hyperthyroidism. In a recent study PTU tablets were mixed in mineral oil, and then with cocoa butter, and frozen, to produce 1 gm suppositories each containing 400mg PTU. Suppositories given 4 times daily maintained a therapeutic blood level(121.1). Jongjaroenprasert et al compared the effectiveness of a 400 mg dose of PTU in 90 ml of water vs. 400 mg of PTU given in polyethylene glycol suppositories. Both methods were effective treatments, but the enema appeared to provide greater bioavailability (121.2).

Long Term Therapeutic Program After the initial period of high-dose therapy, the amount of drug administered daily is gradually reduced to a level that maintains the patient in a euthyroid condition, as assessed by clinical evaluation and serial observations of serum T4 , FT4, or T3 . These tests should appropriately reflect the metabolic status of the patient. Measurement of TSH level is useful when the FT4 falls, to make sure that the patient has not been overtreated, but, as noted previously, TSH may remain suppressed for many weeks after thyrotoxicosis is alleviated. Serum T3 levels can also be monitored and are occasionally still elevated when the T4 level is in the normal range. During the course of treatment, the thyroid gland usually remains the same in size or becomes smaller. If the gland enlarges, the patient has probably become hypothyroid with TSH elevation; this condition should be ascertained by careful clinical and laboratory evaluation. If the patient does become hypothyroid, the dose of antithyroid drug should be reduced. Decrease in size of the thyroid under therapy is a favorable prognostic sign, and more often than not means that the patient will remain euthyroid after the antithyroid drugs have been discontinued. The dose is gradually reduced as the patient reaches euthyroidism, and often one-half or one-third of the initial dose is sufficient to maintain control. The interval between doses -- typically 8-12 hours initially -- can be extended, and patients can often be maintained on twice- or once-a-day therapy with methimazole [122]. Alternatively, antithyroid drugs can be maintained at a higher dose, and thyroxine can be added to produce euthyroidism. Occasionally ingestion of large amounts of iodide interferes with antithyroid drug therapy.

Duration of Treatment- The appropriate duration of antithyroid drug therapy is uncertain, but usually it is maintained for one year. Treatment for six months has been effective in some clinics but is not general practice [123]. Longer treatment -- such as one to three years -- does gradually increase the percentage of responders [124], but this increase must be balanced against the added inconvenience to the patient [125, 126]. Azizi and coworkers have reported treatment of a group of 26 patients for ten years, during which time no serious problems occurred, and the cost approximated that of RAI therapy(126.1). At least one study suggests that treatment with large doses of antithyroid drugs may increase the remission rate, perhaps because of an immunosuppressive action [125]. Body mass, muscle mass, and bone mineral content gradually recover, although bone mass remains below normal [126.2]. Risedronate treatment has been demonstrated to help restore bone mass in osteopenia/osteoporosis associated with Graves’ disease (126.3).

After the patient has taken the antithyroid drugs for a year, the medication is gradually withdrawn over one to two months, and the patient is observed at intervals thereafter. Elevated TRAbs at the time ATDs are to be withdrawn strongly (but imperfectly) suggest relapse will occur (110.1). Most of those who will ultimately have an exacerbation of the disease do so within three to six months; others may not develop recurrent hyperthyroidism for several years [127]. Some patients may have a recurrence after discontinuing the drug that lasts for a short time, and then a remission without further therapy [128]. Addition  of iodide therapy is also a useful possibility, as noted below. A report that administration of iodide increases the relapse rate after drug therapy is withdrawn has not been confirmed [129].

Hashizume and co-workers reported that administration of T4 to suppress TSH for a year after stopping antithyroid drugs produced a very high remission rate [130]. Similar results were found when T4 treatment was given after a course of antithyroid drugs during pregnancy. [131]. These studies engendered much interest because of the uniquely high remission rate obtained by the continuation of thyroxine treatment to suppress TSH for a year or more after the usual course of antithyroid drug therapy. Possibly such treatment is beneficial since it inhibits the release of thyroid antigens. However subsequent studies have not found a beneficial effect of added T4 therapy [131.1,131.2]. It appears that the results are, for some reason, peculiar to this study group.

The probability of prolonged remission correlates with reduction in gland size, disappearance of thyroid-stimulating antibodies from serum[132, 133], (Fig.11-8) return of T3 suppressibility, decrease in serum TG, and a haplotype other than HLA-DR3 [130 -136]. However, none of these markers predict recovery or continued disease with an accuracy rate above 60-70% [136.1]. Long after apparent clinical remission, many patients show continued abnormal thyroid function, including partial failure of T3 suppression, or absent or excessive TRH responses [127-140]. These findings probably indicate the tenuous balance controlling immune responses in these patients.

Breast feeding- Lactating women taking PTU have PTU levels of up to 7.7 µg/ml in blood, but in milk the level is much lower, about 0.7 µg/ml [141]. Only 1-2 mg PTU could be transferred to the baby daily through nursing; this amount is inconsequential except for the possibility of reactions to the drug. Azizi et al. studied intellectual development of children whose mothers took methimazole during lactation, and found that there was no evident effect on physical and intellectual development, at least in children whose mothers took up to 20 mg of MMI daily [141a].

Hypothyroidism- It has long been known that some patients with Graves’ disease eventually develop spontaneous hypothyroidism [68]. Reports have shown that most patients who become euthyroid after antithyroid drug therapy, if followed long enough, also develop evidence of diminished thyroid function [69]. In a prospective study, Lamberg et al [139]found that the annual incidence in these patients of subclinical hypothyroidism was 2.5%, and of overt hypothyroidism 0.6%.

 

TOXIC REACTIONS TO ANTITHYROID DRUGS


The use of antithyroid drugs may be accompanied by toxic reactions, depending on the drug and dose, in 3 - 12% of patients[ 117, 118, 142- 146]. Most of these reactions probably represent drug allergies[ 147- 148]. Chevalley et al., in a study of 180 patients given methimazole[ 143], found an incidence of toxicity of 4.3%, broken down as follows: Total reactions 4.3%; Pruritus 2.2%; Granulocytopenia 1.6%; Urticaria 0.5%.Methimazole may be the drug least likely to cause a toxic reaction, but there is little difference between it and PTU. When the antithyroid drugs are prescribed, the patient should be apprised of the possibility of reactions, and should be told to report phenomena such as a sore throat, fever, or rash to the physician and to discontinue the drug until the cause of the symptoms has been evaluated. These symptoms may herald a serious reaction.

Allergic rash-If a patient taking a thiocarbamide develops a mild rash, it is permissible to provide an antihistamine and continue using the drug to see whether the reaction subsides spontaneously, as it commonly does. If the reaction is more severe or if neutropenia occurs, another drug should be tried or the medication withdrawn altogether. Usually a switch is made to another thiocarbamide, because cross-reactions do not necessarily occur between members of this drug family. Alternatively, the program of therapy may be changed to the use of RAI, which may be given after the patient has stopped taking the antithyroid drug for 48 hours, or the patient may be prepared for surgery by the administration of iodides and propranolol.

The incidence of agranulocytosis in a large series of patients was 0.4% [149]. It occurs most frequently in older patients and those given large amounts of the drug (20-30 mg methimazole every eight hours) [117].Reactions tend to be most frequent in the first few months of therapy but can occur at any time, with small doses of drug, and in patients of all ages [117]. The most common reactions are fever and a morbilliform or erythematous rash with pruritus. Reactions similar to those of serum sickness, with migratory arthralgias, jaundice, lymphadenopathy, polyserositis, and episodes resembling systemic lupus erythematosus have also been observed [147]. Pyoderma gangrenosum can occur (147.1). Neutropenia and agranulocytosis are the most serious complications. These reactions appear to be due to sensitization to the drugs, as determined by lymphocyte reactivity in vitro to the drugs [148]. Occasionally agranulocytosis can develop even though the total WBC remains within normal ranges- a hazard to be remembered and differential counts should be  done. Fortunately, even these problems almost always subside when the drug is withdrawn. Aplastic anemia with marrow hypoplasia has been reported (perhaps 10 cases), again with spontaneous recovery in 2-5 weeks in 70%, but fatal outcome in 3 patients [149]. Thrombocytopenia and/or anemia may accompany the neutropenia. Vasculitis is a fortunately rare complication during treatment with antithyroid drugs.
Neutropenia-It is probably wise to see patients receiving the thiocarbamides at least monthly during the initiation of therapy and every two to three months during the entire program. Neutropenia can develop gradually but often comes on so suddenly that a routine white cell count offers only partial protection. A white cell count must be taken whenever there is any suggestion of a reaction, and especially if the patient reports malaise or a sore throat. A white cell count taken at each visit will detect the gradually developing neutropenia that may occur. While many physicians do not routinely monitor these levels, the value of monitoring is suggested by the study of Tajiri et al [144]. Fifty-five of 15398 patients treated with antithyroid drugs developed agranulocytosis, and 4/5 of these were detected by routine WBC at office visits. Low total leukocyte counts are common in Graves’ disease because of relative neutropenia, and for this reason a baseline WBC and differential should be performed before starting anti-thyroid drugs. However, total polymorphonuclear counts below 2,000 cells/mm3 should be carefully monitored; below 1,200 cells/mm3 it is unsafe to continue using the drugs.

In the event of severe neutropenia or agranulocytosis, the patient should be monitored closely, given antibiotics if infection develops, and possibly adrenal steroids. There is no consensus on the use of glucocorticoids, since they have not been shown to definitely shorten the period to recovery. Administration of recombinant human granulocyte colony stimulating factor (75 µg/day given IM) appears to hasten neutrophile recovery in most patients who start with neutrophile counts > 0.1 X 109/L [150-151]. Antithymocyte globulin and cyclosporin have also been used [151]. Care must be taken to ensure against exposure to infectious agents, and some physicians prefer not to hospitalize their patients for this reason. If the patient is hospitalized, he or she should be placed in a special-care room with full bacteriologic precautions.

ANCA antibodies- Patients may develop antineutrophil cytoplasmic antibodies, either pericytoplasmic or cytoplasmic, during treatment, with or without vasculitis. Most cases appear to be associated with the use of propylthiouracil, and therapy includes cessation of the drug, sometimes treatment with steroids or cyclophosphamide for renal involvement, and rarely plasmapheresis. The commonest cutaneous lesion associated is leukocytoclastic vasculitis associated with purpuric lesions. Symptoms may include fever, myalgia, arthralgia, and lesions in the kidneys and lungs. Prognosis is usually good if the medication is discontinued, although death has occurred. ANCA positivity (pericytoplasmic, cytoplasmic, directed to myeloperoxidase, proteinase3, or human leukocyte elastase) can occur in patients on antithyroid drugs associated with vasculitis. It is also found without clinical evidence of vasculitis, and the significance of this finding is unclear [151.1]. Guma et al recently reported that, in a series of patients with Graves’ disease, 67% were found to be ANCA positive before medical treatment, and that 19% remained positive after one year of antithyroid treatment. This data suggests that ANCA antibodies reflect in some way the autoimmunity associated with Graves’ hyperthyroidism, rather than simply being a manifestation due to the treatment with antithyroid drugs (151.2). In addition to suppression of hematopoiesis and agranulocytosis, methimazole has been associated in one patient with massive plasmocytosis, in which 98% of the cells in the bone marrow were plasma cells. After discontinuation of the drug, and treatment with dexamethasone and G-CSF, the patient’s marrow recovered to normal (151.3).

Liver damage-Thiocarbamides can also cause liver damage ranging from elevation of enzymes, through jaundice, to fatal hepatic necrosis. Toxic hepatitis (primarily with propylthiouracil) and cholestatic jaundice (primarily with methimazole) are fortunately uncommon [150].Toxic hepatitis can be severe or fatal, but the incidence of serious liver complications is so low that routine monitoring of function tests has not been advised[ 1514, 152]. Liver transplantation has been used with success in several patients [152.1]. As noted above, any sign of liver damage must be carefully monitored, and progress of abnormalities in liver function tests demand cessation of the drug[147, 152].

Diffuse interstitial pneumonitis has also been produced by propylthiouracil [153].

Pregnancy-(Please also see chapter on Thyroid Regulation and Dysfunction in the Pregnant Patient). Methimazole should be avoided in early pregnancy as disc ussed above. Very rare cases of esophageal atresia, omphalocele, and choanal atresia occurred in Sweden almost only in infants whose mothers took methimazole during early pregnancy.This is thought to be a true, although fortunately very infrequent, complication of methimazole use. Their observations obviously suggest that methimazole should best not be given during early months of pregnancy (153.1). As noted elsewhere in this web-book, various options are available, including 1) arranging definitive treatment before pregnancy, 2) switching to propylthiouracil as soon as possible and use of that drug during the first trimester, and leaving mild hyperthyroidism untreated (wich associated risks).  iodide treatment can be tried instead of ATD, and is reported to be  significantly more safe, although experience with this approach is inadequate for recommendation (154)

 

Potassium Perchlorate, Lithium,and Cholestyramine

Potassium perchlorate was introduced into clinical use after it was demonstrated that several monovalent anions, including nitrates, have an antithyroid action. Perchlorate was the only member of the group that appeared to have sufficient potency to be useful. This drug, in doses of 200 - 400 mg every six hours, competitively blocks iodide transport by the thyroid. Accordingly, therapeutic doses of potassium iodide will overcome its effect. Institution and control of therapy with perchlorate are similar to those discussed for the thiocarbamides. Toxic reactions to this agent occur in about 4% of cases [155] and usually consist of gastric distress, skin rash, fever, lymphadenopathy, or neutropenia; they usually disappear when the drug is discontinued. The reaction rate is higher when doses of more than 1 g/day are given [155]. Nonfatal cases of neutropenia or agranulocytosis have been reported, and four cases of fatal aplastic anemia have been associated with the use of this drug [156]. Because of toxic reactions, perchlorate is not used at present for routine therapy. It has found a role in therapy of thyrotoxicosis induced by amiodarone [157]. Apparently blocking of iodide uptake is an effective antithyroid therapy in the presence of large body stores of iodide, while in this situation, methimazole and propylthiouracil are not effective alone.

Lithium ion inhibits release of T4 and T3 from the thyroid and has been used in the treatment of thyrotoxicosis, but is most effective when used with a thiocarbamide drug. It does not have a well-established place in the treatment of Graves’ disease[ 157, 158]. It has possible value in augmenting the retention of 131-I [159] and in preparing patients allergic to the usual antithyroid drugs or iodide for surgery, although propranolol is generally used for the latter problem.

Cholestyramine (4gm, q8h) for a month has been shown to hasten return of T4 to normal [159.1] by binding hormone in the gut. It can be used as an adjunct to help speed return of hormone levels to normal, and may be especially beneficial in thyroid storm.

Iodine treatment- Plummer originally observed that the administration of iodide to thyrotoxic patients resulted in an amelioration of their symptoms. This reaction is associated with a decreased rate of release of thyroid hormone from the gland and with a gradual increase in the quantity of stored hormone. The effect of iodide on thyroid hormone release and concentration in blood is apparent in Figure 7. The mechanism of action may be by inhibition of generation of cAMP, and involves inhibition of TG proteolysis, but is not fully understood. Therapeutic quantities of iodide also have an effect on hormone synthesis through inhibition of organification of iodide. Iodide has similar but less intense effects on the normal thyroid gland, apparently because of adaptive mechanisms.

Administration of large amounts of iodide to laboratory animals or humans blocks the synthesis of thyroid hormone and results in an accumulation of trapped inorganic iodide in the thyroid gland (the Wolff-Chaikoff effect, see Ch 2). The thyrotoxic gland is especially sensitive to this action of iodide. Raising the plasma iodide concentration to a level above 5 µg/dl results in a complete temporary inhibition of iodide organification by the thyrotoxic gland. In normal persons elevation of the inorganic 127-I level results, up to a point, in a progressive increase of accumulation of iodide in the gland. When the plasma concentration is above 20 µg/dl, organification is also inhibited in the normal gland [160]. The sensitivity of the thyrotoxic gland, in comparison with that of the euthyroid gland, may be due to an increased ability to concentrate iodide in the thyroid, and its failure to "adapt" by decreasing the iodide concentrating mechanism.

When iodine is to be used therapeutically in Graves’ disease, one usually prescribes a saturated solution of potassium iodide (which contains about 50 mg iodide per drop) or Lugol’s solution (which contains about 8.3 mg iodide per drop). Thompson and co- workers [161] found that 6 mg of I- or KI produces a maximum response. This fact was reemphasized by Friend, who pointed out that the habit of prescribing the 5 drops of Lugol’s or SSKI three times daily is unnecessary [162]. Two drops of Lugol’s solution or 1 drop of a saturated solution of potassium iodide two times daily is more than sufficient.

The therapeutic response to iodide begins within two to seven days and is faster than can be obtained by any other methods of medical treatment. Only 3% of patients so treated fail to respond. Men, older persons, and those with nodular goiter are in the group less likely to have a response to iodide. Although almost all patients initially respond to iodide, about one-third respond partially and remain toxic, and another one-third initially respond but relapse after about six weeks [163].

Because of the partial responses and relapse rate, use of iodide as definitive therapy for thyrotoxicosis has been replaced by the modalities described  above. Currently Iodides are given sometimes after 131-I therapy to control hyperthyroidism, and are usually given as part of treatment before thyroidectomy. However some recent reports suggest iodide might have a larger role to play. Addition  of iodine (38 mg/day) to methimazole (15mg/d) accelerated response over methimazole alone (154), and long term iodine treatment induced remission in 38% of patients who were given this treatment because of adverse reactions to ATD (164). In a study of 30 drug-naïve patients with “mild” GD, all but 3 were controlled on iodine alone (165.). Use of iodides instead of methimazole during the first trimester of pregnancy reduced major anomalies from 4.1% to 1.5% in one study (165.1). Iodine treatment is not currentty considered standard, but this may change soon.

 

Adjunctive Therapy for Graves’ Disease
Propranolol, metopranol, atenolol

Beta-adrenergic blocking agents have won a prominent position in the treatment of thyrotoxicosis. Although they alleviate many of the signs and symptoms, they have little effect on the fundamental disease process[ 166, 167]. Palpitations, excessive sweating, and nervousness improve, and tremor and tachycardia are controlled. Many patients feel much improved, but others are psychologically depressed by the drug and prefer not to take it. Improvement in myocardial efficiency and reduction in the exaggerated myocardial oxygen consumption have been demonstrated [168]. Propranolol lowers oxygen consumption [169, 170] and reverses the nitrogen wasting of thyrotoxicosis, although it does not inhibit excess urinary calcium and hydroxyproline loss. Propranolol is useful in symptomatic treatment while physician and patient are awaiting the improvement from antithyroid drug or 131-I therapy [171]. Some patients appear to enter remission after using this drug alone for six months or so of therapy[ 169, 172]. It has been useful in neonatal thyrotoxicosis [173] and in thyroid storm [174]. The drug must be used cautiously when there is evidence of severe thyrotoxicosis, or heart failure, but often control of tachycardia permits improved circulation. Beta blockade can induce cardiovascular collapse in patients with or without heart failure, and asystolic arrest (174.1,174.2). Administration of beta blocker was shown by Ikram to reduce CO by 13% in patients with uncontrolled CHF, and apparently this reduction in CO can be near fatal in rare patients.
Some surgical groups routinely prepare patients for thyroidectomy with propranolol for 20 - 40 days and add potassium iodide during the last week [175]. The BMR and thyroid hormone level remain elevated at the time of operation, but the patient experiences no problems. We prefer conventional preoperative preparation with thiocarbamides, with or without iodide, and would use propranolol as an adjunct, or if the patient is allergic to the usual drugs.

Propranolol is usually given orally as 20 - 40 mg every four to six hours, but up to 200 mg every six hours may be needed. In emergency management of thyroid storm (see also Chapter 12) or tachycardia, it may be given intravenously (1 - 3 mg, rarely up to 6 mg) over 3 - 10 minutes and repeated every four to six hours under electrocardiographic control. Atropine (0.5 - 1.0 mg) is the appropriate antidote if severe brachycardia is seen.

Reserpine and Guanethidine Drugs such as reserpine [177] and guanethidine [178] that deplete tissue catecholamines were used extensively in the past as adjuncts in the therapy for thyrotoxicosis, but fell into disuse as the value of beta -sympathetic blockade with propranolol became recognized.

Glucocorticoids, Ipodate, and Other Treatments As described elsewhere, potassium iodide acts promptly to inhibit thyroid hormone secretion from the Graves’ disease thyroid gland. PTU, propranolol, glucocorticoids [181], amiodarone, and sodium ipodate (Oragrafin Sodium) inhibit peripheral T4 to T3 conversion, and glucocorticoids may have a more prolonged suppressive effect on thyrotoxicosis [182]. Orally administered resins bind T4 in the intestine and prevent recirculation [183]. All of these agents have been used for control of thyrotoxicosis [ 184, 185]. Combined dexamethasone, potassium iodide, and PTU can lower the serum T3 level to normal in 24 hours, which is useful in severe thyrotoxicosis. Prednisone has been reported to induce remission of Graves’ disease, but at the expense of causing Cushing’s syndrome [187]. Ipodate (0.5 - 1 g orally per day) acts to inhibit hormone release because of its iodine content, in addition to its action to inhibit T4 to T3 conversion. This dose of ipodate given to patients with Graves’ disease reduced the serum T3 level by 58% and the T4 level by 20% within 24 hours, and the effect persisted for three weeks[188, 189]. This dose of ipodate was more effective than 600 mg of PTU, which decreased the T3 level by only 23% during the first 24 hours, whereas the T4 level did not drop. Ipodate may prove to be a useful adjunct in the early therapy of hyperthyroidism, but will increase total body and thyroidal iodine. However, when the drug is stopped, the RAIU in Graves’ patients usually returns to pre- treatment levels within a week [189]. Because it is the most effective agent available in preventing conversion of T4 to T3, it has a useful role in managing thyroid storm.

Immunosuppressive Therapy- Development of new targeted and relatively safe immune suppressive treatments has allowed their extension to Graves’ disease. Rituximab, an anti CD20  B cell lymphocyte depleting monoclonal antibody, was initially found to induce remission in Graves’ ophthalmopathy. It also mediates decreases in anti thyroid antibodies, and is currently employed in a Phase II trial for therapy of mild, relapsing Graves’ disease (189.1, 189.2). Significant adverse events during therapy with rituximab (“serum sickness”, mild colitis, iridocyclitis, polyarthritis) have been reported, and will probably limit its usefulness (189.3) Use of agents of this type, that work by increasing function of regulatory T cells, will probably become common in the next few years. Another approach has been pioneered by Gershengorn and colleagues, who devised a small molecule that is an “allosteric inverse agonist” of TSHR, and inhibits stimulation of TSH receptor activation by TSAbs (189.4 ). Such agents are used in current clinical trials, and should offer entirely new treatment stategies in the future.

 

SURGICAL THERAPY

Subtotal thyroidectomy is an established and effective form of therapy for Graves’ disease, providing the patient has been suitably prepared for surgery. In competent hands, the risk of hypoparathyroidism or recurrent nerve damage is under 1%, and the discomfort and transient disability attendant upon surgery may be a reasonable price to pay for the rapid relief from this unpleasant disease. In some clinics it is the therapy of choice for most young male adults, especially if a trial of antithyroid drugs has failed. Total thyroidectomy may be preferred in patients with serious eye disease or high TRAb levels, in order to help the eye disease and to keep down the incidence of recurrence [190-194].
As with other effective methods available, it is necessary for the physician and the patient to decide on the form of therapy most suitable for the case at hand. Because of the potential but unproved risks of 131-I therapy, it is not always possible to make an entirely rational choice; the fears and prejudices of the physician and the patient will often enter into the decision. Surgery is clearly indicated in certain patients. Among these are (1) patients who have not responded to prolonged antithyroid drug therapy, or who develop toxic reactions to the drug and for whatever reason are unsuitable for 131-I therapy; (2) patients with huge glands, which frequently do not regress adequately after 131-I therapy; and (3) patients with thyroid nodules that raise a suspicion of carcinoma. Stocker et al have reviewed the problem of nodules in Graves’ glands (195). They found that 12% of Graves’ patients had cold defects on scan, and among these half were referred for surgery. Six of 22, representing 2% of all Graves’ patients, 15% of patients with cold nodules, 25% of patients with palpable nodules, and 27% of those going to surgery, had papillary cancer in the location corresponding to the cold defect. Of these patients, one had metastasis to bone and two required multiple treatments with radioiodine. These authors argue for evaluating patients with a thyroid scintigram and further diagnostic evaluation of cold defects. Subtotal or near total thyroidectomy is often the treatment of choice for patients with amiodarone induced thyrotoxicosis, since response to ATDs is typically poor, and RAIU can not be given (196). Surgery may also have a place in therapy of older patients with thyroid storm and/or cardio-respiratory failure, who do not respond rapidly to intensive medical therapy(197).

 

Surgery in patients with ophthalmopathy


Contemporary data indicate that exophthalmos may be exacerbated by RAI therapy [80],although in some studies appearance of progressive ophthalmopathy was about the same after treatment with 131-I as with surgery [79]. Thus, in the presence of serious eye signs, treatment with antithyroid drugs followed by surgery is an important alternative to consider, and total thyroidectomy is preferred [ 80-82]. The preferential use of surgery rather than radioactive iodide in the management of patients with severe Graves’ ophthalmopathy, and the greater, more frequent exacerbation of eye disease after RAI therapy, has been supported in a number of studies including those by Torring et al [36.2], Moleti et al [44.4], and De Bellis et al [44.5] and others documented above. Marcocci et al, in contrast, report that near-total thyroidectomy had no efffect on the course of ophthalmopathy in a group of patients who had absent or non-severe preexisting ophthalmopathy. The relevance of this to patients with more severe ocular disease is uncertain, since it is logical to expect that in these patients there would be no effect of removing antigens, if the patients  lacked any tendency to develop ophthalmopathy [44.6]. Moleti et al recently reported on 55 patients with Graves’ disease and mild to moderate Graves’ ophthalmopathy, who underwent near-total thyroidectomy, and of whom 16 had standard ablative doses of radioactive iodide. They found that the course of ophthalmopathy, both short and long term after treatment, was significantly better in the group of patients who underwent thyroidectomy and 131-I ablation, and suggest that this is a more effective means of inducing and maintaining ophthalmopathic inactivity (44.7). In a randomized, prospective study, total thyroidectomy, rather than partial thyroidectomy, was followed by a better outcome of GO in patients given iv glucocorticoids. Radioiodine uptake test and thyroglobulin assay showed complete ablation in the majority of total, but not of partial thyroidectomy patients(44.6) .

The rate of patient rehabilitation is probably quickest with surgery. Although the source of hormone is directly and immediately removed by surgery, the patient usually must undergo one to three months of preparation before operation. The total time from diagnosis through operative convalescence is thus three to four months. Antithyroid drugs, in contrast, provide at best only 30 - 40% permanent control after one year of therapy. Iodine-131 can assuredly induce prompt remission, but low dose protocols, as noted, are plagued by a need for medical management and retreatment over one to three years before all patients are euthyroid. Treatment with higher doses provides more certain remission at the expense of more certain hypothyroidism.

There are several strong contraindications to surgery, including previous thyroid surgery, severe coincident heart or lung disease, the lack of a well-qualified surgeon, and pregnancy in the third trimester, since anesthesia and surgery may induce premature labor.

More enthusiastic surgeons have in the past recommended surgery for all children as the initial approach, claiming that there is less interference with normal growth and development than with prolonged antithyroid drug treatment [191]. Therapy for childhood thyrotoxicosis is discussed further below.

Preparation for Surgery

Antithyroid drugs of the thiocarbamide group are employed to induce a euthyroid state before subtotal thyroidectomy when surgery is the desired form of treatment. Two approaches are used. Mmethimazole (or PTU if used) may be administered until the patient becomes euthyroid. After this state has been reached, and while the patient is maintained on full doses of thiocarbamides, Lugol’s solution or a saturated solution of potassium iodide is administered for 7 - 10 days. This therapy induces an involution of the gland and decreases its vascularity, a factor surgeons find helpful in the subsequent thyroidectomy. In one study Lugol solution treatment resulted in a 9.3-fold decreased rate of intraoperative blood loss. Preoperative Lugol solution treatment decreased the rate of blood flow, thyroid vascularity measured by histomorphometry , and intraoperative blood loss during thyroidectomy(198).

The iodide should be given only while the patient is under the effect of full doses of the antithyroid drug; otherwise, the iodide may permit an exacerbation of the thyrotoxicosis. Alternatively patients may be prepared by combined treatment with antithyroid drugs and thyroxine. It is not obvious that one method is superior to the other. Severely ill patients can be prepared for surgery rapidly by combining several treatments-iopanoic acid 500mg bid, dexamethasone 1mg bid, antithyroid drugs, and beta-blockers(199).

Pre-treatment should have the patient in optimal condition for surgical thyroidectomy. By this time the patient has gained weight, the nutritional status has been improved, and the cardiovascular manifestations of the disease are under control. At the time of surgery, the anesthesia is well tolerated without the risk of hypersensitivity to sympathoadrenal discharge characteristic of the thyrotoxic subject. The surgeon finds that the gland is relatively avascular. Convalescence is customarily smooth. The stormy febrile course characteristic of the poorly prepared patient in past years is rarely seen.

Reactions to the thiocarbamide drugs occasionally occur during preparation for surgery. If the problem is a minor rash or low-grade fever, the drug is continued, or a change is made to a different thiocarbamide. More severe reactions (severe fever or rash, leukopenia, jaundice, or serum sickness) necessitate a change to another form of therapy, but no entirely satisfactory alternative is available. One course is to administer iodide and propranolol and proceed to surgery. In some patients, it is best to proceed directly to 131-I therapy if difficulties arise in the preparation with antithyroid drugs.

Propranolol has been used alone or in combination with potassium iodide [199] in preparation for surgery, and favorable results have generally been reported[200-201]. This procedure is doubtless safe in the hands of a medical team familiar and experienced with this protocol and willing to monitor the patient carefully to ensure adequate dosage. It is safe to use in young patients with mild disease, but is not advised as a standard protocol. Propranolol is used as an adjunct, or combined with potassium iodide as the sole therapy only when complications with antithyroid drugs preclude their use and surgery is strongly preferred to treatment with 131-I.

Amiodarone induced hyperthyroidism is typically difficult to manage, as described in Chapter 13. Administration of iopanoic acid, 1 gm daily for 13 days, has been shown to provide successful pre-operative therapy, reducing T3 levels to normal (196). Propranolol is the usual drug used for preparation of patients with amiodarone induced hyperthyroidism going to surgery.

Surgical Techniques and Complications

The standard operation is a one-stage subtotal thyroidectomy. General anesthesia is standard, but cervical plexus block and out-patient surgery is employed by some surgeons [202]. The amount of tissue left behind is about 4-10 grams, but this amount is variable. Taylor and Painter [203] found that the average volume of this remnant in 43 patients achieving a remission was about 8 ml, and Sugino et al recommended leaving 6 grams of tissue [204]. The toxic state recurred in only two patients in their series, and in these twice the amount of tissue mentioned above was left. Ozaki also noted the importance of the amount of thyroid remaining as the principal predictor of eu- or hypo-thyroidism [205].There seems however to be no relation between the original size of the thyroid and the size of the remnant necessary to maintain normal metabolism.

Motivated in part by economic considerations, there has been in recent years a reevaluation of thyroidectomy done under local anesthesia as a day-surgery proceedure. Pros and cons have recently been discussed. In proper hands local anesthesia and prompt discharge seem acceptable, but most surgeons opt for the standard in hospital approach since it offers a more controlled operative setting and an element of safety the night after surgery. Some clinicians argue for total-thyroidectomy in an effort to reduce recurrence rates (206, 207), and point out that this operation seems to reduce anti-thyroid autoimmunity and reduces the chance of exacerbation of ophthalmopathy. Permanent cure of the hyperthyroidism is produced in 90 - 98% of patients treated this way.

 

Complications of Surgery

Although surgery of the thyroid has reached a high degree of perfection, it is not without problems even in excellent hands. The complication rates at present are low [208]. Among 254 patients operated on at three Nashville hospitals in the decade before 1970, there was no mortality, only minor wound problems, a 1.9% incidence of permanent hypoparathyroidism, and a 4.2% recurrence rate [209]. Hypo-parathyroidism is the major undesirable chronic complication. Surgical therapy at the Mayo Clinic has [210] been associated with a 75% rate of hypothyroidism but only a 1% recurrence rate, as an effort was made to remove more tissue and prevent recurrences. There is typically an inverse relationship between these two results of surgery. In the recent experience of the University of Chicago Clinics, the euthyroid state has been achieved by surgery in 82%; 6% became hypothyroid, and the recurrence rate was 12% [200]. Palit et al. published a meta analysis of collected series of patients treated for Graves’ disease, either by total thyroidectomy or subtotal thyroidectomy. Overall, the surgery controlled hyperthyroidism in 92% of patients. There was no difference in complication rates between the two kinds of operations, with permanent laryngeal nerve injury occurring in 0.7 - 0.9% of patients, and permanent hypoparathyroidism in 1 – 1.6% of patients. Since many surgeons have become more familiar with and capable of total thyroidectomy, and this avoids the possible recurrence of disease, although possibly slightly increasing the risk of nerve or parathyroid damage, total thyroidectomy has become a common or even preferred alternative to subtotal thyroidectomy for managing hyperthyroidism. Recurrence rates are higher in patients with progressive exophthalmos or strongly positive assays for TRAb, suggesting that total thyroidectomy may be preferred in these cases [207]. Geographic differences in iodine ingestion have been related to the outcome.

Death rates are now approaching the vanishing point [206-210] Of the nonfatal complications, permanent hypoparathyroidism is the most serious, and requires lifelong medical supervision and treatment. Experienced surgeons have an incidence under 1%. Unfortunately, the general experience is near 3%. More patients, perhaps 10%, develop transient post-operative hypocalcemia but soon recover apparently normal function. Perhaps these patients have borderline function that may fail in later years.

Unilateral vocal cord paralysis rarely causes more than some hoarseness and a weakened voice, but bilateral injury leads to permanent voice damage even after corrective surgery. Bilateral recurrent nerve injury may be associated with severe respiratory impairment when an acute inflammatory process supervenes and may be life-threatening. Fortunately, it is now extremely rare after subtotal thyroidectomy. Damage to the superior external laryngeal nerve during surgery may alter the quality of the voice and the ability to shout without causing hoarseness. One may speculate whether declining skills in the techniques of subtotal thyroidectomy, attendant upon a dramatic fall in the use of this procedure, may lead to an increase in the hazards of the procedure.

Hypothyroidism, whether occurring after surgery or 131-I therapy, can be readily controlled. Transient hypothyroidism is common, with recovery in one to six months. The presence of autoimmunity to thyroid antigens predisposes to the development of hypothyroidism after subtotal thyroidectomy for thyrotoxicosis. A positive test for antibodies to the microsomal/TPO antigen was found years ago by Buchanan et al [211] to correlate with an increased incidence of postoperative hypothyroidism. The incidence of hypothyroidism is certainly of importance in weighing the virtues of 131-I and surgical therapy. The ability of surgical therapy to produce a euthyroid state in many patients over long-term follow-up gives it one advantage over RAI therapy, but this must be weighed against the risk of hypoparathyroidism and recurrent nerve damage.

Course After Surgery --

In the immediate postoperative period, patients should be followed closely. They should ideally have a special duty nurse or family member providing watch during the first 24 hours, and a tracheotomy set and calcium chloride or gluconate for infusion should be at the bedside. During this period, undetected hemorrhage can lead to asphyxiation. Current use of drains with constant suction helps protect against this problem.
Transient hypocalcemia is common, resulting from trauma to the parathyroid glands and their blood supply and also possibly to rapid uptake of calcium by the bones, which have been depleted of calcium by the thyrotoxicosis [212,213]. Oral or intra- venous calcium supplementation suffices in most instances to control the symptoms. The calcium may be given slowly intravenously as calcium gluconate or calcium chloride in a dose ranging from 0.5 to 1.0 g every 4-8 hours, as indicated by clinical observation and determination of Ca2+.

 

Replacement thyroid hormone-


If sub-total throidectomy has been performed, thyroid hormeone replacement may not be needed. In 50-70% of patients, the residual gland is able to form enough hormone to prevent even transient clinical hypothyroidism. Serum hormone levels should be determined every two to four months until it is clear that the patient does not need replacement. Some surgeons give their patients thyroxine for an indefinite period after the operation in an attempt to avoid transient hypothyroidism and to remove any stimulus to regeneration of the gland.
If total thyroidectomy has been performed, as is increasingly the case, full replacement doses of thyroxine (1.7 ug/kg BW, or about 1ug/pound of lean body mass) should be instituted immediately, and T4 levels checked in about 2 weeks for adjustment. Patients should be informed that they will need this treatment for life, and that they should re regularly checked, and consistent in their daily dosage.

 

Long Term Follow-Up

Probably the thyroid remnant is not normal. It has a rapid 131-I turnover rate and a small pool of stored organic iodine. Suppressibility by T3 administration returns within a few months of operation in some patients. TSAb tend to disappear from the blood in the ensuing 3 - 12 months [214-2156]. After subtotal thyroidectomy, thyrotoxicosis recurs in 5 - 10% of patients, often many years after the original episode. The long term outcome of thyroid surgery for hyperthyroidism was reviewed by the Department of Surgery at Karolinska Institute. Of 380 patients observed and treated by surgery for thyrotoxicosis, primarily by subtotal thyroidectomy, 1% developed permanent hypoparathyroidism. Recurrent disease occurred in 2%. The operators intended to leave less than two grams of thyroid tissue, which presumably accounts for the low recurrence rate (216).

Finally, adequate follow-up must be carried out after any kind of treatment of Graves’ disease. Recurrence is always possible, either early or late, and there is always the threat that the ophthalmopathic problems may worsen when all else in the progress of the patient seems favorable. A surprisingly large proportion of patients who have had subtotal thyroidectomy for Graves’ disease and who are clinically euthyroid can be shown to have an abnormal TRH response (excessive or depressed), and up to a third have elevated serum TSH levels [217, 218]. Some of them are undoubtedly mildly hypothyroid, whereas others are close to euthyroid but require the stimulation of TSH to maintain this state. These patients should have replacement T 4 therapy if the elevated TSH persists. Over subsequent years the residual thyroid fails in more patients, due either to reduced blood supply, fibrosis from trauma, or continuing autoimmune thyroiditis. After 10 years, and depending on the extent of the original surgery, 20 - 40% are hypothyroid. This continuing thyroid failure is also seen after antithyroid drug therapy with 131-I and represents the natural evolution of Graves’ disease.

SPECIAL CONSIDERATIONS IN THE TREATMENT OF THYROTOXICOSIS IN CHILDREN

Thyrotoxicosis may occur in any age group but is unusual in the first five years of life. The same remarkable preponderance of the disease in females over males is observed in children as in the adult population, and the signs and symptoms of the disease are similar in most respects. Behavioral symptoms frequently predominate in children and produce difficulty in school or problems in relationships within the family. Thyrotoxic children are tall for their age, probably as an effect of the disease. These children are restored to a normal height/age ratio after successful therapy for the thyrotoxicosis. Permanent brain damage and craniosynostosis are reported as complications of early childhood thyrotoxicosis ( 219). Bone age is also often advanced [220].

No more is known about the cause of the disease in children than in adults. Diagnosis rests upon eliciting a typical history and signs and upon the standard laboratory test results. Normal values for children are not the same as for adults during the first weeks of life, and these differencesshould be taken into account.

 

Therapy of Childhood Graves’ Disease

131-I Treatment- In some clinics, RAI is used in the treatment of thyrotoxicosis in children. In an early report, 73 children and adolescents were so treated. Hypothyroidism developed in 43. Subsequent growth and development were normal [221]. In another group of 23 treated with 131-I, there were 4 recurrences, at least 5 became hypothyroid, and one was found to have a papillary thyroid cancer 20 months after the second dose [222]. Safa et al. [40] reviewed 87 children treated over 24 years and found no adverse effects except the well-known occurrence of hypothyroidism. Hamburger has examined therapy in 262 children ages 3 - 18 and concluded 131-I therapy to be the best initial treatment [42]. Read et al (223) reviewed  experience with 131-I over a 36 year period, including six children under age 6, and 11 between 6 and 11 years. No adverse effects on the patients or their offsprings were found, and they advocate 131-I as a safe and effective treatment.
Nevertheless, most physicians remain concerned about the risks of carcinogenesis, and the experience of Chernobyl has accentuated this concern. This problem was more fully discussed earlier in this chapter. Others believe that the risks of surgery and problems with antithyroid drug administration outweigh the potential risk of 131-I therapy. This problem was critically reviewed by Rivkees et al [224]. They point out the significant risks of reaction to antithyroid drugs, and of surgery. Surgery may have a mortality rate in hospital in children of about one per thousand operations, although this may have decreased in recent years. Among problems with radioactive iodide therapy, they note the whole body radiation, possibly worsening of eye disease, and the apparent lack of significant thyroid cancer risk so far reported among children treated with I-131 for Graves’ disease. They assumed that risk would be lower in children after age five, and especially after age ten, and if all thyroid cells were destroyed. They advise using higher doses of radioiodine to minimize residual thyroid tissue, and avoiding treatment of children under age five, but they believe that RAI is a convenient, effective, and useful therapy in children with Graves’ disease. However, as noted above in the section on risks related to use od 131-i, Rivkees own data indicate that treatnment of children with conventional doses of RAI may induce a lifetime risk of any fatal cancer of over 2%, a very serious consideration (44.2) .Concern about the potential long term induction of cancer by RAI given to children is discussed above. Many physicians remain reluctant to use 131-I in children under age 15-18 as a first line therapy.

Surgery in children- Although 131-I therapy may gain acceptance, the most common choice for therapy is between antithyroid drugs and subtotal thyroidectomy [225-227]. Proponents of antithyroid drug therapy believe that there is a greater tendency for remission of thyrotoxicosis in children compared to adults and that antithyroid drug therapy avoids the psychic and physical problems caused by surgery in this age group. With drugs the need for surgery (or 131-I) can be delayed almost indefinitely until conditions become favorable.

As arguments against surgery, one must consider the morbidity and possible, although rare, mortality. Surgery means a permanent scar, and the recurrence rate is much higher (up to 15%) than that observed in adults. If the recurrence rate is kept acceptably low by performing near-total thyroidectomies, there is always an attendant rise in the incidence of permanent hypothyroidism, and greater potential for damage to the recurrent laryngeal nerves and parathyroid glands. Damage to the parathyroids necessitates a complicated medical program that may be permanent, and is one of the major reasons for opposing routine surgical therapy in this disease. However Rudberg et al [228] reported that, in a series of 24 children treated surgically, only one had permanent hypoparathyroidism, and two recurred within 12 years. Soreide et al [229] operated on 82 children and had no post-op nerve palsy, no tetany, nor mortality, and point out that surgery can provide a prompt, safe, and effective treatment. Childhood Graves’ disease was managed by near-total thyroidectomy in 78 patients of average age 13.8 years as reported by Sherman et al. Transient hypoparathyroidism and RCN damage were seen. Only three patients required subsequent 131-I treatment. Eighty-five % of those with ophthalmopathy were improved after surgery. The authors conclude that the treatment is safe and effective when performed by experienced surgeons (230).Others have pointed out the high relapse rate with all forms of therapy in the pediatric age group.

The main argument favoring surgery is that it may correct the thyrotoxicosis with surety and speed, and result in less disruption of normal life and development than is associated with long-term administration of antithyroid drugs and the attendant constant medical supervision. Often children are unable to maintain the careful dosage schedule needed for control of the disease.

If surgery is elected, the patient should be prepared with an antithyroid drug such as methimazole in a dosage and duration sufficient to produce a euthyroid state, and then should be given iodide for seven days before surgery. Lugol’s solution, or a saturated solution of potassium iodide, 1 or 2 drops twice daily, is sufficient to induce involution of the gland.

Anti-thyroid drug therapy in  children-  Antithyroid drug therapy is the usual preferable initial therapy in children. Favorable indications for its use are mild thyrotoxicosis, a small goiter, recent onset of disease, and especially the presence of some obvious emotional problem that seems to be related to precipitation of the disease. Antithyroid drug administration necessitates much supervision by the physician and the parents, the permanent remission rate will be 50% or less, and there is always the possibility of a reaction to the medication.

There is no consensus on secondary treatment if antithyroid drugs fail.. Some physicians favor surgery if the patient and parents seem incapable of following a regimen requiring frequent administration of medicine for a prolonged period or drug reactions occur. A factor that must be remembered in selecting the appropriate course of therapy is the experience of the available surgeon. Lack of experience contributes to a high rate of recurrence, permanent hypothyroidism, or permanent hypoparathyroidism. Other physicians believe the possible but unproven risks of 131-I are more than outweighed by the known risks of operation, and 131-I treatment is increasingly accepted for patients over age 15.

If antithyroid drugs are chosen as primary therapy, the patient is initially given a course of treatment for one or two years, according to the dosage schedule shown in Table 11-9. The dosage of PTU (if used) needed is usually 120 - 175 mg/m2 body surface area daily divided into three equal doses every eight hours. Methimazole can be used in place of PTU; approximately one-tenth as much, in milligrams, is required. Methimazole is now the preferred drug. During therapy the dosage can usually be gradually reduced. Many patients will be satisfactorily controlled by once-a-day treatment. Although the plasma half-life of methimazole in children is only 3-6 hours, the drug is concentrated in the thyroid and maintains higher levels there for up to 24 hours after a dose [231].

The program is similar to that employed in adult thyrotoxicosis. It is sensible to see the child once each month, and at that time to make sure that the program is being followed and progress made. Any evidence of depression of the bone marrow should prompt a change to an alternative drug or a different form of treatment, as discussed below.

At the end of one or two years the medication is withdrawn. If thyrotoxicosis recurs, a second course of treatment lasting for one year or more may be given. A decrease in the size of the goiter during therapy is good evidence that a remission has been achieved. Progressive enlargement of the gland during therapy implies that hypothyroidism has been produced. This enlargement can be controlled by reduction in the dose of antithyroid drug or by administration of replacement thyroid hormone. There is no adequate rule for deciding when medical therapy has failed. After courses of antithyroid drug therapy totaling two to six years and attainment of age 15, if the patient still has not entered a permanent remission it is probably best to proceed with surgical or 131-I treatment. Barrio et al (225) reported on truly long term antithyroid drug therapy, which achieved 40% remissions in pediatric patients, with average time to remission of 5.4 years. Non-remitters were cured by RAI or surgery. Leger reported a similar program with 50% of children appearing to enter a permanent remission (232). In an other study 72% of children treated for 2 years relapsed. Occasionally a drug reaction develops while the condition is being controlled with an antithyroid drug. A change to another thiocarbamide may be satisfactory, but patients should be followed carefully. If a reaction is seen again, or if severe neutropenia occurs, it is usually best to stop antithyroid drug therapy and (1) give potassium iodide and an agent such as propranolol and to proceed with surgery, or (2) to give 131-I.  RAI therapy will be necessary if surgery is contraindicated by uncontrollable thyrotoxicosis,for whatever reason, or with prior thyroidectomy.

 Table 9

Surface area-M2 Weight (lbs) Approximate daily dose of MMI (mg)
0.1 5 2-3
0.2 10 2-5
0.5 30 5-10
0.75 60 10
1.0 90 10-15
1.25 110 15-20
1.5 140 20
2.0 200 20-25

INTRAUTERINE AND NEONATAL THYROTOXICOSIS

Thyrotoxicosis in utero is a rare but recognized syndrome occurring in pregnant women with very high TSH-R stimulating Ab in serum, due to transplacental passage of antibodies. It can also develop in the neonate. It is possible to screen for this risk by assaying TSAb in serum of pregnant women with known current or prior Graves’ Disease. Intra-uterine thyrotoxicosis causes fetal tachycardia, failure to grow, acceleration of bone age, premature closure of sutures, and occasionally fetal death. Multiple sequential pregnancies with this problem have been recorded. Clinical diagnosis is obviously inexact. Antithyroid drugs can be given, but control of the dosage is uncertain [233]. Propylthiouracil is considered to be the safest drug to use in the first trimester, because of fetal anomalies attributed to methimazole exposure in early pregnancy( 234), with switching to MMI in the second and third trimesters..

Luton et al (233) provided their extensive experience in managing these difficult cases. Measurement of TSAb is important. Mothers with negative TSAb assay, and not on ATD, rarely have any fetal problem. Mothers with positive TSAb or on ATD must be monitored by following maternal hormone and TSH levels, fetal growth, heart rate, and by ultrasound for evidence of goiter or other signs of fetal hyper- or hypothyroidism. If maternal hormone levels are low and TSH elevated, with fetal goiter and evidence of hypothyroidism, ATD therapy is reduced and intra-amniotic T4 may be given. If maternal T4 levels are high and TSH low, with fetal goiter and signs of fetal hyperthyroidism, increased doses of ATD are suggested. If the probable metabolic status of the fetus is not clear, fetal blood sampling is feasible although carrying significant risk to the fetus. Plasmapheresis to reduce maternal TRAb has been recommended, but few facts are available.

Thyrotoxicosis is rare in the newborn infant and is usually associated with past or present maternal hyperthyroidism [235,236]. Neonatal hypermetabolism usually arises from transplacental passage of TSAb. Frequently the infant is not recognized as thyrotoxic at birth, but develops symptoms of restlessness, tachycardia, poor feeding, occasionally excessive hunger, excessive weight loss, and possibly fever and diarrhea a few days after birth. The fetus converts T4 to T3 poorly in utero, but switches to normal T4 to T3 deiodination at birth. This phenomenon may normally provide a measure of protection in utero that is lost at birth, allowing the development of thyrotoxicosis in a few days. The syndrome may persist for two to five weeks, until the effects of the maternal antibodies have disappeared. The patient may be treated with propranolol, antithyroid drugs given according to the schedule above, and iodide. The antithyroid drug can be given parenterally if necessary in saline solution after sterilization by filtration through a Millipore filter. Newborn infants with thyrotoxicosis are frequently extremely ill, and ancillary therapy, including sedation, cooling, fluids in large amounts, electrolyte replacement, and oxygen, are probably as important in management as specific therapy for the thyrotoxicosis. Propranolol is used to control the tachycardia (236). Because of the increased metabolism of such infants, attention to fluid balance and adequacy of nutrition are important.

The patient usually survives the thyrotoxicosis, and the disease is typically self-limiting, with the euthyroid state being established in one or two months. Antithyroid medication can be gradually withdrawn at this time.

Graves’ disease can also occur in the newborn because the same disturbance that is causing the disorder in the mother is also occurring independently in the child. Hollingsworth et al [2379] described their experience in such patients. The mothers did not necessarily have active disease during pregnancy. Graves’ disease persisted in these patients from birth far beyond the time during which TSAb of maternal origin could persist. Advanced bone age was one feature of the disorder. Behavioral disturbances were later found in some of these children at a time when they were euthyroid.

General Therapeutic Relationship of the Patient and Physician

The foregoing discussion explains several methods for specifically decreasing thyroid hormone formation. They are, in a sense, both unphysiologic and traumatic to the patient. As a good physician realizes in any problem, but especially in Graves’ disease, attention to the whole patient is mandatory.

During the initial and subsequent interviews, the physician caring for a patient with Graves’ disease should recognize any  psychological and physical stresses. Frequently major emotional problems come to light after the patient recognizes the sincere interest of the physician. Typically the problem involves interpersonal relationships and often is one of matrimonial friction. The upset may be deep-seated and may involve very difficult adjustments by the patient, but characteristically it is related to identifiable factors in the environment. To put it another way, the problem is not an endogenous emotional reaction but a difficult adjustment to real external problems. On the other hand, one must be aware that the emotional lability of the thyrotoxic patient may be a trial for those with whom he or she must live, as well as for the patient. Thus thyrotoxicosis itself may create interpersonal problems. From whatever cause they arise, these problems are dealt with insofar as possible by the wise physician.

We have been unimpressed by the benefits of formal psychiatric care for the average thyrotoxic patient, but are certain that sympathetic discussion by the physician, possibly together with assistance in environmental manipulation, is an important part of the general attack on Graves’ disease. In other cases, personal problems may play a less important etiologic role but may still strongly affect therapy by interfering with rest or by causing economic hardship.

In addition to providing assistance in solving personal problems, two other general therapeutic measures are important. The first is rest. The patient with Graves’ disease should have time away from normal duties to help in reestablishing his or her psychic and physiologic equilibria. Patients can and do recover with appropriate therapy while continuing to work, but more rapid and certain progress is made if a period away from the usual occupation can be provided. Often a mild sedative or tranquilizer is helpful.

Another important general measure is attention to nutrition. Patients with Graves’ disease are nutritionally depleted in proportion to the duration and severity of their illness. Until metabolism is restored to normal, and for some time afterward, the caloric and protein requirements of the patient may be well above normal. Specific vitamin deficiences may exist, and multivitamin supplementation is indicated. The intake of calcium should be above normal.

REFERENCES

  1. Wayne EJ: The diagnosis of thyrotoxicosis. Br Med J 1:411, 1954.

1.1.Sawin CT, Geller A, Kaplan MM, Bacharach P, Wilson PWF, Hershman JM. Low serum thyrotropin (thyroid stimulating hormone) in older persons without hyperthyroidism. Arch Intern Med 151:165-168, 1991.

  1. Morley JE, Shafer RB, Elson MK, Slag MF, Raleigh MJ, Brammer GL, Yuwiler A, Hershman JM: Amphetamine-induced hyperthyroxinemia. Ann Intern Med 93:707, 1980.
  2. Brown ME, Refetoff S: Transient elevation of serum thyroid hormone concentration after initiation of replacement therapy in myxedema. Ann Intern Med 92:491, 1980.
  3. Kaptein EM, Macintyre SS, Weiner JM, Spencer CA, Nicoloff JT: Free thyroxine estimates in nonthyroidal illness: Comparison of eight methods. J Clin Endocrinol Metab 52:1073, 1981.
  4. Engler D, Donaldson EB, Stockigt JR, Taft P: Hyperthyroidism without triiodothyronine excess: An effect of severe nonthyroidal illness. J Clin Endocrinol Metab 46:77, 1978.
  5. Mayfield RK, Sagel J, Colwell JA: Thyrotoxicosis without elevated serum triiodothyronine levels during diabetic ketoacidosis. Arch Intern Med 140:408, 1980.
  6. Sobrinho LG, Limbert ES, Santos MA: Thyroxine toxicosis in patients with iodine induced thyrotoxicosis. J Clin Endocrinol Metab 45:25, 1977.
  7. Sterling K, Refetoff S, Selenkow HA: T3 thyrotoxicosis due to elevated serum triiodothyronine levels. J Amer Med Assn 213:571, 1970.
  8. Hollander CS, Nihei N, Burday SZ, Mitsuma T, Shenkman L, Blum M: Clinical and laboratory observations in cases of triiodothyronine toxicosis confirmed by radioimmunoassay. Lancet 1:609, 1972.
  9. Hollander CS, Mitsuma T, Shenkman L, Stevenson C, Pineda G, Silva E: T3 toxicosis in an iodine-deficient area. Lancet 2:1276, 1972.
  10. Hollander CS, Mitsuma T, Kastin AJ, Shenkman L, Blum M, Anderson DG: Hypertriiodothyroninemia as a premonitory manifestation of thyrotoxicosis. Lancet 2:731, 1971.
  11. Ormston BJ, Alexander L, Evered DC, Clark F, Bird T, Appleton D, Hall R: Thyrotropin response to thyrotropin-releasing hormone in ophthalmic Graves' disease: Correlation with other aspects of thyroid function, thyroid suppressibility and activity of eye signs. Clin Endocrinol 2:369, 1973.
  12. Franco PS, Hershman JM, Haigler ED, Pittman JA: Response to thyrotropin-releasing hormone compared with thyroid suppression tests in euthyroid Graves' disease. Metabolism 22:1357, 1973.
  13. Clifton-Bligh P, Silverstein GE, Burke G: Unresponsiveness to thyrotropin-releasing hormone (TRH) in treated Graves' hyperthyroidism and in euthyroid Graves' disease. J Clin Endocrinol Metab 38:531, 1974.

15.1. Murakami M, Miyashita K, Kakizaki S, Saito S, Yamada M, Iriuchijima T, Takeuchi T, Mori M. Clinical usefulness of thyroid-stimulating antibody measurement using Chinese hamster ovary cells expressing human thyrotropin receptors. Eur J Endocrinol 1995:80-86, 1995.

15.2  Costagliola S, Morgenthaler NG, Hoermann R, Badenhoop K, Struck J, Freitag D, Poertl S, Weglohner W, Hollidt JM, Quadbeck B, Dumont JE, Schumm-Draeger PM, Bergmann A, Mann K, Vassart G, Usadel KH.  Second generation assay for thyrotropin receptor antibodies has superior diagnostic sensitivity for Graves' disease.J Clin Endocrinol Metab. 1999 Jan;84(1):90-7.

15.3. Takasu N, Oshiro C, Akamine H, Komiya I, Nagata A, Sato Y, Yoshimura H, Ito K. Thyroid-stimulating antibody and TSH-binding inhibitor immunoglobulin in 277 Graves patients and in 686 normal subjects. J Endocrinol Invest 20:452-461, 1997.

15.4. Feldt-Rasmussen U, Schleusener H, Carayon P. Meta-analysis evaluation of the impact of thyrotropin receptor antibodies on long-term remission after medical therapy of Graves disease. J Clin Endocrinol Metab 78:98-102, 1994.

  1.  Watson SG, Radford AD, Kipar A, Ibarrola P, Blackwood L Somatic mutations of the thyroid-stimulating hormone receptor gene in feline hyperthyroidism: parallels with human hyperthyroidism.J Endocrinol. 2005 Sep;186(3):523-37.
  2. Namba H, Ross JL, Goodman D, Fagin JA: Solitary polyclonal autonomous thyroid nodule: A rare cause of childhood hyperthyroidism. J Clin Endocrinol Metab 72:1108-1112, 1991.
  3. Shimaoka K, Van Herle AJ, Dindogru A: Thyrotoxicosis secondary to involvement of the thyroid with malignant lymphoma. J Clin Endocrinol Metab 43:64, 1976.
  4. Emerson CH, Utiger RD: Hyperthyroidism and excessive thyrotropin secretion. N Engl J Med 287:328, 1972
  5. Spanheimer RG, Bar RS, Hayford JC: Hyperthyroidism caused by inappropriate thyrotropin hypersecretion. Studies in patients with selective pituitary resistance to thyroid hormone. Arch Intern Med 142:1283-1286, 1982.
  6. Gershengorn ML: Thytropin-induced hyperthyroidism caused by selective pituitary resistance to thyroid hormone. A new syndrome of "inappropriate secretion of TSH". J Clin Invest 56:271, 1975.
  7. Woolf PD, Daly R: Thyrotoxicosis with painless thyroiditis. Am J Med 60:73, 1976.
  8. Gluck FB, Nusynowitz ML, Plymate S: Chronic lymphocytic thyroiditis, thyrotoxicosis, and low radioactive iodine uptake. N Engl J Med 293:624, 1975.
  9. Ginsberg J, Walfish PG: Post-partum transient thyrotoxicosis with painless thyroiditis. Lancet 1:1125, 1977.
  10. Amino N, Yabu Y, Miyai K, Fujie T, Azukizawa M, Onishi T, Kumahara Y: Differentiation of thyrotoxicosis induced by thyroid destruction from Graves' disease. Lancet 1:344, 1978.
  11. Inada M, Nishikawa M, Naito K, Ishii H, Tanaka K, Imura H: Reversible changes of the histological abnormalities of the thyroid in patients with painless thyroiditis. J Clin Endocrinol Metab 52:431, 1981.
  12. Yabu Y, Amino N, Mori H, Miyai K, Tanizawa O, Takai S-I, Kumahara Y, Matsusuka F, Kuma K: Postpartum recurrence of hyperthyroidism and changes of thyroid-stimulating immunoglobulins in Graves' disease. J Clin Endocrinol Metab 51:1454, 1980.
  13. Amino N, Miyai K, Azukizawa M, Yabu Y, Fujie T, Onishi T, Kumahara Y: Differentiation of thyrotoxicosis induced by thyroid destruction from Graves' disease. Lancet 2:344, 1978.
  14. Shigemasa C, Ueta Y, Mitani Y, Taniguchi S, Urabe K, Tanaka T, Yoshida A, Mashiba H: Chronic thyroiditis with painful tender thyroid enlargement and transient thyrotoxicosis. J Clin Endocrinol Metab 70:385, 1990.

29.1. Sarlis NJ, Brucker-Davis F, Swift JP, Tahara K, Kohn LD. Graves’ disease following thyrotoxic painless thyroiditis. Analysis of antibody activities against the thyrotropin receptor in two cases. Thyroid 7:829, 1997

29.12. Yoshimura M, Hershman JM. Thyrotropic action of human chorionic gonadotropin. Thyroid 5:425-434, 1995.

29.13 Hershman JM. 1999 Human chorionic gonadotropin and the thyroid: hyperemesis gravidarum and trophoblastic tumors. Thyroid 9:653.

29.14. Goodwin TM, Hershman JM. Hyperthyroidism due to inappropriate production of human chorionic gonadotropin. Clin Obstet Gynecol 40:32-44, 1997.

29.2. Rodien P. Bremont C. Sanson ML. Parma J. Van Sande J. Costagliola S. Luton JP. Vassart G. Duprez L. Familial gestational hyperthyroidism caused by a mutant thyrotropin receptor hypersensitive to human chorionic gonadotropin. New England Journal of Medicine. 339(25):1823-6, 1998 .

29.3. Fuhrer D, Wonerow P, Willgerodt H, Paschke R. Identification of a new thyrotropin receptor germline mutation (Leu629 Phe) in a family with neonatal onset of autosomal dominant nonautoimmune hyperthyroidism. J Clin Endocrinol Metab 82:4234-4238,

29.4. Ginsberg J, Lewanczuk RZ, Honore LH.  Hyperplacentosis:  a novel cause of hyperthyroidism.  Thyroid 11:393-396, 20011997.

  1. Cave WT Jr, Dunn JT: Choriocarcinoma with hyperthyroidism: Probable identity of the thyrotropin with human chorionic gonadotropin. Ann Intern Med 85:60, 1976.
  2. Nagataki S, Mizuno M, Sakamoto S, Irie M, Shizume K, Nakao K, Galton VA, Arky RA, Ingbar SH: Thyroid function in molar pregnancy. J Clin Endocrinol Metab 44:254, 1977.
  3. Tsuruta E, Tada H, Tamaki H, Kashiwai T, Asahi K, Takeoka K, Mitsuda N, Amino N. Pathogenic role of asialo human chorionic gonadotropin in gestational thyrotoxicosis. J Clin Endocrinol Metab 80:350-355, 1995.
  4. Miyai K, Tanizawa O, Yamamoto T, Azukizawa M, Kawai Y, Noguchi M, Ishibas K, Kumahara Y: Pituitary-thyroid function in trophoblastic disease. J Clin Endocrinol Metab 42:254, 1976.

33.1 Doi F, Kakizaki S, Takagi H, Murakami M, Sohara N, Otsuka T, Abe T, Mori M .Long-term outcome of interferon-alpha-induced autoimmune thyroid disorders in chronic hepatitis C.Liver Int. 2005 Apr;25(2):242-6

33.2 Chen F, Day SL, Metcalfe RA, Sethi G, Kapembwa MS, Brook MG, Churchill D, de Ruiter A, Robinson S, Lacey CJ, Weetman AP.Characteristics of autoimmune thyroid disease occurring as a late complication of immune reconstitution in patients with advanced human immunodeficiency virus (HIV) disease.Medicine (Baltimore). 2005 Mar;84(2):98-106.

33a. Sgarbi, JA; Villaca, FG; Garbeline, B; Villar, HE; Romaldini, JH.          The effects of early antithyroid therapy for endogenous subclinical hyperthyroidism in clinical and heart abnormalities.                 J Clin Endocrinol Metab     88           1672-1677                2003

33b Woeber KA.Observations concerning the natural history of subclinical hyperthyroidism.Thyroid. 2005 Jul;15(7):687-91

  1. Lahey FH: Apathetic thyroidism. Ann Surg 93:1026, 1931.
  2. Philip JR, Harrison MT, Ridley EF, Crooks J: Treatment of thyrotoxicosis with ionizing radiation. Lancet 2:1307, 1968.

36.1. Franklyn, JA. The management of hyperthyroidism. New Engl J Med 330:1731-1738, 1994.

36.2. Torring O, Tallstedt L, Wallin G, Lundell G, Lunggren J-G, Taube A, Saaf M, Hamberger B, Thyroid Study Group. Graves’ hyperthyroidism: treatment with antithyroid drugs, surgery, or radioiodine – A prospective, randomized study. J Clin Endocrinol Metab 81:2986-2993, 1996.

36.3. Vitti P, Rago T, Chiovato L, Pallini S, Santini F, Fiore E, Rocchi R, Martino E, Pinchera A. Clinical features of patients with Graves’ disease undergoing remission after antithyroid drug treatment. Thyroid 7:369, 1997.

36.31 Burch HB1, Cooper DS2. Management of Graves Disease: A Review.JAMA. 2015 Dec 15;314(23):2544-54. doi: 10.1001/jama.2015
36.4 Scholz, GH; Hagemann, E; Arkenau, C; Engelmann, L; Lamesch, P; Schreiter, D; Schoenfelder, M; Olthoff, D; Paschke, R.  Is there a place for thyroidectomy in older patients with thyrotoxic storm and cardiorespiratory failure?      Thyroid    13            933          2003

  1. Glinoer D, Hesch D, LaGasse R, Laurberg P: The management of hyperthyroidism due to Graves' disease in Europe in 1986. Results of an international survey. Proceedings of the Symposium held during the 15th Annual Meeting of the European Thyroid Association in Stockholm, June - July, 1986, 37 pages.
  2. Solomon B, Glinoer D, LaGasse R, Wartofsky L: Current trends in the management of Graves' disease. J Clin Endocrinol Metab 70:1518-1524, 1990.
  3. Dobyns BM, Sheline GE, Workman JB, Tompkins EA, McConahey WM, Becker DV: Malignant and benign neoplasms of the thyroid in patients treated for hyperthyroidism: a report of the Cooperative Thyrotoxicosis Therapy Follow-up Study. J Clin Endocrinol Metabl 37:976-998, 1974.
  4. Safa AM, Schumacher P, Rodriguez-Antunez A: Long-term follow-up results in children and adolescents treated with radioactive iodine (131-Iodine). N Engl J Med 292:167-171, 1975.
  5. Holm LE, Dahlqvist I, Israelsson A, Lundell GM: Malignant thyroid tumors after 131-iodine therapy. N Engl J Med 303:188-191, 1980.
  6. Hamburger JI: Management of hyperthyroidism in children and adolescents. J Clin Endocrinol Metab 60:1019, 1985.
  7. Freitas JE, Swanson DP, Gross MD, Sisson JS: Iodine-131: Optimal therapy for hyperthyroidism in children and adolescents? J Nucl Med 20:847, 1979.
  8. Hayek A, Chapman E, Crawford JD: Long-term results of treatment of thyrotoxicosis in children and adolescents with radioactive iodine. N Engl J Med 283:949, 1970.

44.1.Franklyn JA, Maisonneuve P, Sheppard M, Betteridge PB. Cancer indicdence and mortality after radioiodine treatment for hyperthyroidism: a population based cohort study. Lancet 353:2111-15, 1999

44.2 Rivkees SA, Dinauer C. An optimal treatment for pediatric Graves' disease is radioiodine.J Clin Endocrinol Metab. 2007 Mar;92(3):797-800.

44.3. Torring O, Tallstedt L, Wallin G, Lundell G, Ljunggren J-G, Taube A, Saaf M, Hamberger B, and The Thyroid Study Group. Graves' hyperthyroidism: Treatment with antithyroid drugs, surgery, or radioiodine-a prospective, randomized study. J Clin Endocrinol Metab 81:2986-2993,

44,4. Moleti M, Mattina F, Lo Presti VP, Baldari CS, Bonanno N, Trimarchi F, Vermiglio F. Role of residual thyroid tissue ablation after thyroidectomy for Graves' disease. Its effects on the course of related ophthalmopathy. J Endocrinol Invest 23:37, 2000.

44.5. De Bellis A204c. De Bellis A, Bizzarro A, Perrino S, Coronclla C, Iorio S, Pepe M, Guaglione M, Wall JR, Bellastella A. Improvement of severe ophthalmopathy and decrease of antibodies against extraocular muscles, G2s, and Fp subunit of succinate dehydrogenase after near-total thyroidectomy in Graves' disease. J Endocrinol Invest 23:14, 2000.

44.6. Marcocci C204d. Marcocci C, Bruno-Bossio G, Manetti L, Tanda ML, Miccoli P, Iacconi P, Bartolomei MP, Nardi M, Pinchera A, Bartalena L. The course of Graves' ophthalmopathy is not influenced by near-total thyroidectomy; a case-control study. Clin Endocrinol 51:503-508, 1999.

44.7  Moleti, M; Mattina, F; Salamone, I; Violi, MA; Nucera C; Baldari, S; Schiavo, MGL; Regalbuto, C; Trimarchi, F; Vermiglio, F. Effects of thyroidectomy alone or followed by radioiodine ablation of thyroid remnants on the outcome of Graves’ ophthalmopathy. Thyroid 13 653-658 2003.

  1. Chapman EM: History of the discovery and early use of radioactive iodine. J Amer Med Assn 250:2042-2044, 1983.
  2. Chapman EM, Maloof F: The use of radioactive iodine in the diagnosis and treatment of hyperthyroidism: Ten years' experience. Medicine 34:261, 1955.
  3. Blahd W, Hays MT: Graves' disease in the male. A review of 241 cases treated with an individually calculated dose of sodium iodide 131-I. Arch Intern Med 129:33, 1972.

47.1 Bajnok L, Mezosi E, Nagy E, Szabo J, Sztojka I, Varga J, et al. 1999 Calculation of the radioiodine dose for the treatment of Graves' hyperthyroidism: Is more than seven-thousand rad target dose necessary? Thyroid 9:865.

47.2  Leslie, WD; Ward, L; Salamon, EA; Ludwig, S; Rowe, RC; Cowden, EA.  A randomized comparison of radioiodine doses in Graves’ hyperthyroidism.           J Clin Endocrinol Metab           88            978-983   2003

  1. Hagen F, Ouelette RP, Chapman EM: Comparison of high and low dosage levels of 131-I in the treatment of thyrotoxicosis. N Engl J Med 277:559, 1967.
  2. Cevallos JL, Hagen GA, Maloof F, Chapman EM: Low-dosage 131-I therapy of thyrotoxicosis (diffuse goiters). N Engl J Med 290:141, 1974.
  3. Ross DS, Daniels GH, De Stafano P, Maloof F, Ridgway EC: Use of adjunctive potassium iodide after radioactive iodine (131-I) treatment of Graves' hyperthyroidism. J Clin Endocrinol Metab 57:250, 1983.
  4. Reinwein D, Schaps D, Berger H, Hackenberg K, Horster FA, Klein E, Von Zur Muhlen A, Wendt RU, Wildmeister W: Hypothyreoserisyiko nach fraktionierter Radiojodtherapie. Dtsch Med Wochenschr 98:1789, 1973.
  5. Rapoport B, Caplan R, DeGroot L: Low-dose sodium iodide 131-I therapy in Graves' disease. J Amer Med Assn 224:1610, 1973.
  6. Roudebush CP, Hoye KE, DeGroot LJ: Compensated low-dose 131-I therapy of Graves' disease. Ann Intern Med 87:441, 1977.
  7. Sridama V, McCormick M, Kaplan EL, Fauchet R, DeGroot LJ: Long-term follow-up study of compensated low-dose 131-I therapy for Graves' disease. N Engl J Med 311:426-432, 1984.
  8. Glennon JA, Gordon ES, Sawin CT: Hypothyroidism after low- dose I131 treatment of hyperthyroidism. Ann Intern Med 76:721, 1972.
  9. Saito S, Sakurada T, Yamamoto M, Yoshida K, Kaise K, Kaise N, Yoshinaga K: Long term results of radioiodine 131-I therapy in 331 patients with Graves' disease. Tokyo J Exp Med 132:1-10, 1980.
  10. DeGroot LJ, Mangklabruks A, McCormick M: Comparison of RA 131-I treatment protocols for Graves' disease. J Endocrinol Invest 13:111-118, 1990.

57.1.  Rini JN, Vallabhajosula S, Zanzonico P, Hurley JR, Becker DV, Goldsmith SJ. Thyroid uptake of liquid versus capsule 131-I tracers in hyperthyroid patients treated with liquid 131-I. Thyroid 9:347, 1999

57.2. Berg G, Michanek A, Holmberg E, Nystrom E. Clinical outcome of radioiodine treatment of hyperthyroidism: a follow-up study. J Intern Med 239:165-171, 1996.

57.3-Schiavo M, Bagnara MC, Calamia I, Bossert I, Ceresola E, Massaro F, Giusti M, Pilot A, Pesce G, Caputo M, Bagnasco M.A study of the efficacy of radioiodine therapy with individualized dosimetry in Graves' disease: need to retarget the radiation committed dose to the thyroid. J Endocrinol Invest. 2011 Mar;34(3):201-5. Epub 2010 Dec 15

57.4 Chen DY, Schneider PF, Zhang XS, He ZM, Jing J, Chen TH Striving for euthyroidism in radioiodine therapy of Graves' disease: a 12-year prospective, randomized, open-label blinded end point study. Thyroid. 2011 Jun;21(6):647-54. doi: 10.1089/thy.2010.0348. Epub 2011 May 12

  1. Wise PH, Burnet RB, Ahmad A, Harding PE: Intentional radioiodine ablation in Graves' disease. Lancet 2:1231, 1975.

58.1 Sapienza MT1, Coura-Filho GB, Willegaignon J, Watanabe T, Duarte PS, Buchpiguel CA. Clinical and Dosimetric Variables Related to Outcome After Treatment of Graves' Disease With 550 and 1110 MBq of 131I: Results of a Prospective Randomized Trial. Clin Nucl Med. 2015 Sep;40(9):715-9. doi: 10.1097

58.2.  Allahabadia A, Daykin J, Sheppard MC, Gough SCL, Franklyn JA.  Radioiodine treatment of hyperthyroidism—prognostic factors for outcome.  J Clin Endocrinol Metab 86:3611-3617, 2001.

58.3:  Leslie WD, Ward L, Salamon EA, Ludwig S, Rowe RC, Cowden EA.  A randomized comparison of radioiodine doses in Graves' hyperthyroidism.J Clin Endocrinol Metab. 2003 Mar;88(3):978-83.

58.4 Santos RB, Romaldini JH, Ward LS A randomized controlled trial to evaluate the effectiveness of 2 regimens of fixed iodine (¹³¹I) doses for Graves disease treatment.Clin Nucl Med. 2012 Mar;37(3):241-4.

58.5 Bogazzi F, Giovannetti C, Fessehatsion R, Tanda ML, Campomori A, Compri E, Rossi G, Ceccarelli C, Vitti P, Pinchera A, Bartalena L, Martino E. J Clin Endocrinol Metab. 2010 Jan;95(1):201-8 Impact of lithium on efficacy of radioactive iodine therapy for Graves' disease: a cohort study on cure rate, time to cure, and frequency of increased serum thyroxine after antithyroid drug withdrawal.

58.6 Bonnema SJ, Bennedbaek FN, Veje A, Marving J, Hegedus L.Continuous methimazole therapy and its effect on the cure rate of hyperthyroidism using radioactive iodine: an evaluation by a randomized trial. J Clin Endocrinol Metab. 2006 Aug;91(8):2946-51.

  1. Marcocci C, Gianchecchi D, Masini I, Golia F, Ceccarelli C, Bracci E, Fenzi GF, Pinchera A: A reappraisal of the role of methimazole and other factors on the efficacy and outcome of radioiodine therapy of Graves' hyperthyroidism. J Endocrinol Invest 13:513-520, 1990.

59.1. Glinoer D, Verelst J. Use of 131-Iodine for the treatment of hyperthyroidism in adults. Annales d Endocrinologie 57:177-185, 1996.

59.2 Nakazato N, Yoshida K, Mori K, Kiso Y, Sayama N, Tani J-I, et al. 1999 Antithyroid drugs inhibit radioiodine-induced increases in thyroid autoantibodies in hyperthyroid Graves' disease. Thyroid 9:775.

59.3.   Andrade VA, Gross JL, Maia AL.  The effect of methimazole pretreatment on the efficacy of radioactive iodine therapy in Graves’ hyperthyroidism:  one-year follow-up of a prospective, randomized study.  J Clin Endocrinol Metab 86:3488-3493, 2001.

59.31. Burch HB, Solomon BL, Cooper DS, Ferguson P, Walpert N, Howard R.  The effect of antithyroid drug pretreatment on acute changes in thyroid hormone levels after 131-I ablation for Graves’ disease.  J Clin Endocrinol Metab 86:3016-3021, 2001.

59.4 Zakavi SR1, Khazaei G, Sadeghi R, Ayati N, Davachi B, Bonakdaran S, Jabbari Nooghabi M, Moosavi Z. Methimazole discontinuation before radioiodine therapy in patients with Graves' disease. Nucl Med Commun. 2015 Dec;36(12):1202-7. doi: 10.1097.

  1. Greig WR, Gillespie FC, Thomson JA, McGirr EM: Iodine-125 treatment for thyrotoxicosis. Lancet 1:755, 1969.
  2. Editorial: Radioiodine treatment of thyrotoxicosis. Lancet 1:23,1972.
  3. Bremmer WF, Greig WR, McDougall IR: Results of treating 297 thyrotoxic patients with 125I. Lancet 2:281, 1973.

62.1. Liu B1, Tian R1, Peng W1, He Y1, Huang R1, Kuang A1.Radiation Safety Precautions in (131)I Therapy of Graves' Disease Based on Actual Biokinetic Measurements. J Clin Endocrinol Metab. 2015 Aug;100(8):2934-41. doi: 10.1210/jc.2015-1682

  1. Benua RS, Dobyns BM: Isolated compounds in the serum, disappearance of radioactive iodine from the thyroid, and clinical response in patients treated with radioactive iodine. J Clin Endocrinol Metab 15:118, 1955.

63.1. Stensvold AD, Jorde R, Sundsfjord J. Late and transient increases in free T4 after radioiodine treatment for Graves’ disease. J Endocrinol Invest 20:580-584, 1997.

  1. Stanbury JB, Janssen MA: The iodinated albumin-like component of the plasma of thyrotoxic patients. J Clin Endocrinol Metab 22:978, 1962.

64.1.  Dale J, Daykin J, Holder R, Sheppard MC, Franklyn JA.  Weight gain following treatment of hyperthyroidism.  Clin Endocrinol 55:233-239, 2001.

  1. Goldsmith RE: Radioisotope therapy for Graves' disease. Mayo Clin Proc 47:953, 1972.
  2. Slingerland WD, Hershman JM, Dell E, Burrows B: Thyrotropin and PBI in radioiodine-treated hyperthyroid patients. J Clin Endocrinol Metab 35:912, 1972.
  3. Gordin A, Wagar G, Hernberg CA: Serum thyrotropin and response to thyrotrophin-releasing hormone in patients who are euthyroid after radioiodine treatment for hyperthyroidism. Acta Med Scand 194:335, 1973.
  4. Baldwin WW: Graves' disease succeeded by atrophy. Lancet 1:145,1895.
  5. Wood LC, Ingbar SH: Hypothyroidism as a late sequela in patients with Graves' disease treated with antithyroid drugs. J Clin Invest 64:1429, 1979.

69a. Tigas S, Idiculla J, Beckett G, Toft A. Is excessive weight gain after ablative treatment of hyperthyroidism due to inadequate thyroid hormone therapy? Thyroid 10:1107, 2000.

  1. Creutzig H, Kallfelz I, Haindl J, Thiede G, Hundeshagen H: Thyroid storm and iodine-131 treatment. Lancet 2:145, 1976.
  2. Creutzig H, Kallfelz I, Haindl J, Thiede G, Hundeshagen H: Thyroid storm and iodine-131 treatment. Lancet 2:145, 1976.
  3. Lamberg BA, Hernberg CA, Wahlberg P, Hakkila R: Treatment of toxic nodular goiter with radioactive iodine. Acta Med Scand 165:245, 1959.
  4. McDermott MT, Kidd GS, Dodson Jr LE, Hofeldt FD: Radioiodine-induced thyroid storm. Case report and literature review. Amer J Med 75:353, 1983.
  5. Townsend JD: Hypoparathyroidism following radioactive iodine therapy for intractable angina pectoris. Ann Intern Med 55:662, 1961.
  6. Gilbert-Dreyfus MZ, Gali P: Cataract due to tetany following radioactive iodine therapy. Sem Hop Paris 34:1301, 1958.
  7. Fulop M: Hypoparathyroidism after 131-I therapy. Ann Intern Med 75:808, 1971.

76.1 Ceccarelli C, Canale D, Battisti P, Caglieresi C, Moschini C, Fiore E, Grasso L, Pinchera A, Vitti P.Testicular function after 131-I therapy for hyperthyroidism.Clin Endocrinol (Oxf). 2006 Oct;65(4):446-52.

76.2. Franklyn JA, Sheppard MC, Maisonneuve P. Thyroid function and mortality in patients treated for hyperthyroidism. JAMA. 2005 Jul 6;294(1):71-80.

  1. Fenzi G, Hashizume K, Roudebush C, DeGroot LJ: Changes in thyroid stimulating immunoglobulins during antithyroid therapy. J Clin Endocrinol Metab 48:572, 1979.
  2. Teng CS, Yeung RTT, Khoo RKK, Alagaratnam TT: A prospective study of the changes in thyrotropin binding inhibitory immunoglobulins in Graves' disease treated by subtotal thyroidectomy or radioactive iodine. J Clin Endocrinol Metab 50:1005, 1980.
  3. Sridama V, DeGroot LJ: Treatment of Graves' disease and the course of ophthalmopathy. Amer J Med 87:70-73, 1989.
  4. Tallstedt L, Lundell G, Torring O, Wallin G, Ljunggren J-G, Blomgren H, Taube A. Occurrence of ophthalmopathy after treatment for Graves' disease. N Engl J Med 326:1733-1738, 1992.

80.1. Fernandez Sanchez JR, Rosell Pradas J, Carazo Martinez O, Torres Vela E, Escobar Jimenez F, Garbin Fuentes I, Vara Thorbeck R. Graves’ ophthalmopathy after subtotal thyroidectomy and radioiodine therapy. Brit J Surg 80:1134-1136, 1993.

80.2 Vannucchi G, Campi I, Covelli D, Dazzi D, Currò N, Simonetta S, Ratiglia R, Beck-Peccoz P, Salvi M.

J Clin Endocrinol Metab. 2009 Sep;94(9):3381-6 Graves' orbitopathy activation after radioactive iodine therapy with and without steroid prophylaxis.

Eur J Endocrinol. 2016 Apr;174(4):491-502. doi: 10.1530/EJE-15-1099. Epub 2016 Jan 15.

80.3.Taïeb D1, Bournaud C2, Eberle MC2, Catargi B2, Schvartz C2, Cavarec MB2, Faugeron I2, Toubert ME2, Benisvy D2, Archange C2, Mundler O2, Caron P2, Abdullah AE2, Baumstarck K2. Eur J Endocrinol. 2016 Apr;174(4):491-502. doi: 10.1530/EJE-15-1099. Quality of life, clinical outcomes and safety of early prophylactic levothyroxine administration in patients with Graves' hyperthyroidism undergoing radioiodine therapy: a randomized controlled study.

 

  1. Gamstedt A, Wadman B, Karlsson A: Methimazole, but not betamethasone, prevents 131-I treatment-induced rises in thyrotropin receptor autoantibodies in hyperthyroid Graves' disease. J Clin Endocrinol Metab 62:773-777, 1986.
  2. Bartalena L, Marcocci C, Bogazzi F, Panicucci M, Lepri A, Pinchera A: Use of corticosteroids to prevent progression of Graves' ophthalmopathy after radioiodine therapy for hyperthyroidism. N Engl J Med 321:1349-1352, 1989.

82.1. Bartalena L, Marcocci C, Bogazzi F, Manetti L, Tanda ML, Dell’Unto E, Bruno-Bossio G, Nardi M, Bartolomei MP, Lepri A, Rossi G, Martino E, Pinchera A. Relation between therapy for hyperthyroidism and the course of Graves’ ophthalmopathy. N Engl J Med 338:73-78, 1998.

82.2 Jensen BE, Bonnema SJ, Hegedus L.Glucocorticoids do not influence the effect of radioiodine therapy in Graves' disease.Eur J Endocrinol. 2005 Jul;153(1):15-21.

82.3 Takamura Y, Nakano K, Uruno T, Ito Y, Miya A, Kobayashi K, Yokozawa T,Matsuzuka F, Kuma K, Miyauchi A.  Changes in serum TSH receptor antibody (TRAb) values in patients with Graves' disease after total or subtotal thyroidectomy. Endocr J. 2003 Oct;50(5):595-601.

82.4 De Bellis A, Conzo G, Cennamo G, Pane E, Bellastella G, Colella C, Iacovo AD, Paglionico VA, Sinisi AA, Wall JR, Bizzarro A, Bellastella ATime course of Graves' ophthalmopathy after total thyroidectomy alone or followed by radioiodine therapy: a 2-year longitudinal study.Endocrine. 2012 Apr;41(2):320-6. Epub 2011 Nov 16.

82.5 Leo M, Marcocci C, Pinchera A, Nardi M, Megna L, Rocchi R, Latrofa F, Altea MA, Mazzi B, Sisti E, Profilo MA, Marinò M. Outcome of Graves' orbitopathy after total thyroid ablation and glucocorticoid treatment: follow-up of a randomized clinical trial.J Clin Endocrinol Metab. 2012 Jan;97(1):E44-8. Epub 2011 Oct 26.

82.6 Bojic T1, Paunovic I2,3, Diklic A4,5, Zivaljevic V6,7, Zoric G8, Kalezic N9,10, Sabljak V11,12, Slijepcevic N13, Tausanovic K14, Djordjevic N15,16, Budjevac D17, Djordjevic L18, Karanikolic A19,20. Total thyroidectomy as a method of choice in the treatment of Graves' disease - analysis of 1432 patients. BMC Surg. 2015 Apr 9;15:39. doi: 10.1186/s12893-015-0023-3.

  1. DeGroot LJ, Paloyan E: Thyroid carcinoma and radiation. A Chicago endemic. J Amer Med Assn 225:487, 1973.
  2. Clark DW: Association of irradiation with cancer of the thyroid in children and adolescents. J Amer Med Assn 159:1007, 1955.
  3. Simpson CL, Hempelmann LH, Fuller LM: Neoplasia in children treated with x-rays in infancy for thymic enlargement. Radiology 64:840, 1955.
  4. Doniach I: The effect of radioactive iodine alone and in combination with methylthiouracil upon tumor production in the rat's thyroid gland. Br J Cancer 7:181, 1953.
  5. Sampson RJ, Key CR, Buncher CR, Iijuma S: Thyroid carcinoma in Hiroshima and Nagasaki. J Amer Med Assn 209:65, 1969.
  6. Conard RA, Dobyns BM, Sutow W: Thyroid neoplasia as late effect of exposure to radioactive iodine in fallout. J Amer Med Assn 214:316, 1970.
  7. Pacini F, Vorontsova T, Demidchik E, Molinaro E, Agate L, Romei C, Shavrova E, Cherstvoy E, Ivashkevitch Y, Kuchinskaya E, Schlumberger M, Rouga G, Felesi M, Pinchera A. Post-Chernobyl thyroid carcinoma in Belarus children and adolescents: comparison with naturally occurring thyroid carcinoma in Italy and France. J Clin Endocrinol Metab 82:3563-3569, 1997.

89.2-Cardis E, Kesminiene A, Ivanov V, Malakhova I, Shibata Y, Khrouch V, Drozdovitch V, Maceika E, Zvonova I, Vlassov O, Bouville A, Goulko G, Hoshi M, Abrosimov A, Anoshko J, Astakhova L, Chekin S, Demidchik E, Galanti R, Ito M, Korobova E, Lushnikov E, Maksioutov M, Masyakin V, Nerovnia A, Parshin V, Parshkov E, Piliptsevich N, Pinchera A, Polyakov S, Shabeka N, Suonio E, Tenet V, Tsyb A, Yamashita S, Williams D. Risk of thyroid cancer after exposure to 131I in childhood.J Natl Cancer Inst. 2005 May 18;97(10):724-32.

  1. Pochin EE: Leukemia following radioiodine treatment of thyrotoxicosis. Br Med J 2:1545, 1960.
  2. Saenger EL, Thoma GE, Tompkins EA: Incidence of leukemia following treatment of hyperthyroidism. J Amer Med Assn 205:855, 1968.
  3. Biological effects of ionizing radiation V. The health effects of exposure to low levels of ionizing radiation, report of the Advisory Committee on the Biological Effects of Ionizing Radiation, National Research Council, Washington, DC, Natl Acad Press, 1990.
  4. Russell WL, Kelly EM: Mutation frequencies in male mice and the estimation of genetic hazards of radiation in men. Proc Natl Acad Sci USA 79:542, 1982.
  5. Webster EW, Merrill OE: Radiation hazards: II. Measurements of gonadal dose in radiographic examination. N Engl J Med 257:811, 1957.
  6. Robertson J, Gorman CA: Gonadal radiation dose and its genetic significance in radioiodine therapy of hyperthyroidism. J Nucl Med 17:826, 1976.
  7. Sarkar SD, Bierwaltes WH, Gill SP, Cowley BJ: Subsequent fertility and birth histories of children and adolescents treated with 131-I for thyroid cancer. J Nucl Med 17:460, 197
  8. Hollingsworth JW: Delayed radiation effects in survivors of the atomic bombings: A summary of the findings fo the Atomic Bomb Casualty Commission, 1947-1959. N Engl J Med 263:481, 1960.
  9. Plummer HS: Results of administering iodine to patients having exophthalmic goiter. J Amer Med Assn 80:1955, 1923.
  10. MacKenzie CG, MacKenzie JB: Effect of sulfonamides and thiourea on the thyroid gland and basal metabolism. Endocrinology 32:185, 1943.
  11. Astwood EB: Treatment of hyperthyroidism with thiourea and thiouracil. J Amer Med Assn 122:78, 1943.
  12. Marcocci C, Chiovato L, Mariotti S, Pinchera A: Changes of circula.1ing thyroid autoantibody levels during and after therapy with methi-mazole in patients with Graves' disease. J Endocrinol Invest 5:13, 1982.
  13. Pinchera A, Liberti P, Martino E, Fenzi GF, Grasso L, Rovis I, Baschieri L: Effects of antithyroid therapy on the long-acting thyroid stimulator and the antithyroglobulin antibodies. J Clin Endocrinol Metab 29:231, 1969.
  14. MacGregor AM, Ibbertson HK, Smith BR, Hall R: Carbimazole and autoantibody synthesis in Hashimoto's thyroiditis. Br Med J 281:968, 1980.
  15. McGregor AM, Petersen MM, McLachlan SM, Rooke P, Smith BR, Hall R: Carbimazole and the autoimmune response in Graves' disease. N Engl J Med 303:302, 1980.
  16. Hallengren B, Forsgren A, Melander A: Effects of antithyroid drugs on lymphocyte function in vitro. J Clin Endocrinol Metab 51:298, 1980.

105.1. Weetman AP. The immunomodulatory effects of antithyroid drugs. Thyroid 4:145-146, 1994.

  1. Ludgate ME, McGregor AM, Weetman AP, Ratanachaiyavong S, Lazarus JH, Hall R, Middleton GW: Analysis of T cell subsets in Graves' disease: alterations associated with carbimazole. Br Med J 288:526, 1984.
  2. Totterman TH, Karlsson FA, Bengtsson M, Mendel-Hartvig IB: Induction of circulating activated suppressor-like T cells by methimazole therapy for Graves' disease. N Engl J Med 316:15, 1987.
  3. Sridama V, Pacini F, DeGroot LJ: Decreased suppressor T- lymphocytes in autoimmune thyroid diseases detected by monoclonal antibodies. J Clin Endocrinol Metab 54:316, 1982.
  4. Volpe R. Evidence that the immunosuppressive effects of antithyroid drugs are mediated through actions on the thyroid cell, modulating thyrocyte-immunocyte signaling: A review. Thyroid 4:217-223, 1994.

109.1 Bahn RS, Burch HS, Cooper DS, Garber JR, Greenlee CM, Klein IL, Laurberg P, McDougall IR, Rivkees SA, Ross D, Sosa JA, Stan MN. Thyroid. 2009 Jul;19(7):673-4. The Role of Propylthiouracil in the Management of Graves' Disease in Adults: report of a meeting jointly sponsored by the American Thyroid Association and the Food and Drug Administration.

110.Laurberg P, Hansen PEB, Iversen E, Jensen SE, Weeke J: Goiter size and outcome of medical treatment of Graves' disease. Acta Endocrinol 111:39-43, 1986.

110.1. Laurberg P1, Krejbjerg A, Andersen SL. Relapse following antithyroid drug therapy for Graves' hyperthyroidism. Curr Opin Endocrinol Diabetes Obes. 2014 Oct;21(5):415-21. doi: 10.1097

  1. Shizume K, Irie M, Nagataki S, Matsuzaki F, Shishiba Y, Suematsu H, Tsushima T: Long-term result of antithyroid drug therapy for Graves' disease: Follow-up after more than 5 years. Endocrinol Jpn 17:327, 1970.
  2. Shizume K: Long term antithyroid drug therapy for intractable cases of Graves' disease. Endocrinol Jpn 25:377, 1978.
  3. Wartofsky L: Low remission after therapy for Graves' disease. Possible relation of dietary iodine with antithyroid therapy results. J Amer Med Assn 226:1083, 1973.
  4. Hedley AJ, Young RE, Jones SJ, Alexander WD, Bewsher PD: Antithyroid drugs in the treatment of hyperthyroidism of Graves' disease: Long-term follow-up of 434 patients. Clin Endocrinol 31:209-218, 1989.
  5. Alexander WD, McG Harden R, Koutras DA, Wayne E: Influence of iodine intake after treatment with antithyroid drugs. Lancet 2:866, 1965.
  6. McMurray JF Jr, Gilliland PF, Ratliff CR, Bourland PD: Pharmacodynamics of propylthiouracil in normal and hyperthyroid subjects after a single oral dose. J Clin Endocrinol Metab 41:362, 1975.
  7. Cooper DS: Which antithyroid drug? Amer J Med 80:1165- 1168, 1986.
  8. Cooper DS, Goldminz D, Levin AA, Ladenson PW, Daniels GH, Molitch ME, Ridgway EC: Agranulocytosis associated with antithyroid drugs. Ann Int Med 98:26-29, 1983.
  9. Barnes V, Bledsoe T: A simple test for selecting the thioamide schedule in thyrotoxicosis. J Clin Endocrinol Metab 35:250, 1972.
  10. Cooper DS, Saxe VC, Meskell M, Maloof F, Ridgway EC: Acute effects of propylthiouracil (PTU) on thyroidal iodide organification and peripheral iodothyronine deiodination: Correlation with serum PTU levels measured by radioimmunoassay. J Clin Endocrinol Metab 54:101, 1982.
  11. Nabil N, Miner DJ, Amatruda JM: Methimazole: An alternative route of administration. J Clin Endocrinol Metab 54:180, 1982.

121.1. Zweig SB, Schlosser JR, Thomas SA, Levy CJ, Fleckman AM .Rectal administration of propylthiouracil in suppository form in patients with thyrotoxicosis and critical illness: case report and review of literature. Endocr Pract. 2006 Jan-Feb;12(1):43-47.

121.2Jongjaroenprasert W, Akarawut W, Chantasart D, Chailurkit L, Rajatanavin R.  Rectal administration of propylthiouracil in hyperthyroid patients:  comparison of suspension enema and suppository form.  Thyroid 12:627-631, 2002

  1. Greer MA, Meihoff WC, Studer H: Treatment of hyperthyroidism with a single daily dose of propylthiouracil. N Engl J Med 272:887, 1965.
  2. Greer MA, Kammer H, Bouma DJ: Short-term antithyroid drug therapy for the thyrotoxicosis of Graves' disease. N Engl J Med 297:173, 1977.
  3. Allannic H, Fauchet R, Orgiazzi J, Madec AM, Genetet B, Lorcy Y, Le Guerrier AM, Delambre C, Derennes V: Antithyroid drugs and Graves' disease: A prospective randomized evaluation of the efficacy of treatment duration. J Clin Endocrinol Metab 70:675, 1990.
  4. Romaldini JH, Bromberg N, Werner RS, Tanaka LM, Rodrigues HF, Werner MC, Farah CS, Reiss LCF: Comparison of high and low dosage regimens of antithyroid drugs. J Clin Endocrinol Metab 57:563, 1983.
  5. Yamamoto M, Totsuka Y, Kojima I, Yamashita N, Togawa K, Sawaki N, Ogata E: Outcome of patients with Graves' disease after long-term medical treatment guided by triiodothyronine (T3) suppression test. Clin Endocrinol 19:467-476, 1983.

126.1 Azizi F, Ataie L, Hedayati M, Mehrabi Y, Sheikholeslami F.Effect of long-term continuous methimazole treatment of hyperthyroidism: comparison with radioiodine.Eur J Endocrinol. 2005 May;152(5):695-701

126.2. Jodar E, Munoz-Torres M, Escobar-Jimenez F, Quesada M, Luna JD, Olea N. Antiresorptive therapy in hyperthyroid patients: Longitudinal changes in bone and mineral metabolism. J Clin Endocrinol Metab 82:1989-1994, 1997.

126.3 Majima T, Komatsu Y, Doi K, Takagi C, Shigemoto M, Fukao A, Morimoto T, Corners J, Nakao K.Clinical significance of risedronate for osteoporosis in the initial treatment of male patients with Graves' disease.J Bone Miner Metab. 2006;24(2):105-13.

  1. Solomon DH, Beck JC, Vanderlaan WP, Astwood EB: Prognosis of hyper-thyroidism treated by antithyroid drugs. J Amer Med Assn 152:201, 1953.
  2. McLarty DG, Alexander WD, McHarden R, Robertson JWK: Self-limiting episodes of recurrent thyrotoxicosis. Lancet 1:6, 1971.
  3. Thalassinos NC, Fraser TR: Effect of potassium iodide on relapse rate of thyrotoxicosis treated with antithyroid drugs. Lancet 2:183, 1971.
  4. Hashizume K, Ichikawa K, Sakurai A, Suzuki S, Takeda T, Kobayashi M, Miyamoto T, Arai M, Nagasawa T: Administration of thyroxine in treated Graves' disease. Effects on the level of antibodies to thyroid-stimulating hormone receptors and on the risk of recurrence of hyperthyroidism. N Engl J Med 324:947-953, 1991.
  5. Hashizume K, Ichikawa K, Nishii Y, Kobayashi M, Sakurai A, Miyamoto T, Suzuki S, Takeda T. Effect of administration of thyroxine on the risk of postpartum recurrence of hyperthyroid Graves' disease. J Clin Endocrinol Metab 75:6-10, 1992.

131.1. Rittmaster RS, Zwicker H, Abbott EC, Douglas R, Givner ML, Gupta MK, Lehmann L, Reddy S, Salisbury SR, Shlossberg AH, Tan MH, York SE. Effect of methimazole with or without exogenous L-thyroxine on serum concentrations of thyrotropin receptor antibodies in patients with Graves’ disease. J Clin Endocrinol Metab 81:3283-3288, 1996.

131.2. Lucas A, Salinas I, Rius F, Pizarro E, Granada ML, Foz M, Sanmarti A. Medical therapy of Graves’ disease: does thyroxine prevent recurrence of hyperthyroidism? J Clin Endocrinol Metab 82:2410-2413, 1997.

  1. Zakarija M, McKenzie JM, Banovac K: Clinical significance of assay of thyroid-stimulating antibody in Graves' disease. Ann Intern Med 93:28, 1980.
  2. Werner RS, Romaldini JH, Farah CS, Werner MC, Bromberg N. Serum thyroid-stimulating antibody, thyroglobulin levels, and thyroid suppressibility measurement as predictors of the outcome of combined methimazole and triiodothyronine therapy in Graves' disease. Thyroid 1:293, 1991.
  3. Schleusener H, Schwander J, Fischer C, Holle R, Holl G, Badenhoop K, Hensen J, Finke R, Bogner U, Mayr WR, Schernthaner G, Schatz H, Pickardt CR, Kotulla P: Prospective multicenter study on the prediction of relapse after antithyroid drug treatment in patients with Graves' disease. Acta Endocrinologica (Copenh) 120:689-701, 1989.
  4. Farid NR (ed): HLA in Endocrine and Metabolic Disorders. New York, Academic Press, 1981, p 357.
  5. Allannic H, Fauchet R, Lorcy Y, Gueguen M, Le Guerrier A- M, Genetet B: A prospective study of the relationship between relapse of hyperthyroid Graves' disease after antithyroid drugs and HLA haplotype. J Clin Endocrinol Metab 57:719, 1983.

136.1. Michelangeli V, Poon C, Taft J, Newnham H, Topliss D, Colman P. The prognostic value of thyrotropin receptor antibody measurement in the early stages of treatment of Graves’ disease with antithyroid drugs. Thyroid 8:119, 1998.

  1. Irvine WJ, Gray RS, Toft AD, Lidgard FP, Seth J, Cameron EHD: Spectrum of thyroid function in patients remaining in remission after antithyroid drug therapy for thyrotoxicosis. Lancet 1:179, 1977.
  2. Buerklin EM, Schimmel M, Utiger RD: Pituitary-thyroid regulation in euthyroid patients with Graves' disease previously treated with antithyroid drugs. J Clin Endocrinol Metab 43:419, 1976.
  3. Lamberg BA, Salmi J, Wagar G, Makinen T: Spontaneous hypothyroidism after antithyroid treatment of hyperthyroid Graves' disease. J Endocrinol Invest 4:399, 1981.
  4. Hirota Y, Tamai H, Hayashi Y, Matsubayashi S, Matsuzuka F, Kuma K, Kumagai LF, Nagataki S: Thyroid function and histology in forty-five patients with hyperthyroid Graves' disease in clinical remission more than ten years after thionamide drug treatment. J Clin Endocrinol Metab 62:165, 1986.
  5. Kampmann JP, Johansen K, Hansen JM, Helwig J: Propylthiouracil in human milk: Revision of a dogma. Lancet 1:736, 1980.

141a. Azizi F, Khoshniat M, Bahrainian M, Hedayati M. Thyroid function and intellectual development of infants nursed by mothers taking methimazole. J Clin Endocrinol Metab 85:3233-3238, 2000.

  1. Wiberg JJ, Nuttall FQ: Methimazole toxicity from high doses. Ann Intern Med 77:414, 1972.
  2. Chevalley J, McGavack TH, Kenigsberg S, Pearson S: A four- year study of the treatment of hyperthyroidism with methimazole. J Clin Endocrinol Metab 14:948, 1954.
  3. Tajiri J, Noguchi S, Murakami T, Murakami N. Antithyroid drug-induced agranulocytosis. The usefulness of routine white blood cell count monitoring. Arch Intern Med 150:621-624, 1990.
  4. Tamai H, Takaichi Y, Morita T, Komaki G, Matsubayashi S, Kuma K, Walter Jr RM, Kumagai LF, Nagataki S: Methimazole- induced agranulocytosis in Japanese patients with Graves' disease. Clin Endocrinol 30:525-530, 1989.
  5. Amrhein JA, Kenny F, Ross D: Granulocytopenia, lupus-like syndrome, and other complications of propylthiouracil therapy. J Pediatr 76:54, 1970.
  6. Pacini F, Sridama V, Refetoff S: Multiple complications of propylthiouracil treatment: Granulocytopenia, eosinophilia, skin reaction, and hepatitis with lymphocyte sensitization. J Endocrinol Invest 5:403-407, 1982.

147.1 Darben T, Savige J, Prentice R, Paspaliaris B, Chick J. 1999 Pyoderma gangrenosum with secondary pyarthrosis following propylthiouracil. Australasian J Dermatol. 40:144-146.

  1. Wall JR, Fang SL, Kuroki T, Ingbar SH, Braverman LE: In vitro immunoreactivity to propylthiouracil, methimazole, and carbimazole in patients with Graves' disease: A possible cause of antithyroid drug-induced agranulocytosis. J Clin Endocrinol Metab 58:868-872, 1984.
  2. Biswas N, Ahn Y-H, Goldman JM, Schwartz JM: Case report: Aplastic anemia associated with antithyroid drugs. Am J Med Sci 301:190-194, 1991.
  3. Escobar-Morreale HF, Bravo P, Garcia-Robles R, Garcia-Larana J, de la Calle H, Sancho JM. Methimazole-induced severe aplastic anemia: unsuccessful treatment with recombinant human granulocyte-monocyte colony-stimulating factor. Thyroid 7:67-70, 1997
  4. Tamai H, Mukuta T, Matsubayashi S, Fukata S, Komaki G, Kuma K, Kumagai LF, Nagataki S. Treatment of methimazole-induced agranulocytosis using recombinant human granulocyte colony-stimulating factor (rhG-CSF). J Clin Endocrinol Metab 77:1356-1360, 1993

151.1 Gunton JE, Stiel J, Caterson RJ, McElduff A. Antithyroid drugs and antineutrophil cytoplasmic antibody positive vasculitis. A case report and review of the literature. J Clin Endocrinol Metab 84:13-16, 1999

151.2 Guma M, Salinas I, Reverter JL, Roca J, Valls-Roc M, Juan M, Olive A.  Frequency of antineutrophil cytoplasmic antibody in Graves’ disease patients treated with methimazole.  J Clin Endocrinol Metab 88:2141-2146, 2003.

151.3.  Breier DV, Rendo P, Gonzalez  J, Shilton G, Stivel M, Goldztein S.  Massive plasmocytosis due to methimazole-induced bone marrow toxicity.  Amer J Hematol 67:259-261, 2001.

151.4. Cooper DS: Antithyroid drugs. N Engl J Med 311:1353- 1362, 1984.

  1. Weiss M, Hassin D, Bank H: Propylthiouracil-induced hepatic damage. Arch Intern Med 140:1184-1185, 1980.

152.1. Williams KV, Nayak S, Becker D, Reyes J, Burmeister LA. Fifty years of experience with propylthiouracil-associated hepatotoxicity: What have we learned? J Clin Endocrinol Metab 82:1727-1733, 1997.

  1. Miyazono K, Okazaki T, Uchida S, Totsuka Y, Matsumoto T, Ogata E, Terakawa K, Kurihara N, Takeda T: Propylthiouracil-induced diffuse interstitial pneumonitis. Arch Intern Med 144:1764-1765, 1984.

.

153.1 Karlsson, FA; Axelsson, O; Melhus, H. Severe embryopathy and exposure to methimazole in early pregnancy. J Clin Endocrinol Metab 87 947-948 2002.

  1. Sato S1, Noh JY, Sato S, Suzuki M, Yasuda S, Matsumoto M, Kunii Y, Mukasa K, Sugino K, Ito K, Nagataki S, Taniyama M. Comparison of efficacy and adverse effects between methimazole 15 mg+inorganic iodine 38 mg/day and methimazole 30 mg/day as initial therapy for Graves' disease patients with moderate to severe hyperthyroidism. Thyroid. 2015 Jan;25(1):43-50. doi: 10.1089/thy.2014.0084.
  2. Godley AF, Stanbury JB: Preliminary experience in the treatment of hyperthyroidism with potassium perchlorate. J Clin Endocrinol Metab 14:70, 1954.
  3. Krevans JR, Asper SP Jr, Rienhoff WF Jr: Fatal aplastic anemia following use of potassium perchlorate in thyrotoxicosis. J Amer Med Assn 181:162, 1962.
  4. Georges JL, Normand JP, Lenormand ME, Schwob J. Life-threatening thyrotoxicosis induced by amiodarone in patients with benign heart disease. European Heart Journal 13:129-132, 1992.
  5. Lazarus JH, Addison GM, Richards AR, Owen GM: Treatment of thyrotoxicosis with lithium carbonate. Lancet 2:1160, 1974.
  6. Turner JG, Brownlie BEW, Rogers TGH: Lithium as an adjunct to radioiodine therapy for thyrotoxicosis. Lancet 1:614, 1976.

159.1. Mercado M, Mendoza-Zubieta V, Bautista-Osorio R, Espinoza-De Los Monteros AL. Treatment of hyperthyroidism with a combination of methimazole and cholestyramine. J Clin Endocrinol Metab 81:3191-3193, 1996.

  1. Reinwein D, Klein E: Der Einfluss des anorganischen Blutjodes auf den Jodumstaz der menschlichen Schilddruse. Acta Endocrinol 35:485, 1960.
  2. Thompson WO, Thorp EG, Thompson PK, Cohen AC: The range of effective iodine dosage in exophthalmic goiter. II. The effect on basal metabolism of the daily administration of one-half drop of compound solution of iodine. Arch Intern Med 45:420, 1930.
  3. Friend DG: Iodide therapy and the importance of quantitating the dose. N Engl J Med 263:1358, 1960.
  4. Emerson CH, Anderson AJ, Howard WJ, Utiger RD: Serum thyroxine and triiodothyronine concentrations during iodide treatment of hyperthyroidism. J Clin Endocrinol Metab 40:33, 1975.
  5. Okamura K1, Sato K, Fujikawa M, Bandai S, Ikenoue H, Kitazono T. Remission after potassium iodide therapy in patients with Graves' hyperthyroidism exhibiting thionamide-associated side effects. J Clin Endocrinol Metab. 2014 Nov;99(11):3995-4002. doi: 10.1210/jc.2013-4466.

165.. Uchida T1, Goto H, Kasai T, Komiya K, Takeno K, Abe H, Shigihara N, Sato J, Honda A, Mita T, Kanazawa A, Fujitani Y, Watada H. Therapeutic effectiveness of potassium iodine in drug-naïve patients with Graves' disease: a single-center experience. Endocrine. 2014 Nov;47(2):506-11. doi: 10.1007/s12020-014-0171-8.

165.1. Yoshihara A1, Noh JY1, Watanabe N1, Mukasa K1, Ohye H1, Suzuki M1, Matsumoto M1, Kunii Y1, Suzuki N1, Kameda T1, Iwaku K1, Kobayashi S1, Sugino K1, Ito K1. Substituting Potassium Iodide for Methimazole as the Treatment for Graves' Disease During the First Trimester May Reduce the Incidence of Congenital Anomalies: A Retrospective Study at a Single Medical Institution in Japan. Thyroid. 2015 Oct;25(10):1155-61. doi: 10.1089/thy.2014.058

166. Shanks RG, Hadden DR, Lowe DC, McDevitt DB, Montgomery DAD: Controlled trial of propranolol in thyrotoxicosis. Lancet 2:1969.

  1. Mazzaferri EL, Reynolds JC, Young RL, Thomas CN, Parasi AF: Propranolol as primary therapy for thyrotoxicosis. Arch Intern Med 136:50, 1976.
  2. Wiener L, Stout BD, Cox JW: Influence of beta sympathetic blockade with propranolol on the hemodynamics of hyperthyroidism. Am J Med 46:227, 1969.

.169. Saunders J, Hall SEH, Crowther A, Sonksen PH: The effect of propranolol on thyroid hormones and oxygen consumption in thyrotoxicosis. Clin Endocrinol 9:67, 1978.

  1. Georges LP, Santangelo RP, Mackin JF, Canary JJ: Metabolic effects of propranolol in thyrotoxicosis. I. Nitrogen, calcium, and hydroxyproline. Metabolism 24:11, 1975.
  2. Hadden DR, Montgomery DAD, Shanks RG, Weaver JA: Propranolol and iodine-131 in the management of thyrotoxicosis. Lancet 2:852, 1968.
  3. Pimstone N, Marine N, Pimstone B: Beta-adrenergic blockade in thyrotoxic myopathy. Lancet 2:1219, 1968.
  4. Smith CS, Howard NJ: Propranolol in treatment of neonatal thyrotoxicosis. J Pediatr 83:1046, 1973.
  5. Rosenberg I: Thyroid storm. N Engl J Med 283:1052, 1970.

174.1. Fraser T, Green D.  Weathering the storm:  beta-blockade and the potential for disaster in severe hyperthyroidism.  Emergency Med 13:376-380, 2001

174.2. Ikram H.  Haemodynamic effects of beta-adrenergic blockade in hyperthyroid patients with and without heart failure.  Br Med J 1:1505-1507, 1977.

  1. Bewsher PD, Pegg CAS, Steward DJ, Lister DA, Michie W: Propranolol in the surgical management of thyrotoxicosis. Ann Surg 180:787, 1974.
  2. Canary JJ, Schaaf M, Duffy BJ, Kyle LH: Effects of oral and intramuscular injection of reserpine in thyrotoxicosis. N Engl J Med 257:435, 1957.
  3. Moncke C: Treatment of thyrotoxicosis with reserpine. Med Monatsschr Pharm 50:1742, 1955.
  4. Lee WY, Bronsky D, Waldenstein SS: Studies of the thyroid and sympathetic nervous system interrelationships: Effect of guanethidine on manifestations of hyperthyroidism. J Clin Endocrinol Metab 22:879, 1962.
  5. deGroot WJ, Leonard JJ, Paley HW, Johnson JE, Warren JV: The importance of autonomic integrity in maintaining the hyperkinetic circulatory dynamics of human hyperthyroidism. J Clin Invest 40:1033, 1961.
  6. Dillion PT, Babe J, Meloni CR, Canary JJ: Reserpine in thyrotoxic crisis. N Engl J Med 283:1020, 1970.
  7. DeGroot LJ, Hoye K: Dexamethasone suppression of serum T3 and T4. J Clin Endocrinol Metab 4:976, 1976.
  8. Williams DE, Chopra IJ, Orgiazzi J, Solomon DH: Acute effects of corticosteroids on thyroid activity in Graves' disease. J Clin Endocrinol Metab 41:354, 1975.
  9. Witztum JL, Jacobs LS, Schonfeld G: Thyroid hormone and thyrotropin levels in patients placed on colestipol hydrochloride. J Clin Endocrinol Metab 46:838-840, 1978.
  10. Boehm TM, Burman KD, Barnes S, Wartofsky L: Lithium and iodine combination therapy for thyrotoxicosis. Acta Endocrinol 94:174, 1980.
  11. Sharp B, Reed AW, Tamagna EI, Gefner DL, Hershman JM: Treatment of hyperthyroidism with sodium ipodate (oragraffin) in addition to propylthiouracil and propranolol. J Clin Endocrinol Metab 53:622, 1981. 186. Croxson MS, Hall TD, Nicoloff JT: Combination drug therapy for treatment of hyperthyroid Graves' disease. J Clin Endocrinol Metab 45:623, 1977.
  12. Werner SC, Platman SR: Remission of hyperthyroidism (Graves' disease) and altered pattern of serum-thyroxine binding induced by prednisone. Lancet 2:752, 1965.
  13. Wu S-Y, Shyh T-P, Chopra IJ, Solomon DH, Huang H-W, Chu P-C: Comparison of sodium ipodate (oragrafin) and propylthiouracil in early treatment of hyperthyroidism. J Clin Endocrinol Metab 54:630, 1982.
  14. Shen D-C, Wu S-Y, Chopra IJ, Huang H-W, Shian L-R, Bian T-Y, Jeng C-Y, Solomon DH: Long term treatment of Graves' hyperthyroidism with sodium ipodate. J Clin Endocrinol Metab 61:723, 1985.189.1. Bal CS, Kumar A, Pandey RM.  A randomized controlled trial to evaluate the adjuvant effect of lithium on radioiodine treatment of hyperthyroidism.  Thyroid 12:399-405, 2002.

189.1 El Fassi D, Banga JP, Gilbert JA, Padoa C, Hegedüs L, Nielsen CH.

Clin Immunol. 2009 Mar;130(3):252-8 Treatment of Graves' disease with rituximab specifically reduces the production of thyroid stimulating autoantibodies.
189.2 Heemstra KA, Toes RE, Sepers J, Pereira AM, Corssmit EP, Huizinga TW, Romijn JA, Smit JW.Eur J Endocrinol. 2008 Nov;159(5):609-15 Rituximab in relapsing Graves' disease, a phase II study.

189.3 El Fassi D, Nielsen CH, Junker P, Hasselbalch HC, Hegedüs L. Systemic adverse events following rituximab therapy in patients with Graves' disease.J Endocrinol Invest. 2011 Jul-Aug;34(7):e163-7.

189.4 Neumann S, Eliseeva E, McCoy JG, Napolitano G, Giuliani C, Monaco F, Huang W, Gershengorn MCA new small-molecule antagonist inhibits Graves' disease antibody activation of the TSH receptor.J Clin Endocrinol Metab. 2011 Feb;96(2):548-54. Epub 2010 Dec 1.

  1. Klementschitsch P, Shen K-L, Kaplan EL: Reemergence of thyroidectomy as treatment for Graves' disease. Surg Clin North Am 59:35, 1979.
  2. Palit TK, Miller CC, Miltenburg DM. The efficacy of thyroidectomy for Graves' disease: A meta-analysis. J Surg Res 90:161-165, 2000.
  3. Winsa B, Rastad J, Akerstrom G, Johansson H, Westermark K, Karlsson FA. Retrospective evaluation of the effect of subtotal and total thyroidectomy in the treatment of Graves' disease with and without endocrine ophthalmopathy. Thesis, Brita Winsa, The University of Upsala, Upsala, Sweden, 1993.

193. Sundaresh V1, Brito JP, Wang Z, Prokop LJ, Stan MN, Murad MH, Bahn RS. Comparative effectiveness of therapies for Graves' hyperthyroidism: a systematic review and network meta-analysis. J Clin Endocrinol Metab. 2013 Sep;98(9):3671-7. doi: 10.1210/jc.2013-1954.

  1. Hamilton RD, Mayberry WE, McConahey WM, Hanson KC: Ophthalmopathy of Graves' disease: A comparison between patients treated surgically and patients treated with radioiodide. Mayo Clin Proc 42:812, 1967.
  2. Stocker DJ, Foster SS, Solomon BL, Shriver CD, Burch HB.  Thyroid cancer yield in patients with Graves’ disease selected for surgery on the basis of cold scintiscan defects.  Thyroid 12:305, 2002.
  3. Bogazzi F, Miccoli P, Berti P, Cosci C, Brogioni S, Aghini-Lombardi F, Materazzi G, Bartalena L, Pinchera A, Braverman LE, Martino E.  Preparation with iopanoic acid rapidly controls thyrotoxicosis in patients with amiodarone-induced thyrotoxicosis before thyroidectomy.  Surgery 132:1114-1117, 2002.

197:  Scholz GH, Hagemann E, Arkenau C, Engelmann L, Lamesch P, Schreiter D,Schoenfelder M, Olthoff D, Paschke R.  Is there a place for thyroidectomy in older patients with thyrotoxic storm and cardiorespiratory failure?Thyroid. 2003 Oct;13(10):933-40.

198 Erbil Y, Ozluk Y, Giriş M, Salmaslioglu A, Issever H, Barbaros U, Kapran Y, Ozarmağan S, Tezelman S. Effect of lugol solution on thyroid gland blood flow and microvessel density in the patients with Graves' disease.J Clin Endocrinol Metab. 2007 Jun;92(6):2182-9

  1. Feek CM, Stewart J, Sawers A, Irvine WJ, Beckett GJ, Ratcliffe WA, Toft AD: Combination of potassium iodide and propranolol in preparation of patients with Graves' disease for thyroid surgery. N Engl J Med 302:883, 1980.
  2. Bewsher BD, Pegg CAS, Stewart DJ, Lister DA, Michie W: Propranolol in the surgical management of thyrotoxicosis. Ann Surg 180:787, 1974.
  3. Toft AD, Irvine WJ, Campbell RWF: Assessment by continuous cardiac monitoring of minimum duration of preoperative propranolol treatment in thyrotoxic patients. Clin Endocrinol 5:195, 1976.
  4. Kulkarni RS, Braverman LE, Patwardhan NA. Bilateral cervical plexus block for thyroidectomy and parathyroidectomy in healthy and high risk patients. J Endocrinol Invest 19:714-718, 1996.
  5. Taylor GW, Painter NS: Size of the thyroid remnant in partial thyroidectomy for toxic goiter. Lancet 1:287, 1962.
  6. Sugino K, Mimura T, Toshima K, Iwabuchi H, Kitamura Y, Kawano M, Ozaki O, Ito K. Follow-up evaluation of patients with Graves' disease treated by subtotal thyroidectomy and risk factor analysis for post-operative thyroid dysfunction. J Endocrinol Invest 16:195-199, 1993.
  7. Ozaki O, Ito K, Mimura T, Sugino K, Ito K. Factors affecting thyroid function after subtotal thyroidectomy for Graves’ disease: Case control study by remnant-weight matched-pair analysis. Thyroid 7:555, 1997.
  8. Miccoli P, Vitti P, Rago T, Iacconi P, Bartalena L, Bogazzi F, Fiore E, Valeriano R, Chiovato L, Rocchi R, Pinchera A. Surgical treatment of Graves’ disease: Subtotal or total thyroidectomy? Surgery 120:1020-1025, 1996
  9. Menconi F, Marinò M, Pinchera A, Rocchi R, Mazzi B, Nardi M, Bartalena L, Marcocci C.Effects of total thyroid ablation versus near-total thyroidectomy alone on mild to moderate Graves' orbitopathy treated with intravenous glucocorticoids. J Clin Endocrinol Metab. 2007 May;92(5):1653-8.
  10. Wilhelm SM, McHenry CR.World J Surg. 2009 Dec 23. [Total Thyroidectomy Is Superior to Subtotal Thyroidectomy for Management of Graves’ Disease in the United States.
  11. Sawyers JL, Martin CE, Byrd BF Jr, Rosenfield L: Thyroidectomy for hyperthyroidism. Ann Surg 175:939, 1972.
  12. Farnell MB, van Heerden JA, McConahey WM, Carpenter HA, Wolff LH Jr: Hypothyroidism after thyroidectomy for Graves’ disease. Amer J Surg 142:535, 1981

211 Buchanan WW, Koutras DA, Crooks J, Alexander WD, Brass W, Anderson JR, Goudie RB, Gray KG: The clinical significance of the complement-fixation test in thyrotoxicosis. J Endocrinol 24:115, 1962.

  1. Yamashita H, Noguchi S, Tahara K, Watanabe S, Uchino S, Kawamoto H, Toda M, Murakami N. Postoperative tetany in patients with Graves’ disease: A risk factor analysis. Clin Endocrinol 47:71-77, 1997.
  2. Michie W, Duncan T, Hamer-Hodges DW, Bewsher PD, Stowers JM, Pegg CAS, Hems G, Hedley AJ: Mechanism of hypocalcemia after thyroidectomy for thyrotoxicosis. Lancet 1:508, 1971.
  3. Hardisty CA, Talbot CH, Munro DS: The effect of partial thyroidectomy for Graves' disease on serum long-acting thyroid stimulator protector (LATS-P). Clin Endocrinol 14:181, 1981.
  4. Bech K, Feldt-Rasmussen U, Bliddal H, Date J, Blichert-Toft M: The acute changes in thyroid stimulating immunoglobulins, thyroglobulin, and thyroglobulin antibodies following subtotal thyroidectomy. Clin Endocrinol 16:235, 1982.
  5. Werga-Kjellman P, Zedenius J, Tallstedt L, Traisk F, Lundell G, Wallin G.  Surgical treatment of hyperthyroidism:  A ten year experience.  Thyroid 11:187-192, 2001.
  6. Fukino O, Tamai H, Fujii S, Ohsako N, Matsubayashi S, Kuma K, Nagataki S: A study of thyroid function after subtotal thyroidectomy for Graves' disease: particularly on TRH tests, T3 suppression tests and antithyroid antibodies in euthyroid patients. Acta Endocrinol 103:28-33, 1983.
  7. Hedley AJ, Hall R, Amos J, Michie W, Crooks J: Serum-thyrotropin levels after subtotal thyroidectomy for Graves' disease. Lancet 1:455, 1971.

219 Segni M, Leonardi E, Mazzoncini B, Pucarelli I, Pasquino AM. 1999 Special features of Graves' disease in early childhood. Thyroid 9:871.

  1. Perrild H, Jacobsen BB. Thyrotoxicosis in childhood. Europ J Endocrinol 134:678-679, 1996
  2. Starr P, Jaffe HL, Oettinger L Jr: Late results of 131-I treatment of hyperthyroidism in seventy-three children and adosescents. J Nucl Med 5:81, 1964.
  3. Kogut MD, Kaplan SA, Collipp PJ, Tiamsic T, Boyle D: Treatment of hyperthyroidism in children. N Engl J Med 272:217, 1965.

223:  Read CH Jr, Tansey MJ, Menda Y. A 36-year retrospective analysis of the efficacy and safety of radioactive iodine in treating young Graves' patients. J Clin Endocrinol Metab. 2004 Sep;89(9):4229-33

  1. Rivkees SA, Sklar C, Freemark M. The management of Graves’ disease in children, with special emphasis on radioiodine treatment. J Clin Endocrinol Metab 83:3767-3776.

224.1 Rivkees SA, Controversies in the management of Graves’ disease in children. J Endocrinol Invest 2016 Nov;39(11):1247-1257

  1. Barrio R, Lopez-Capape M, Martinez-Badas I, Carrillo A, Moreno JC, Alonso M.Graves' disease in children and adolescents: response to long-term treatment.Acta Paediatr. 2005 Nov;94(11):1583-9

226. Léger J, Gelwane G, Kaguelidou F, Benmerad M, Alberti C; French Childhood Graves' Disease Study Group Positive impact of long-term antithyroid drug treatment on the outcome of children with Graves' disease: national long-term cohort study.J Clin Endocrinol Metab. 2012 Jan;97(1):110-9.

 

  1. Jevalikar G, Solis J, Zacharin M. Long-term outcomes of pediatric Graves' disease. J Pediatr Endocrinol Metab. 2014 Nov;27(11-12):1131-6. doi: 10.1515/jpem-2013-0342.
  2. Rudberg C, Johansson H, Akerstrom G, Tuvemo T, Karlsson FA. Graves’ disease in children and adolescents. Late results of surgical treatment. Europ J Endocrinol 134:710-715, 1996.
  3. Soreide JA, van Heerden JA, Lo CY, Grant CS, Zimmerman D, Ilstrup DM. Surgical treatment of Graves’ disease in patients younger than 18 years. World J Surg 20:794-800, 1996

230 Sherman J, Thompson GB, Lteif A, Schwenk WF 2nd, van Heerden J, Farley DR, Kumar S, Zimmerman D, Churchward M, Grant CS.Surgical management of Graves disease in childhood and adolescence: an institutional experience.Surgery. 2006 Dec;140(6):1056-61

  1. Okuno A, Yano K, Inyaku F, Suzuki Y, Sanae N, Kumai M, Naitoh Y. Pharmacokinetics of methimazole in children and adolescents with Graves’ disease. Acta Endocrinol (Copenh) 115:112-118, 1987.
  2. Léger J, Gelwane G, Kaguelidou F, Benmerad M, Alberti C; French Childhood Graves’ Disease Study Group Positive impact of long-term antithyroid drug treatment on the outcome of children with Graves’ disease: national long-term cohort study.J Clin Endocrinol Metab. 2012 Jan;97(1):110-9.

233 Luton D, Le Gac I, Vuillard E, Castanet M, Guibourdenche J, Noel M, Toubert ME, Leger J, Boissinot C, Schlageter MH, Garel C, Tebeka B, Oury JF, Czernichow P, Polak M.Management of Graves' disease during pregnancy: the key role of fetal thyroid gland monitoring.J Clin Endocrinol Metab. 2005 Nov;90(11):6093-8

  1. Clementi M, Di Gianantonio E, Pelo E, Mammi I, Basile RT, Tenconi R. 1999 Methimazole embryopathy:   delineation of the phenotype. Amer J Medical Genet. 83:43-46.
  2. Cheek JH, Rezvani I, Goodner D, Hopper B: Prenatal treatment of thyrotoxicosis to prevent intrauterine growth retardation. Obstet Gynecol 60:122, 1982.
  3. Zimmerman D. 1999 Fetal and neonatal hyperthyroidism. Thyroid 9:727.
  4. Hollingsworth DR, Mabry CC, Eckerd JM: Hereditary aspects of Graves' disease in infancy and childhood. J Pediatr 81:446, 1972.

 

 

 

 

Effects of the Environment, Chemicals and Drugs on Thyroid Function

ABSTRACT

The sensitive and tightly regulated feedback control system, thyroid gland autoregulation, and the large intrathyroidal and extrathyroidal storage pools of thyroid hormone serve to provide a constant supply of thyroid hormone to peripheral tissues in the face of perturbations imposed by the external environment, chemicals and drugs, and a variety of diseases processes. The thyroid is subject to a great number of exogenous and endogenous perturbations. The same agent may produce alterations in various aspects of thyroid hormone economy. For this reason, it is difficult to precisely classify all external and internal influences according to their mode of action. This chapter reviews effects on the thyroid produced by changes in the external environment, chemicals and drugs. The effects of non-thyroidal illness are reviewed in Chapter 5b. The effects of the more important factors and chemical agents and drugs are discussed individually.

 

RESPONSES TO ALTERATIONS IN THE EXTERNAL ENVIRONMENT

Environmental Temperature

Changes in environmental temperature may cause alterations in TSH secretion and in the serum concentration of thyroid hormones and their metabolism. The changes are probably mediated through the hypothalamus and the pituitary and by peripheral effects on the pathways and rates of thyroid hormone degradation and fecal losses and alterations in thyroid hormone action. The in vitro effects of temperature on the firmness of binding of T4 to its transport serum proteins conceivably also play a role in vivo.1  The overall effects of environmental temperature have been more obvious and easier to demonstrate in animals than in humans but differences in thermal regulation 1a may mean that findings in animal models may not apply to humans. Additionally, studies of individuals with prolonged residence in Arctic and Antarctic regions may be confounded by other alterations in daylight, activity levels, living conditions and sleep deprivation. 1b,1c

Effects of Cold

Dramatic, although transient, increases in serum TSH levels have been observed in infants and young children during surgical hypothermia.2  Also, a prompt and important secretion of TSH occurs in the newborn, in the first few hours after birth, accompanied by an increase in thyroid hormone secretion and clearance.3,4  Since this TSH surge is partially prevented by maintaining infants in a warm environment, postnatal cooling appears to be responsible in part for the rise in TSH secretion. In most studies, exposure of adults to cold or even intensive hypothermia has produced no changes,5,6 or at best minimal increases7 in serum TSH. More prolonged exposure to cold generally results in maintenance of the total T4 (TT4) and free T4 (fT4) levels with maintenance of a normal or decreased total T3 (TT3) and free T3 (fT3) levels. 7a,7b , however, others have shown prolonged arctic residence leads the increase in TSH to be associates with an increase in, thyroglobulin and T3..7c These alterations may be partly the consequence of a direct effect of temperature on the rate and pathways of thyroid hormone metabolism with more rapid production and clearance of T3. Altered kinetics have been demonstrated in humans 7d, but have been more thoroughly studied in animals.8,9,9a,9b    It has been more difficult to show a clear seasonal variation in serum hormone concentration. However, the variation demonstrated in several studies10,11 has been that T4 and T3 values are higher during the colder months.

Cold exposure in animals leads to thyroid gland hyperplasia, enhanced hormonal secretion, degradation, and excretion, accompanied by an increased demand for dietary iodine. All of these effects are presumably due to an increased need for thyroid hormone by peripheral tissues. The prompt activation of pituitary TSH secretion after cold exposure of the rats12,13 is possibly due in part to a direct effect on the hypothalamus.14  Exposure to cold has also resulted in augmented TRH production, and serum levels,16 and blunted responses of TSH to exogenous TRH.17  These effects have not been reproduced by other laboratories13,18 although an increase in thyroid hormone secretion has been clearly demonstrated.6,19,20  In the rat, it is associated with augmented rates of T4 and T3 deiodination, increased conversion of T4 to T3, and enhanced hepatic binding and biliary and fecal clearance of the iodothyronines.8,9,9a,21,22  Finally, thyroid hormone effects may be enhanced by alterations in co-activators which enhance the activity of thyroid hormone receptors on gene activation. 22a

Effects of Heat

In general, an increase in ambient temperature has produced effects opposite to those observed during cold exposure, although the effects of heat have not been extensively investigated. As indicated above, thyroid hormone levels in serum tend to be lower during the summer months. A decrease in the serum T3 concentration, with reciprocal changes in the levels of rT3, have been observed in normal subjects acutely exposed to heat and during febrile illnesses.23,24  In the latter condition, the contribution of the rise in body temperature relative to other effects of systemic illness cannot be dissociated. A decrease in the elevated serum TSH level associated with primary hypothyroidism has been induced by increases in body temperature.25

High Altitude and Anoxia

Acute elevations in serum T4 and T3 concentrations occur in humans during the early period of exposure to high altitude.26  Increases in the rate of T4 degradation and thyroidal RAIU have also been reported.27,28 At very high elevations (5400-6300 m), elevations in T4, fT4, T3, and TSH with a normal fT3 have been reported.28a When compared to those residing at sea level, individuals adapted to altitude were noted to have a lower T4 with higher fT4 and fT3 levels and a normal TSH response to TRH.28b Moderate, transient increases in oxygen consumption, not a result of sympathetic activation, were found in one study.28

The responses of rats exposed to high altitude or anoxia seem to be quite different. Thyroidal iodinative activity and T4 formation are diminished.29-31  The partial reversal of these changes by the administration of TSH led the authors of these studies to conclude that the primary effect is probably diminished TSH secretion.

Alterations in Light

Pinealectomy induces a moderate increase in thyroid weight,32 and continuous light exposure33 increases the T4 secretion rate of rats by about 20%. In squirrels, continuous darkness produces a decrease in thyroid weight and T4 levels33a, but this effect is blocked by pinealectomy.33a These studies suggest that melatonin has an inhibitory effect on thyroid gland function.33a,34 A nocturnal increase in Type II deiodinase activity Is blocked by exposure to continuous light.34a   Although the retinas of rat pups reared in total darkness are totally devoid of TRH, the content of TRH in the hypothalamus remains unaltered.35  The diurnal variation in hypothalamic TRH content, reflecting both rhythmic synthesis and secretion, is, however, blunted in the absence of cyclical light changes. Little is known about the effect of light on the thyroid in humans. The normal TSH rhythm can be reset by a pulse of light.35a

Nutrition

Since thyroid hormone plays a central role in the regulation of total body metabolism, it is not surprising that nutritional factors may profoundly alter the regulation, supply, and disposal of this thermogenic hormone. Although many dietary changes can affect the thyroid economy, the most striking and important effects are related to alterations in total caloric intake and the supply of iodine. The changes associated with caloric deprivation appear homeostatic in nature producing alterations in thyroid hormones which would conserve energy through a reduction in catabolic expenditure. The changes observed with a deficiency or excess of iodine supply generally serve to maintain an adequate synthesis and supply of thyroid hormone, principally through modifications in thyroidal iodide accumulation and binding.

 

Starvation and Fasting

Multiple alterations in thyroid hormone regulation and metabolism have been noted during caloric restriction. The most dramatic effect is a decrease in the serum TT3 within 24-48 hours of the initiation of fasting.36-40b  Because changes in the free T3 fraction are usually small, the absolute concentration of FT3 is also reduced, clearly into the hypothyroid range The marked reduction in serum T3 is caused by a reduction in its generation from T4 rather than by an acceleration in its metabolic clearance rate.41,42 The decline in T3 concentration is accompanied by a concomitant and reciprocal change in the concentration of total and free rT3. The increase in the serum rT3 concentration tends to begin later and to return to normal at the time serum T3 is being maintained at a low level with continuous calorie deprivation.38,39 Little change occurs in the concentrations of TT4 and FT4 and the production and metabolic clearance rates of T4.38,39,41,42  When small changes have been observed, they were generally in the direction of an increase in the FT4 concentration. They are attributed to decreased concentration of the carrier proteins in serum, as well as to their diminished association with the hormone caused by the inhibitory effect of free fatty acids (FFA) the level of which increases during fasting.40,43

Decreased outer ring monodeiodination (5'-deiodinase activity) would explain both the decreased generation of T3 from T4 and the excess accumulation of rT3. This hypothesis seems to be fully supported by in vitro studies using liver tissue from fasted fats.44  It is further supported by the finding of increased generation and serum concentration of 3',5'-T2 and 3'-T1 and decreased 3,5-T2 and 3,3'-T2.44-47  However, a less important increase in the monodeiodination of the inner ring of T4 (5-deiodination)42 explains the temporal dissociation of changes in serum T3 and rT3 concentration. A decrease in plasma T3 after fasting with an increase in hepatic type III deiodinase activity and mRNA has also been noted in chickens. 47a An increase in the nondeiodinative pathway of T4 degradation with the formation of Tetrac has been also reported.48

Considerable controversy remains regarding the mechanisms responsible for the observed changes in the rates of the deiodinative pathways of iodothyronines. Decreased generation of nonprotein sulfhydryls (NP-SH) as a cause of the reduction in 5'-deiodinase activity was suggested on the basis of the observed enhancement in enzyme activity by the in vitro addition of dithiothreitol. Reduced glutathione and NADPH had a similar effect.49  Although Chopra's50 direct measurements of NP-SH in tissue during fasting seemed to confirm this hypothesis, the precise mechanism is likely more complex. Decreased tissue NP-SH content does not always correlate with the inhibition of T3 generation, which may be restored by glucose refeeding independently of changes in NP-SH content.50,51

Composition of the diet rather than reduction in the total calorie intake seems to determine the occurrence of decreased T3 generation in peripheral tissues during food deprivation. The dietary content of carbohydrate appears to be the key ingredient since as little as 50 g glucose reverses toward normal the fast-induced changes in T3 and rT3.52  Replacement of dietary carbohydrate with fat results in changes typical of starvation.39,53  Refeeding of protein may partially improve the rate of T3 generation, but the protein may be acting as a source of glucose through gluconeogenesis.54  Yet, dietary glucose is not the sole agent responsible for all changes in iodothyronine metabolism associated with starvation. For example, the increase in serum rT3 concentration may not be solely dependent on carbohydrate deprivation since a pure protein diet partially restores the level of rT3 but not that of T339 (Fig. 5-1). The composition of the antecedent diet also has an effect on the magnitude of the serum T3 fall during fasting.39,52  It is possible that the cytoplasmic redox state, measured in terms of the lactate/pyruvate ratio rather than glucose itself, regulates the rate of deiodinative pathways of iodothyronines.55

The basal serum TSH level during calorie deprivation is either normal or low, the response to TRH is blunted37-39 and the normal nocturnal rise in TSH is blunted.40a These changes are quite surprising given the consistent and profound decrease in serum FT3 levels. Several hypothesis have been proposed to explain this paradox. Because the pituitary is able to continue to respond appropriately during fasting to both suppressive and stimulatory signals,56 it has been suggested that starvation only "resets" the set point of feedback regulation. A more plausible hypothesis, supported by experimental data,57,58 proposes that the pituitary is regulated by the intracellular concentration of T3, which may remain unaltered through factors ensuring its continuous local generation during starvation, whereas a decrease is typically found in other tissues.   Further support for this hypothesis comes from a recent study demonstrating that fasting produces a marked increase in hypothalamic Type II Deiodinase mRNA58a which would enhance local T3 production.   This hypothesis gives credence to the preservation of a closer inverse relationship between serum FT4 and TSH than between FT3 and TSH. Hypothalamic TRH content in starved rats has been reported to be normal,59 low60  or even elevated.60a The elevation of TRH was accompanied by normal levels of proTRH mRNA and decreased pituitary TSH; it was suggested that this represented decreased TRH release. 60a In a different study of starved rats, the hypothalamic proTRH mRNA and the TRH content were both decreased,60b but these effects were reversed by adrenalectomy suggesting that they were secondary to increased glucocorticoid levels.60b Neonatal starvation in rats leads to diminished TRH and TSH production, with resultant hypothyroidism and growth retardation.61

Starvation produces a greater than 50% decrease in the maximal binding capacity of T3 to rat liver nuclear receptors within 48 hours.62  Although accompanied by a diminution of almost equal magnitude in the nuclear T3 content, it is unlikely that the observed change represents an alteration of the receptor content by the hormone as the more profound diminution of nuclear T3 content associated with hypothyroidism does not produce changes in the maximal binding capacity of T3 in rat liver nuclei. The reduction in maximal binding capacity has been demonstrated to coincide with a reduction in the level of the thyroid hormone receptors.62a The affinity of the rat liver T3 receptor is not affected by starvation.62,63  Studies in humans have used circulating mononuclear cells and, probably due to the limited choice of tissue, results have been either equivocal or negative.64

Other hormonal and metabolic changes during fasting may account for the observed alterations in the regulation and metabolism of thyroid hormones. Among them are the increase in plasma cortisol and suppression of adrenergic stimuli.65  Both changes are known to induce independently a decrease in the serum T3 concentration by inhibition of T4 to T3 conversion in peripheral tissues (see below). Accordingly, they may be partly responsible for the decrease in T3 neogenesis during starvation. There is likely a highly complex interplay between the changes in thyroid hormone and the many metabolic changes of starvation. In addition to a direct effect of glucose, changes in FFA, ketosis, and the redox state may influence thyroid hormone metabolism, while T3 itself may impact hepatic glucose production.40b

Two major issues of theoretical and practical importance remain unresolved - do the observed changes in thyroid function produce some degree of hypothyroidism, and is this state beneficial to the energy-deprived organism? Although the suppressed serum TSH response to TRH suggests that the starving organism does not suffer from a significant deprivation in thyroid hormone, other observations indicate the contrary. The decreased pulse rate, systolic time interval, oxygen consumption, and decrease in activity of some liver enzymes are suggestive of hypothyroidism at the level of peripheral tissues.66  Furthermore, administration of T3 to restore its serum level to normal during fasting increased the production and excretion of urea and 3-methylhistidine.56,67  Larger doses of T3, given during fasting, had even more profound effects. These effects included dramatic increased in the excretion of urea and creatine, and increased plasma levels of ketones and FFA indicating an accelerated protein and fat breakdown.68  Such evidence leaves little doubt that the decrease in T3 generation during calorie deprivation has an energy- and nitrogen-sparing effect. It is tempting to speculate that the result is beneficial in the adaptation to malnutrition through reduction in metabolic expenditure.

Fasting is not only a useful model for studying the effects of calorie deprivation on thyroid hormone but is also the prototype of the "low T3 syndrome".69  The latter is produced by a number of chemical agents and drugs, and accompanies a variety of nonthyroidal illnesses. It is possible that malnutrition, concomitant in a number of acute and chronic illnesses, is in part responsible for some of the observed changes in thyroid physiology.

 

Protein-Calorie Malnutrition (PCM)

As in the case of starvation, PCM is associated with a low serum T3 concentration and increased rT3 levels, probably due to similar changes in iodothyronine monodeiodination. However, important differences exist between the abnormalities in thyroid function observed in PCM and acute calorie deprivation. Most reports indicate important decreases in TBG and TTR concentrations, and there are also indications of hormone binding abnormalities.70,71  As a consequence, the free concentrations of both T4 and T3 are usually normal.70,72,72a  Recovery is associated with restoration of the level of serum thyroid hormones and binding proteins. Despite an accelerated turnover time, the absolute amount of extrathyroidal T4 disposed each day is reduced. Refeeding restores the T4 kinetics to normal.70  The thyroidal RAIU is reduced due to a defect in the iodine-concentrating mechanism.73  The most striking difference between starvation and PCM is the finding the latter of an exaggerated and sustained TSH response to TRH, with basal TSH levels either elevated or normal.70,72,72a,72b,74

The experimental model of protein malnutrition in the rat yielded different results from those observed in humans. Serum T4 and T3 levels were found to be both elevated.75  However, in the lamb, as in humans, chronic malnutrition leads to a lower rate of T4 utilization.76

 

Overfeeding and Obesity

Overfeeding produces an increase in the serum T3 concentration as a result of an increased conversion of T4 to T3. It is particularly marked when the excess calories are given in the form of carbohydrates.77  Thus, it appears that the effect of overnutrition on iodothyronine metabolism is the opposite of that of starvation. This finding gives further credence to the speculation that changes in thyroid hormone may serve to modulate the homeostasis of energy expenditure.

Although it has been reported that serum T3 concentrations correlate with body weight,78 it appears that this phenomenon reflects the effect of an increase in caloric intake on T3 production. Most studies find that obese subjects have normal thyroid function and hormone metabolism.79  Furthermore, no abnormalities in the hypothalamic-pituitary-thyroid axis have been demonstrated in obese subjects.

 

Minerals

Iodine. Of the many minerals that may affect thyroid function, iodine is the most important. It is an essential substrate for thyroid hormone synthesis and also interacts with the function of the thyroid gland at several levels.

Acute administration of increasing doses of iodide enhances total hormone synthesis until a critical level of intrathyroidal iodide is reached. Beyond this level, iodide organification and hormone synthesis are blocked (the acute Wolff-Chaikoff block). Chronic or repeated administration of moderate to large doses of iodine causes a decrease in iodide transport resulting in a decrease in its intrathyroidal concentration. The latter relieves the Wolff-Chaikoff block and is known as the escape or adaptation phenomenon. Although the exact mechanisms of the block and escape remain unknown, they appear to be autoregulatory in nature since they are independent of pituitary TSH secretion. Iodoloactones may play a role in the induction of the Wolff-Chaikoff block.80   One mechanism through which iodide acts is via desensitization of the thyroid gland to TSH. In TSH stimulated glands, iodine rapidly reduces the level of the mRNA for thyroid peroxidase (TPO) and the Na/I symporter (NIS) but not for thyroglobulin (Tg) or the TSH receptor (TSHr).80a Iodine also antagonizes TSH stimulated thyrocyte proliferation.80a   In FRTL-5 cells, iodine blocks the TSH stimulation of Tg synthesis but does not alter the level of the Tg mRNA.80b   These actions occur without a change in TSH receptor number, and may, in part, be via an action on adenylyl cyclase.80c More detailed description is provided in Chapter 2.

Another effect of large doses of iodine, apparently independent of TSH and hormone synthesis, is the prompt inhibition of hormone release. It has been exploited to achieve rapid amelioration of thyrotoxicosis in Graves' disease and toxic nodular goiters (see Chapters 11 and 13). In normal persons, the inhibitory effect of large doses of iodine on thyroid hormone release produces a transient decrease in the serum concentration of T4 and T3. It causes, in turn, a compensatory increase in serum TSH, which stimulates hormone secretion and thus counteracts the effect of iodine.81,82  The mechanisms of thyroidal autoregulation are believed to serve the purpose of accommodating wide and rapid fluctuations in iodine supply.

The most intriguing effects of iodine are the involution of hyperplasia and the decrease in vascularity that occur when the ion is administered to patients with diffuse toxic goiter. Iodine may be able to induce apoptosis in thyroid cells. 82a,82b Under different circumstances, iodide may intensify the hyperplasia and produce a goiter (Chapter 20).

Iodine deficiency used to be the leading cause of goiter in the world and still remains so in certain regions. When severe, it can cause hypothyroidism and cretinism, described in detail in Chapter 20 . In the United States and the rest of the developed world, untoward effects from excess iodine supplementation or the use of iodine-containing compounds are more common than problems related to iodine deficiency.

Excess iodine can be responsible for the development of goiter, hypothyroidism, and thyrotoxicosis. However, it should be emphasized that these complications usually occur in persons with underlying defects of thyroid function who are unable to utilize the normal adaptive mechanisms. Iodide-induced goiter (iodide goiter), without or with hypothyroidism (iodide myxedema), is encountered with greater frequency in patients with Hashimoto's thyroiditis or previously treated Graves' disease.83,84  Other predisposed persons include those who have undergone partial thyroid gland resection, patients with defects of hormonogenesis, and some with cystic fibrosis.85 Drugs such as phenazone,86,87 lithium,88 sulfadiazine,89 and cycloheximide90 may act synergistically with iodide to induce goiter and/or hypothyroidism.

More rarely, ingestion of excess iodide may cause thyrotoxicosis (iodide-induced thyrotoxicosis or Jodbasedow).90a   This was initially observed with the introduction of iodine prophylaxis in areas of endemic iodine deficiency.91,92  It has also been observed after the administration of iodide in excess to patients with nodular thyroid disease residing in areas of moderate iodine deficiency or even iodine sufficiency.93,94  Although the exact mechanism of induction of thyrotoxicosis remains obscure, it may be related to the stimulation of increased thyroid hormone synthesis in areas of the gland with autonomous nodular activity.

Ingestion of excess iodide by a gravid woman may cause an iodide goiter in the fetus, and if the gland is large enough it may result in asphyxia during the postnatal period (Chapter 20). Consumption of Kombu, the iodine-rich seaweed, is responsible for the occurrence of endemic goiter in the Japanese island of Hokkaido.95  It has also been suggested that the increase in dietary iodine content in the United States during the last three decades is responsible for the higher recurrence rate of thyrotoxicosis in patients previously treated with antithyroid drugs.96

 

Calcium. Calcium is said to be goitrogenic when in the diet in excess. Administration of 2 g calcium per day was associated with decreased iodide clearance by the thyroid.97  The action is unknown, but it may in some way make overt a borderline dietary iodine deficiency. Calcium also acutely and chronically reduces the absorption of thyroxine. 97a, 97b

 

 

Nitrate. Nitrate in the diet (0.3 - 0.9%) can interfere with 131I uptake in the thyroid of rats and sheep.98  This concentration is found in some types of hay and in silages.

 

Bromine. Bromine is concentrated by the thyroid and interferes with the thyroidal 131I uptake in animals99,99a and humans, possibly by competitive inhibition of iodide transport into the gland. Bromine can also induce alterations in cellular architecture, blood supply and can lead to a reduction in T4 and T3 levels.99b

 

Rubidium. Rubidium is goitrogenic in rats.100  However, the mechanism of action is unknown.

 

Fluorine. Fluorine is not concentrated by the thyroid but has a mild antithyroid effect, possibly by inhibiting the iodide transport process.101  In large amounts, it is goitrogenic in animals. The amounts of fluorine consumed in areas with endemic fluorosis are not sufficient to interfere with thyroid function or to produce goiter.102,103  However, other data suggest that dietary fluorine may exacerbate an iodine deficiency and thus modulate the distribution of goiter in areas with low iodine intake.104

 

Cobalt. Cobalt inhibits iodide binding by the thyroid.105  The mechanism is unknown. Cobalt deficiency is associated with a reduction in type I monodeiodinase activity and a fall in T3105a while cobalt excess may produce goiter and decreased thyroid hormone production. 105b It is sufficiently active to have been used in the treatment of thyrotoxicosis.106

 

Cadmium.   Administration of cadmium to rats or mice decreases serum levels of T4 and T3. 106a,106b   It also decrease the activity of hepatic Type I - 5’Deiodinase.106a,106c

 

Lithium Ion. Lithium ion is goitrogenic when used in the treatment of manic-depressive psychosis and can induce myxedema.107  Experimentally, lithium increases thyroid weight and slows thyroid iodine release.108  When lithium carbonate was given to human subjects in doses of 900 mg four times daily, there was a significant decrease in the rate of release of thyroidal iodine in euthyroid and hyperthyroid subjects.109  Lithium also decreases the rate of degradation of T4 in both hyperthyroid and euthyroid subjects.110  Inhibition of thyroid hormone release may be the dominant effect of the ion.110a Therefore, the decrease in serum T3 concentration is greater in hyperthyroid patients, and changes in the rT3 level, if any, are minimal.111-113

A number of mechanisms have been suggested for the effects of lithium. One well-documented phenomenon is a potentiation of an iodide-induced block of binding and hormone release,88,114 perhaps because lithium is concentrated by the thyroid115 and increases the intrathyroidal iodide concentration109,111 (Fig. 5-2). Although it has been shown that lithium inhibits the adenylate cyclase activity in the thyroid gland as well as in other tissues,116 it also blocks the cAMP-mediated translocation of thyroid hormone. The latter effect, which is probably responsible for the inhibition of hormone release, appears to be due to the stabilization of thyroid microtubules promoted by lithium.117 In rat brain, lithium administration decreased both the levels of the Type II 5’Deiodinase and the Type III 5 Deiodinase.117a In the rat, lithium may also lead to an alteration in the distribution of thyroid hormone receptors with the alpha 1 isoform being increased in the cortex and decreased in the hypothalamus while the beta isoform was also decreaseed in the hypothalamus. 117b  

An exaggerated response of TSH to TRH may be seen in a majority of lithium treated patients110a but an elevated basal TSH is usually absent. An increase in the basal serum TSH concentration and its response to TRH most likely represents an early manifestation of hypothyroidism rather than a direct effect of lithium on the hypothalamic-pituitary axis.118  The prevalence of goiter has been reported to be as high as 60%.110a     Based on studies in FRTL-5 cells, lithium may have direct mitogenic effects on the thyroid that are independent of TSH and cAMP. 110b The occurrence of hypothyroidism during lithium therapy occurs in 10-40% of lithium treated patients and is far more frequent in women than men.110a,118a, 118b,118c  

 

Although much less frequent, lithium therapy has been associated with the development of thyrotoxicosis.110a Lithium is also reported to produce exophthalamos during chronic therapy; the condition regresses when treatment is stopped. The phenomenon is a protrusion of the globe but does not involve the other changes of infiltrative ophthalmopathy of Graves' disease.118,119

 

 

Selenium. Selenium is a component of the enzymes glutathione peroxidase (GSH-Px) and superoxide dismutase, both enzymes responsible for protection against free radicals. In addition, Type I 5’Deiodinase also contains selenium.119a 119c Thus, a deficiency of selenium could predispose the thyroid to oxidative injury and lead to decreased peripheral T3 production. In the elderly, reduced selenium levels have been associated with a decreased T3/T4 ratio.119b It has been postulated that the combined deficiency of iodine and selenium in Zaire results in myxedematous rather than neurologic cretinism because the decrease in peripheral conversion to T3 results in greater delivery of T4 into the neonatal developing brain.119c In rats, selenium deficiency led to a decrease in renal but not hepatic Type I 5’ Deiodinase activity and serum T3 levels were unaffected.119d Selenium deficiency led to decrease GSH-Px activity in the liver, kidney and rbc’s but not the thyroid.119d   Serum T4 was normal when both dietary iodine and selenium were both deficient, but was reduced when either was deficient alone.119d   In other studies, brain GSH-Px and Type I deiodinase activity were normal in the presence of iodine or selenium deficiency while brain Type II Deiodinase activity was increased by iodine deficiency and unaffected by selenium deficiency.119e In contrast in brown adipose tissue (BAT), both selenium and iodine deficiency led to decreased deiodinase activity and decreased production of the uncoupling protein.119e

 

            Treatment of goitrous children with combined seleium and iodine deficiency leads to a reduction in serum TSH and goiter size.119f The response, however, was correlated with the selenium level with both the goiter and TSH responses being correlated with the baseline selenium level. 119f  In an epidemeological study in China, low selenium levels were assocated with an increased ididence of goiter, sub-clinical and overt hypothryoidism and thyroiditis. 119g

 

Physical and Emotional Stress

Perhaps the most dramatic study of emotional stress is that reported by Kracht,120 who found that stress provoked thyrotoxicosis in wild rabbits. Although some stress models may prompt secretion of thyroid hormone in animals,120,121 this effect is unlikely to occur in humans, at least for a sustained period of time. The stress-induced increase in adrenocortical activity tends not only to suppress TSH release but also to inhibit T3 production. A major problem in the analysis of available date is the difficulty in separating effects produced by non-specific stress from the effects caused by the agents used to induce the stress. Many of the changes in thyroid function described in this chapter under the headings starvation, temperature, altitude and anoxia may be due, in part, to stress.

 

Surgery

Surgery has been used as a means to study the effect of stress on thyroid physiology in animals.122  Studies in humans have been prompted by the suspicion that thyroid hormone may mediate the postoperative metabolic changes leading to increased oxygen consumption and protein wastage. Some discrepancies in available data stem from lack of uniformity in the groups of patients studied in terms of preoperative state or disease, type of surgery, types of anesthetic agents and other drugs used, and the postoperative course, including nutrition and the period of recovery.

The most striking change in thyroid function is a decrease in the serum TT3 and FT3 concentrations shortly after surgery; rT3 concentrations are elevated in the postoperative period.123,124  The combined findings suggest a diversion in the normal deiodinative pathways of T4. FT4 levels may also be depressed in the postoperative period, but to a lesser degree.124  The TTR but not the TBG level is sharply reduced.125  This clear reduction in the concentration of the active forms of thyroid hormone during the postoperative period is preceded by a small, short-term increase in FT4 and FT3 concentrations during surgery.123,124  The magnitude of the subsequent reduction in T3 level appears to correlate with the severity of trauma and the morbidity during the postoperative course.123  The serum TSH concentration also tends to diminish,124 except during surgery performed in children under the conditions of hypothermia.2

Because surgical trauma produces a prompt elevation in plasma cortisol levels and food intake is curtailed during the pre-, intra-, and postoperative periods, the possibility that glucocorticoids and starvation are the principal contributors to the observed changes in thyroid function has been given strong consideration. However, Brandt et al.126 showed equally profound diminution in the serum T3 concentration when surgery was carried out with epidural anesthesia, which abolishes the plasma cortisol surge. Similarly, the almost routine use of glucose infusion should have been able to prevent the changes in serum T3 and rT3 levels if starvation played a major role in producing the changes observed during surgery.

 

Acute Mental Stress

Data on the effect of emotional stress on thyroid function in humans are principally derived from studies in patients with psychiatric disturbances. Thus, even if only patients with acute psychiatric decompensation are considered, the results are colored by the nature of the mental illness, its antecedent history, and the use of drugs. An early suggestion of enhanced hormonal secretion came from the observation of elevated protein-bound iodine (PBI) levels in the serum of psychiatric patients presumably under emotional stress and in medical students in the course of examinations.127  In more recent studies, elevations of the FT4I have been consistently found during admission of acute psychiatric patients. The incidence ranged from 7 to 18%.128-130  In one study, an equal number of patients (9%) had a low FT4I.128  In most instances, values became normal with time and treatment of the psychiatric illness. The TSH response to TRH is blunted or even absent in most psychiatric patients with elevated FT4I.130 Significant abnormalities in the serum T3 concentration are rare.

 

 

CHEMICALS AND DRUGS

Goitrogens

A number of compounds have the ability to inhibit thyroid hormone synthesis (Fig.5-3). Irrespective of their mechanism of action, they are collectively called goitrogens. As a result of a decrease in serum thyroid hormone levels, TSH secretion is enhanced, causing goiter formation. Some goitrogens occur naturally in food, and others are in drugs with goitrogenic side effects. The least toxic and those possessing the highest thyroid-inhibiting activity are used in the treatment of hyperthyroidism.

 

 

Dietary Goitrogens

The discovery of natural and synthetic substances that impair the synthesis of thyroid hormone are landmarks in the history of pharmacology.131  These substances are discussed in more detail in Chapter 20. Although iodide deficiency is, without doubt, the major cause of endemic goiter and cretinism throughout the world, dietary goitrogens may play a contributing role in some endemics, and may possibly be the dominant factor in certain areas. The dietary goitrogens fall into several categories, more than one of which may occur in the same food.

Certain foods contain cyanogenic glucosides,132 compounds that, upon hydrolysis by glucosidase, release free cyanide. These foods include almond seeds and such important dietary items as cassava, sorghum, maize, and millet. Cassava contains enough cyanogenic glucoside to be lethal if large quantities are consumed raw. Ordinarily, the root is extensively soaked, then dried and powdered. Most of the cyanide is lost in this process; that left in the root is liberated after ingestion and converted to SCN-. Chronic poisoning due to cassava is responsible for a tropical neuropathy in Nigeria133 and Tanzania, and is suspected of being a contributing cause of goiter in Central Africa.134,135

Other important classes of antithyroid compounds arise from hydrolysis of the thioglucosides.132,136,137  These compounds are metabolized in the body to goitrin or thiocyanates and isothiocyanates, and ultimately to other sulfur containing compounds, or are excreted as such. They are important in the goitrogenic activity of seeds of plants of the genus Brassica and the cruciferae, compositae, and unbelliferae. Among the plants containing these compounds are cabbage, kale, brussel sprouts, cauliflower, kohlrabi, turnip, rutabaga, mustard, and horseradish. Myxedema was reported in a woman without previous thyroid disease who consumed extremely large amounts of raw bok choy. 137b    Cattle may ingest these goitrogens and pass them to humans through milk, as observed in Australia,138 Finland,139,140 and England.141 . The isothiocynate, cheiroline, occurs in the leaves of choumoellier and may be related to a focal area of endemic goiter in Australia. The goitrogen is thought to be transmitted from forage to cows, to milk, and finally to children. Although there is considerable circumstantial evidence relating these compounds to endemic goiter, it has been difficult to prove their role with certainty.

Thiocyanate is a well-known inhibitor of iodide trapping when in high concentration in blood. The blood levels obtained by ingestion of dietary goitrogens are rarely of this degree. Inhibition of iodide trapping, and thyroid peroxidase activity, and augmentation of urinary iodide loss, as demonstrated by Delange and Ermans and co-workers, all may play a role in the goitrogenic activity.132,134,135  Thiocyanate may also reduce the iodine content of breast milk or animal milk and thus indirectly impact the thyroid function of young children in areas of marginal iodine sufficiency.141a A study in Thailand found an association between thiocyanate levels and TSH in pregnant women with low iodine excretion. 141b

Astwood et al. and Greer142,143 found that turnips contain progoitrin, which is a mustard oil thioglycoside. It undergoes rearrangement by enzymes in human enteric bacteria, or in the turnip, to be converted to goitrin, an active goitrogenic thioglycoside, L-5-vinyl-2-thio-oxazolidone.144,145  Goitrin inhibits oxidation of iodine and its binding to thyroid protein in the same way as do the thiocarbamides.

Several endemics of goiter have been attributed to dietary goitrogens, usually acting together with iodine deficiency. Goitrin is apparently present in cow's milk in Finland.146  In the Pedgregoso region of Chile, pine nuts of the tree Araucaria americana are made into a flour and consumed in large amounts, and may be related to endemic goiter.147,148  In the Cauca river valley of Colombia, sulfur-containing compounds found in the water supply, derived from sedimentary rocks containing a large amount of organic matter, are believed to be responsible for endemic goiter.149  At least, extracts from these waters are goitrogenic in rats. Pearl millet has been reported to cause goiter development in goats. 149a

Other mechanisms may also contribute to dietary goitrogenicity. Thus, diets high in soybean components or other materials increasing fecal bulk may cause excess fecal loss of T4 and increase the need for this hormone.150-153  These diets are low in iodine content, and soybean has been thought but not proven to contain a goitrogen.

The goitrogens, by blocking hormone synthesis, deplete the thyroid of iodide; this reduction itself increases the sensitivity of the gland to TSH.154  This sensitivity, in turn, further promotes goitrogenicity.

 

Antithyroid Drugs

According to their principal mode of action on thyroidal iodine metabolism, antithyroid drugs are divided into two categories: (1) the monovalent anions, which inhibit iodide transport into the thyroid gland, and (2) a large number of compounds that act through inhibition of thyroidal iodide binding and iodotyrosine coupling. The most important representatives of this latter category of compounds are the group of thionamides. The effect of the drugs in the first category is counteracted by exposure to excess iodine, whereas iodine has no inhibition, and at times even potentiates, the action of drugs in the second category. Other drugs inhibit thyroid hormone secretion or act through yet unknown mechanism. A list of these agents is provided in Table 5-1.

 

Monovalent Anions. Certain monovalent anions (SCN-, Cl04-, NO3-) inhibit transport of iodide into the thyroid gland and thereby depress iodide uptake and hormone formation.164-166  Thiocyanate stimulates efflux of iodide from the thyroid as well,167 and also inhibits iodide binding and probably coupling.168,169  A large number of complex anions, such as monofluorosulfonate, difluorophosphate, and fluoroborate,170 inhibit iodide transport. Of these, fluoroborate,171 and perchlorate,172 are concentrated by the thyroid gland. These ions have a molecular volume and charge similar to those of iodide, and may compete with iodide for transport.170,171  Perchlorate is sufficiently active to be useful clinically.173  Perchlorate and thiocyanate also displace T4 from thyroid hormone-binding serum proteins in vivo and in vitro and cause a transient elevation of free T4.174  In contrast to the pharmacologic effects of perchlorate, concerns have been raised about the potential health effects of environmental perchlorate exposure, especially in municipal water supplies. Several studies have been unable to detect an increase in hypothyroidism 174a. 174b, congenital hypothyroidism 174c, or thyroid cancer 174d in exposed populations, but a study in Thailand found an association bwteen perchlorate levels and TSH in pregnant women. 141b

 

Thionamides. The thionamide and thiourylene drugs do not prevent transport of iodide into the thyroid gland, but rather impair covalent binding of iodide to TG.175-177  They may be competitive substrates for thyroid iodide peroxidase, preventing the peroxidation of iodide by this enzyme. In small doses, the thiocarbamides inhibit formation of iodothyronines from iodotyrosyl precursors. When slightly larger amounts are present, iodination of MIT and tyrosine is prevented.177,178  Minute amounts (10-8 M) have, paradoxically, a stimulatory effect on iodination in thyroid slices.179

The basic structure necessary for the antithyroid action of these drugs is

 

N

|

S

|

-N=C-X-

where X may be C, N, or O180,181 (Fig. 5-3). The thiocarbamides are metabolized in the thyroid gland by transsulfuration.182  The enzyme responsible may also be involved in the iodide peroxidase enzyme system.183  Glands under TSH stimulation metabolize the antithyroid drugs at an accelerated rate, as has been shown for thiourea.184

Iodide is released more rapidly from a gland blocked by PTU than from one blocked by perchlorate.165,185  This action occurs presumably because PTU prevents the utilization of all iodide available to the gland (transported from the blood or formed in the gland by deiodination of iodotyrosines), whereas potassium perchlorate prevents uptake of iodide but does not inhibit reutilization of iodide derived from within the gland. T4 disappears from the PTU-blocked rat thyroid at a faster rate than do iodotyrosines.185

In addition to the effects on the thyroid gland, PTU (and, to a much lesser extent, methimazole) partially inhibits the peripheral deiodination of T4186-191 and its hormonal action.188,192-194  PTU acts directly on body tissues to inhibit the normal formation of T3 from T4.191,195  Coincidentally, fecal excretion of T4 increases.186  In order to inhibit goiter induced by antithyroid drugs in rats, one must maintain the T4 concentration in blood at a higher level that is normal for the species.188,192  Presumably, inhibition of T4 monodeiodination by the antithyroid drug leads to a buildup of T4 in blood and diminishes the availability of T3 in the tissues.191  Higher doses of T4 or higher blood levels may be sufficient to push the reaction toward T3 and allow formation of quantities sufficient to prevent goiter.

Metabolism of the antithyroid drugs has been observed after administration of 35S-labeled drugs. Methimazole is rapidly absorbed from the gastrointestinal tract in humans. It reaches a peak plasma level about an hour after administration, and then declines gradually to near zero levels at 24 hours. These drugs are accumulated and degraded in the thyroid, since they are substrates of the peroxidase.196,197  Carbimazole is accumulated as its metabolic product, methimazole. The concentration ratio between thyroid and plasma for unmetabolized methimazole in rats may approach 25, eight hours after administration of the drug. The metabolic products derived from the drug are excreted in the urine, largely during the first day.

 

Other Goitrogenic Compounds

A number of other drugs, including the aminoheterocyclic compounds and substituted phenols, act as goitrogens principally by impairing TG iodination (Fig. 5-3). They are in general far less potent in their goitrogenic effect than the thionamides. None are used therapeutically as antithyroid drugs; rather, goitrogenesis is an undesirable side effect of their use. Some the compounds have multiple effects and thus influence thyroid physiology at various levels. These compounds are individually discussed in greater detail. A comprehensive list is provided in Table 5-1.

Table 5-1     Agents Inhibiting Thyroid Hormone Synthesis and Secretion

 

Block iodide transport into the thyroid gland  
Substance Common Use
Monovalent anions (SCN-, Cl04-, N03-)a Not in current use; Cl04- test agent
Complex anions (monofluorosulfonate,difluorophosphate, fluoroborate)a -------
Minerals (bromine, fluorine) In diet
Lithiuma Treatment of manic-depressive psychosis
Ethionamide Antituberculosis drug
   
   
Impair TG iodination and iodotyrosine coupling  
Substance Common Use
Thionamides and thiourylenes, (PTU,methimazole, carbimazole)a Antithyroid drugs
Sulfonamides (acetazolamide, sulfadiazine, sulfisoxazole)a Diuretic, bacteriostatic
Sulfonylureas (carbutamide, tolbutamide, metahexamide, ?chloropropamide)a Hypoglycemic agents
Salicylamides (p-aminosalicylic acid, p-aminobenzoic acid)a Antituberculosis drugs
Resorcinol Cutaneous antiseptic
Amphenone Anticonvulsive
Aminoglutethimide Antiadrenal agent
Thiocyanatea No current use; in diet
Antipyrine (phenazone)a Antiasthmatic
Aminotriazole "Cranberry poison”
Amphenidone Tranquilizer
2,3-Dimercaptopropanol (BAL) Chelating agent
Ketoconozole Antifungal agent
   
   
Inhibitors of thyroid hormone secretion  
Substance Common Use
Amiodarone a Antianginal and antiarrhythmic agent
Iodide (in large doses)a Antiseptic, expectorant, and others
Lithiuma Treatment of manic-depressive psychosis
Cause Thryoiditis  
Substance Common Use
Amiodarone a Antianginal and antiarrhythmic agent
Interleukin II a Cancer therapy
g-Interferon a Antiviral and cancer therapy
Sunitinib a Cancer therapy
Sorafenib a Cancer therapy
Ipilmumab a Cancer therapy
Pembrolizumab a Cancer therapy
Nivolumab  a Cancer therapy
   
   
Mechanism unknown  
Substance Common Use
p-bromdylamine maleate a Antihistamine
Phenylbutazone a Antiinflammatory agent
Minerals (calcium, rubidium, cobalt)a -------
Thalidomide396 Cancer therapy
   
   
aReferences given in the text  

 

Sulfonamides. Sulfonamides, particularly those containing an aminobenzene grouping, have antithyroid activity. Acetazoleamide (Diamox), the diuretic agent, has a strong effect on animals and humans.198,199  Its action, prevention of intrathyroidal iodide binding, is not related to carbonic anhydrase inhibition. Sulfadiazine and sulfisoxazole have a similar action, probably through a synergistic effect on iodide.89

 

Sulfonylureas. Sulfonylureas, derivatives of sulfonamides and used as hypoglycemic-antidiabetic agents, also inhibit the synthesis of thyroid hormone. They include carbutamide, tolbutamide, methahexamide, and possibly chlorpropamide, but not the phenylethyl biguanide (Fig. 5-3). They impair thyroidal RAIU and cause goiter in the rat.200,201  Carbutamide is much more potent than tolbutamide. Carbutamide, 2 g/day (but not 1 g/day), may reduce the thyroidal RAIU in humans to 20% of control values, but the uptake gradually rises as treatment is continued and is normal after 20 weeks. From 1 to 2 g tolbutamide per day does not affect RAIU in humans.202  Thus, in the usual dose range, tolbutamide will not depress thyroid function.

Chlorpropamide in large doses (3-7 g) depresses the RAIU in humans; the common therapeutic doses (up to 1 g daily) usually have no effect on serum T4.203  A mild antithyroid action is often reflected in a rise in RAIU, which may be found after the agents are withdrawn.

These drugs inhibit hormone synthesis by inhibition of iodide binding. In most instances, the pituitary compensates for the effect and maintains a euthyroid state by increased synthesis of TSH. Nevertheless, hypothyroidism is said to be more common in diabetic patients on sulfonylureas than in patients treated by other means.204

Sulfonylureas also block binding of T4 to the carrier proteins in serum and thus depress the T4 concentrations.205  This effect is most pronounced after intravenous administration.

 

Polychlorinated Biphenyls   Animal studies have suggested that polchlorinated bihenyls (PCBs) may reduce thyroid hormone levels by decreasing synthesis, increasing biliary excretion of conjugated metabolites and displacing T4 from binding proteins. 205a A review of studies in humans, did not find significant or consistent changes. 205a

  

Effects of Miscellaneous Compounds and Drugs

 

General Mechanisms of Action

A large number of substances may affect thyroid gland function and thyroid hormone metabolism and action. The list continues to grow with the introduction of new diagnostic agents, drugs, and food additives. Drugs affect the transport, metabolism, action and excretion of T4 and its derivatives as well as regulation at all levels of the hypothalamic-pituitary-thyroid axis. Some drugs may induce hypothyroidism or thyrotoxicosis, and if autoimmune mechanisms are involved, the thyroid dysfunction may not resolve with discontinuance of the drug. Some compounds may not have any direct effect on thyroid hormone economy or regulation, but have clinical relevance by interfering in specific diagnostic assays.

Compounds are discussed and listed below based on their major mechanisms of action. Many drugs have more than one mechanism of action and the explanation for observed abnormalities is not always known. Results of experiments conducted in animals or in vitro are not always applicable to human pathophysiology. Compounds which alter thyroid hormone secretion are generally goitrogens or anti-thyroid drugs and were discussed in the preceeding section. Selected compounds with significant effects on the thyroid, wide-spread use or that are of particular interest in understanding the mechanism of drug effects are described in greater detail.

 

Alterations of Thyroid Hormone Transport

Some hormones and drugs may affect thyroid hormone transport in blood by altering the concentration of the binding proteins in serum. Thyroid hormone transport may also be affected by substances that compete with the binding of thyroid hormone to its carrier proteins (Table 5-2). TBG synthesis is increased by estrogens220-223 and decreased by androgens and anabolic steroids.223,224 Estrogen’s effect to increase TBG is blunted or reversed by tamoxifen and raloxifene.224a The most extensively studied compounds that interfere competitively with thyroid hormone binding to the carrier proteins in serum are salicylates, diphenylhydantoin, and heparin.212,225-231,231a,b A clinically significant effect of furosemide211 may only be seen with very high doses and with accumulation with renal failure.

Table 5-2     Compounds that Affect Thyroid Hormone Transport Proteins in Serum

 

Increase TBG concentration  
Substance Common Use
Estrogensa Ovulatory suppressants, anticancer agents, hormone replacement
Heroin and methadone206 Opiates (in addicts)
Clofibrate207 Hypolipemic agent
5-Fluorouracil208 Anticancer agent
Perphenazine209 Tranquilizer
   
   
Decrease TBG concentration  
Substance Common Use
Androgens and anabolic steroidsa Virilizing, anticancer, and anabolic agents
Glucocorticoidsa Antiinflammatory, immunosuppressive, and anticancer agents; decrease intracranial pressure
L-Asparaginase210 Antileukemic agent
Nicotinic acid210a ,210b Hypolipidemic agent
   
   
Interfere with thyroid hormone binding to TBG and/or TTR  
Substance Common Use
Salicylates, 4 amino-salicylic acid and salsalatea Antiinflammatory, analgesic, antipyretic, antituberculosis agents
Diphenylhydantoin and analogsa Anticonvulsive and antiarrhythmic agents
Furosemide211 Diuretic
Sulfonylureasa Hypoglycemic agents
Heparina Anticoagulant
Dinitrophenola Uncouples oxidative phosphorylation
Free fatty acids212,213 --------
o,p'-DDD214 Antiadrenal agent
Phenylbutazone215 Antiinflammatory agent
Halofenate216 Hypolipemic agent
Fenclofenac217 Antirheumatic agent
Orphenadrine218 Spasmolytic agent
Monovalent anions (SCN-, C104-)a Antithyroid agents
Thyroid hormone analogs, including dextroisomers219 Cholesterol reducing agents
   
   
aReferences given in the text  

 

 

 

In general, the effect of increased hormone binding is an increase in the serum concentration of total (bound) T4 and of reduced binding is a decrease in the total (bound) T4, with T3 effected to a lesser extent. There is no significant effect on the absolute concentration of the metabolically active fractions of FT4 and FT3, or usually their free indices (FT4I and FT3I). In the steady state, the quantity of thyroid hormone reaching peripheral tissues and the pathways and amount of hormone degradation remain unaltered. However, before this steady state is reached, an acute perturbation in the equilibrium between free and bound hormone brings about transient changes in thyroid hormone secretion and degradation. The hypothalamic-pituitary-thyroid axis participates in the reestablishment of the new steady state. For example, as illustrated in Figure 5-4, an abrupt increase in the concentration of TBG shifts the equilibrium between total and bound hormone, causing a decrease in the concentration of free hormone. The consequences are fourfold. First, there is a shift in the exchangeable hormone from tissues to blood. Second, a decreased hormone content in tissues diminishes its absolute degradation rate. Third, a decline in hormone concentration in tissues activates the hypothalamic-pituitary axis, causing an increase in TSH secretion. Fourth, the latter acts on the thyroid gland to step up its hormonal secretion and reestablish an appropriate thyroid hormone/TBG ratio. Thus, a normal thyroid hormone concentration in serum and tissues and hormonal production and disposal rates are reestablished. TSH concentration returns to normal, and a new steady state is maintained at the expense of an increased intravascular pool and a decreased fractional turnover rate and total distribution space of thyroid hormone.232,233  The reverse sequence of events accompanies an acute decrease in TBG concentration or binding (Fig. 5-4).

 

Alterations of Thyroid Hormone Metabolism

Agents that may alter the extrathyroidal metabolism of thyroid hormone are listed in Table 5-3. Several drugs with wide use in clinical practice inhibit the conversion of T4 to T3 in peripheral tissues. Glucocorticoids,239,240 amiodarone,241,242 and propranolol243-245 are a few examples. As expected, their most profound effect on thyroid function is a decrease in the serum concentration of T3,239,241,243 usually with a concomitant increase in the rT3 level.239,241  An increase in the serum T4 concentration has also been observed on occasion.241,245  The serum TSH concentration may also occasionally rise,241 provided the drug does not have a direct inhibitory effect on the hypothalamic-pituitary axis.246  In the absence of inherent abnormalities in thyroid hormone secretion or in its regulation, TSH levels should return to normal and hypothyroidism should not ensue from the chronic administration of compounds the only effect of which is to interfere partially with T4 monodeiodination.

Other mechanisms by which some compounds affect the extrathyroidal metabolism of thyroid hormone are acceleration of the overall rates of deiodinative and nondeiodinative routes of hormone disposal. Examples of drugs acting principally through the former mechanism are diphenylhydantoin and phenobarbital,247-249 and via the latter, colestipol237, ferrous sulfate238a, aluminum hydroxide238b and sucralfate238c. Patients receiving these drugs should increase the secretion of hormone from the thyroid gland in order to compensate for the enhanced hormonal loss through degradation or fecal excretion. Thyroid hormone concentration in blood should remain unaltered. However, hypothyroid patients receiving such drugs may require higher doses of exogenous hormone to maintain a eumetabolic state (Chapter 9). In patients on thyroid hormone therapy who are also taking drugs which bind thyroid hormone in the gastrointestinal tract, the administration of the two drugs at different times will markedly reduce or eliminate the effect on thyroid hormone absorption.

Acute increases in serum T4 and FT4 concentration after the injection of insulin or during halothane anesthesia have been attributed to an enhanced release of T4 normally stored in the liver.250,251

 

Table 5-3 Agents that Alter the Extrathyroidal Metabolism of Thyroid

 

Substance Common Use
Inhibit conversion of T4 to T3
PTUa Antithyroid drug
Glucocorticoids (hydrocortisone, prednisone, dexamethasone)a Antiinflammatory and
immunosuppressive; Decrease
intracranial pressure
Propranolola ß-Adernergic blocker
(antiarrhythmic, antihypertensive)

Iodinated contrast agents [ipodate

(orgrafin), iopanoic acid

(Telepaque)]a

Radiologic contrast media
Amiodaronea Antianginal and antiarrhythmic agent
Clomipramine234 Tricylic antidepressant
Stimulators of hormone degradation or fecal excretion
Diphenylhydantoina Anticonvulsive and antiarrhythmic agent
Carbamazepine235 Anticonvulsant
Phenobarbitala Hypnotic, tranquilizing, and
anticonvulsive agent
Cholestyramine236  and colestipol237 Hypolipemic resins
Soybeans151 152 Diet
Rifampin238a Antituberculosis drug
Ferrous Sulfate238 Iron therapy
Aluminum hydroxide238b Antacid
 Sucralfate            238c Anti-ulcer therapy
Imatinib 384 Cancer therapy
Bexarotene 387 Cancer therapy
Sevelemer 393 Phosphate Binder
Colesevelam 394 Hypolipemic resin
Lanthanum Carbonate 394                                Phosphate Binder
Coffee 395 Diet
aReferences given in the text  

 

Alterations of Thyroid Hormone Regulation

The last two decades have seen a prodigious growth in the list of substances that can be shown to act on the hypothalamic-pituitary axis (Table 5-4). Although many of these compounds are used frequently, only a few have significant effects on thyroid function via this central mechanism. Furthermore, patients receiving these drugs rarely have any abnormality of serum TSH although the response of TSH to the administration of TRH may be altered. An effect of these drugs may be seen in patients with untreated or partially treated primary hypothyroidism. In patients with an elevated basal level of serum TSH, addition of these drugs may produce a further increase or a significant diminution.

Although the following paragraphs discuss the general mechanisms of action for these compounds, specific mechanisms are not always known. A major problem in interpretation is the variability of experimental designs. These variables include doses, routes of administration, duration and time of treatment, drug combinations, age and sex of subjects, hormonal status at the time of testing, and time of blood sampling. Furthermore, observed responses may be effected by the method of data analysis. For example, results of TSH responses to TRH have been expressed in terms of changes in the absolute value, increments or decrements from the basal level, and percent of the basal value at either the peak and nadir of the response or the integrated area over the duration of the response.

The most potent suppressors of pituitary TSH secretion are thyroid hormone and its analogs. They act on the pituitary gland by blocking TSH secretion through the mechanisms discussed in Chapter 4. Some TSH-inhibiting agents listed in Table 5-4, such as, fenclofenac and salicylates, may act solely by increasing the free thyroid hormone level through interference with its binding to serum proteins.   Other agents appear to have a direct inhibitory effect on the pituitary and possibly on the hypothalamus. The most notable is dopamine and its agonists. They have been shown to suppress the basal TSH levels in euthyroid persons284,285 and in patients with primary hypothyroidism.267,284-286  More uniformly, they suppress the TSH response to the administration of TRH.268,285,287,288  In contrast, most dopamine antagonists increase TSH secretion.150-155  Increases in the basal TSH and in its response to TRH have been observed in euthyroid persons,252,255 as well as in patients with primary hypothyroidism250-256 who have been given these drugs. A notable exception to this rule, which casts some doubt on the assumed mechanism of action of dopamine antagonists, is neuroleptic dopamine blocker, pimozide, which has been reported to reduce the elevated serum TSH level in patients with primary hypothyroidism.289

 

Table 5-4     Agents that May Affect TSH Secretion

 

Substance Common Use
Increase serum TSH concentration and/or its response to TRH
Iodine (iodide and iodine-containing compounds)a Radiologic contrast media, antiseptic expectorants, antiarrhymic and antianginal agents
Lithiuma Treatment of bipolar psychoses

Dopamine receptor lockers    (metclopramide,252,253 domperidone253 254)

 

Antiemetic
Dopamine-blocking agent(sulpiride255 ) Tranquilizer

Decarboxylase inhibitor        (benserazide256)

 

Dopamine-depleting agent(monoiodotyrosine253)

 

L-Dopa inhibitors(chloropromazine,257 biperidine,258 haloperidol258) Neuroleptic drugs

Cimetidine (histamine receptor blocker)259

 

Treatment of peptic ulcers
Clomifene (antiestrogen)260 Induction of ovulation
Spironolactone261 Antihypertensive agent
Amphetamines262 Anticongestants and antiappetite
 

Decrease serum TSH concentration and/or its response to TRH

 

Thyroid hormones (T4 and T3)

Replacement therapy, antigoitrogenic and anticancer agents

 

Thyroid hormone analogs (D-T4,263 3,3',5-Triac,264 etiroxate-HCl,265 3,5-dimethyl-3-isopropyl-L-thyronine266) Cholesterol-lowering and weight reducing agents
Dopaminergic agents (agonists)  
Dopaminea Antihypotensive agent
L-Dopaa (dopamine precursor) Diagnostic and anti-Parkinsonian agent
2-Brom-alpha-ergocryptinea Antilactation and pituitary tumor suppressive agent
Fusaric acid (inhibitor of dopamine ß-hydroxylase267)
Pyridoxine (coenzyme of dopamine synthesis268) Vitamin and antiheuropathic agent
Other dopaminergic agents  (perbidil,269 apomorphine,269 lisuride270 ) Treatment of cerebrovascular diseases and migraine
Dopamine antagonist (pimozide)a

Neuroleptic agent

 

alpha-Noradrenergic blockers     (phentolamine,271 thioridazine272) Neuroleptic agents
Serotonin antagonists metergoline,273           cyroheptadine,274 methysergide275) Antimigraine agents and appetite stimulators
Serotonin agonist(5-hydroxytryptophan276)
Glucocorticoidsa    Antiinflammatory, immunosuppressive, and anticancer agents. Reduction of intracranial pressure
Acetylsalicylic acida

Antiinflammatory, antipyrexic and

analgesic agent

Growth hormone277 b                                            Growth-promoting agent
Somatostatin278,279                                                                
Octreotide 279a Treatment of carcinoids, acromegaly and other secretory tumors
Opiates (morphine,280 leucine- eukephaline,281 heroin282) Analgesic agents
Clofibrate283 Hypolipemic agent
Fenclofenac216 Antirheumatic agent
Bexarotene a Cancer therapy
Metformin 392                                                          Anti-diabetic agent
Ipilmumab a (autoimmune hypophysitis) Cancer therapy
Pembrolizumab a (autoimmune hypophysitis) Cancer therapy
Nivolumab a (autoimmune hypophysitis) Cancer therapy

 

aReferences given in the text

bIn hyposomatotrophic dwarfs

 

Iodine and some iodide-containing organic compounds cause a rapid increase in the basal and TRH-stimulated levels of serum TSH. This effect is undoubtedly due to a decrease in the serum thyroid hormone concentration either by inhibition of hormone synthesis and secretion by the thyroid gland81,82 or by a selective decrease in the concentration of T3.290  The latter effect is mediated through the inhibition of T3 generation from T4. A more selective, intrapituitary inhibition of T4 to T3 conversion appears to be responsible for the TSH-stimulating effect of the radiographic contrast agent iopanoic acid58 and amiodarone. Iodine does not stimulate TSH secretion in patients in whom it has produced hyperthyroidism. 94  A decrease in the free thyroid hormone concentration in serum, albeit minimal in magnitude, may also be responsible for the increase in TSH levels observed during treatment with clomifene.260

It has been postulated that some agents may act by modifying the effect of TSH on its target tissue. For example, theophylline may potentiate the action of TSH through its inhibitory effect on phosphodiesterase, which may lead to an increase in the intracellular concentration of cAMP.291  In fact, the presence of the pituitary is required to demonstrate that methylxanthines augment the goitrogenic effect of a low-iodine diet in the rat.292  One of the postulated effects of diethyl ether anesthesia in the rat is inhibition of the action of TSH on the thyroid gland,293 although it has also been reported to induce a transient redistribution of T4 between serum and tissues.294

 

Alterations of Thyroid Hormone Action

A handful of drugs seem to act by blocking some of the peripheral tissue effects of thyroid hormone. Others appear to mimic one or several manifestations of the thyroid hormone effect on tissues. Guanethidine releases catecholamines from tissues.295  It has a beneficial effect in thyrotoxicosis, including a decrease in BMR, pulse rate, and tremulousness.296,297  This agent has little effect on the thyroid gland, but depresses manifestations of thyrotoxicosis that are mediated by sympathetic pathways. The sympatholytic agents phentolamine and dibenzyline have been reported both to depress and to stimulate thyroid function in animals. Their action is not clear, and it is of minimal clinical significance.298-300  Among several α-adrenergic blocking agents tested, only phentolamine showed an inhibitory effect on the TSH response to TRH.271

                  Theoretically, thyroid hormone effects could be blocked by drugs which interfere with the tissue uptake of thyroid hormone or binding to its receptors. Inhibition of both cellular uptake and nuclear receptor binding has been demonstrated in vitro for amiodarone in hepatocytes and cultured pituitary cells. Inhibition of cellular thyroid hormone uptake has also been reported for calcium channel blockers and benzodiazapines. Furosemide and non-steroidal anti-inflammatory drugs reduce T3 binfding to cytosolic receptors.   There is, however, no clear evidence that any of these drugs have a clinically significant effect on thyoid hormone action.

Among the multiple effects the ß-adrenergic blocker, propranolol, has on thyroid hormone economy, it appears to reduce the peripheral tissue responses to thyroid hormone (see also Chapters 3 and 11). Dinitrophenol enhances oxygen consumption by a direct effect on tissues and thus mimics one of the actions of thyroid hormone.301

Recent interest has been directed toward compounds which may share some but not all thyroid hormone actions by either selective tissue uptake or receptor binding. The general goal is to develop agents which promote weight loss or decrease lipids without adverse effects on the skeleton, heart (tachycardia) or pituitary (TSH suppression). 300a-300d  Diiodothyropropanoic acid (DIPTA) in short term studies was found to decrease cholesterol and lead to weight loss. 300a However it was also found to increase bone turnover and reduce TSH, T3 and T4. 300a

The drug eprotirome was shown to reduce total and LDL cholesterol, triglyerides and Lp(a) lipoprotein. 300b,300c Eprotirome was not found to have adverse effects on the heart or bone and did not changed levels of TSH or T3 although mild, reversible dops in T4 levels were noted. 300b,300c In a controlled trial, that was terminated prior to completion when adverse cartilage effects were noted in dogs, several patients did devlop transaminase elevations. 300c

 

Specific Agents

Estrogens and selective estrogen receptor modulators (SERMs). Hyperestrogenism, either endogenous (caused by pregnancy, hydatidiform moles, or estrogen-producing tumors) or exogenous (due to the administration of estrogens), is accompanied by an increase in TBG and a decrease in TTR concentrations in serum.220-222  Estrogens are the most common cause of TBG elevation, and this effect can be produced even after their topical application. The magnitude of TBG increase is in part dose related and occurs in women as well as in men. While tamoxifen blocks the estrogen induced increase of TBG224a, tamoxifen alone in post-menopausal women increases TBG and T4 and 3 levels.301a . The selective estrogen receptor modulator (SERM) raloxifene, increases TBG, produces small increase in T4 and insignificant changes in free T4. 301b,301c In a single case report, raloxifene appeared to also alter thyroid hormone absorption. 301d Estrogen increases the complexity of oligosaccharide side chains and, as a consequence, the number of sialic acids in the TBG molecule which in turn prolongs its survival in serum.302  The concentrations of other serum proteins, including several that bind hormones, such as cortisol-binding globulin and sex-hormone binding globulin, are also increased.303

The consequences of increased TBG concentration in serum are higher serum levels of T4, T3 and rT3 and, to a lesser extent, other metabolites of T4 deiodination. The fractional turnover rate of T4 is depressed principally due to an increase in the intravascular T4 pool. On the other hand, the FT4 and FT3 concentrations and the absolute amount of hormone degraded each day remain normal.232,233  Transient changes in these parameters during the early changes in TBG concentration can be anticipated as described above. Some of the effects of pregnancy on thyroid function are also mediated by an estrogen-induced increase in the serum TBG concentration. The effects on thyroidal and renal iodide clearance and BMR are mediated by different mechanisms (see Chapter 3).

The effect of estrogen, if any, on the control of TSH secretion is controversial. Contradictory results suggesting a stimulatory304 and an inhibitory305,306 effect have been obtained by different investigators and both stimulation and inhibition has been shown in a single study depending on the dosage utilized.306a In a study of the effects of Tamoxifen, TSH was elevated at 3 months but not at 6 months.306b Although women show a greater TSH responsiveness to TRH than men,306-308 administration of pharmacologic doses of estrogens does not appear to have a significantly enhancing effect.309,310 During ovarian hyperstimulation for ovulation induction, an increase in TSH and fT4 has been observed and this has been attributed to the marked increase in estrogen. 310a

The effects of estrogens in the rat are not identical to those observed in humans. Estrogens do not induce changes in the concentration of serum T4-binding proteins in the rat.22  Thus, investigations carried out in this species are not always representative in interpreting the effects of estrogens observed in humans

 

Androgens. Androgens decrease the concentration of TBG in serum and thereby reduce the level of T4 and T3.223,311  The TTR concentration, however, is increased.223  As with estrogens, the concentration of free hormone remains unaffected, and the degradation rate of T4 is normal at the expense of an accelerated turnover rate.223  TSH levels are normal.305  Anabolic steroids with weaker androgenic action have the same effect, although similar changes observed during danazol therapy have been attributed to its androgen-like properties.224

 

Salicylates. Acetylsalycilic acid has been identified as the most commonly administered medication which may cause significant alterations in measured parameters of thyroid function.224b,224c   Salicylate and its noncalorigenic congeners (Fig. 5-3) compete for thyroid hormone-binding sites on serum TTR and TBG.225-228  As a result, the serum concentrations of T4 and T3 decline and their free fractions increase.228  The turnover rate of T4 is accelerated, but degradation rates remain normal.225,226  Salicylate and its noncalorigenic congeners also suppress the thyroidal RAIU but do not retard iodine release from the thyroid gland.312   The impaired respone to TRH313 and the hypermetabolic effect314 of salicylates have been attributed to the increase in the FT4 and FT3 fractions. If this were correct, hormonal release from the serum-binding proteins should produce only a temporary suppression of the thyroidal RAIU and transient hypermetabolism, but both effects are observed during chronic administration of salicylates.225,226  In addition, this mechanism of action does not explain the lack of calorigenic effect of some salicylate congeners despite their ability to also displace thyroid hormone from its serum-binding proteins.

In vitro studies have demonstrated an inhibitory effect of salicylate on the outer ring monodeiodination of both T4 and rT3,315 but lack of typical changes in serum iodothyronine levels suggests that this action is less important in vivo.

Acetylsalicylic acid mimics some actions of thyroid hormone, but does not reverse classic manifestations of hypothyroidism.   While salicylate administration may lower serum cholesterol levels,316 it does not provide a therapeutic effect in myxedema, or lower TSH levels.317 Administration of 8 g aspirin daily raises the BMR to normal in myxedema, accelerates the circulation, and increases sweating, but it has no effect on the skin change, the electrocardiogram, or the mental state.316

Because of some analogies between the effects of salicylates and nitrophenol, uncoupling of oxidative phosphorylation has been suggested as one of its possible mechanisms of action. If this were the case, direct chemical action does not appear to be involved since analogs of salicylate that do not uncouple oxidative phosphorylation in vitro are active in vivo.318

 

p-Aminosalicylic acid and p-aminobenzoic acid are closely related chemically to salicylate. They inhibit iodide binding in the thyroid gland and are goitrogenic.319,320  These agents also displace thyroid hormone from its serum protein-binding sites.321  Abnormalities of thyroid function tests have been also reported in patients treated with salsalate.322

 

Heparin. Patients receiving heparin chronically may have increased FT4 and FT3.230,231 Reciprocal changes in serum TSH have been reported.231  While it had been suggested that heparin might interact with the T4-binding proteins to alter the steric configuration of the binding sites and reduce the affinity of the proteins for T4 and T3210, it is now thought that heparin acts via the activation of lipoprotein lipase to increase free fatty acid levels which may displace T4 from binding proteins. This effect is most likely to be significant when the levels of albumin are low and triglycerides are high such as during hyperalimentation with lipid solutions. Even low doses of heparin may be sufficient to cause artifactual, in vitro, increase in T4 especially when measured by equilibrium dialysis.231a Although initially reported with crude heparin preparations, this heparin effect has also been noted with enoxaparin. 231b

 

Glucocorticoids. Physiologic amounts, as well as pharmacologic doses of glucocorticoids influence thyroid function. Their effects are variable and multiple, depending on the dose and on the endocrine status of the individual. The type of glucocorticoid and the route of administration may also influence the magnitude of the effect.323  Known effects include (1) decrease in the serum concentration of TBG and increase in that of TTR;324,325  (2) inhibition of the outer ring deiodination of T4 and probably rT3;239,240  (3) suppression of TSH secretion;246,326,327  (4) a possible disease in hepatic binding of T4; and (5) increase in renal clearance of iodide.328,329

The decrease in the serum concentration of TBG caused by the administration of pharmacologic doses of glucocorticoids results in a decrease in the serum total T4 concentration and an increase in its free fraction and the resin uptake test result. The absolute concentration of FT4 and FT4I remain normal. The more profound decrease in the concentration of serum T3 compared to T4 associated with the administration of pharmacologic doses of glucocorticoids cannot be solely ascribed to their effect on serum TBG. It is due to the decreased conversion of T4 to T3 in peripheral tissues. Thus, glucocorticoids reduce the serum T3/T4 ratio and increase that of rT3/T4 in hypothyroid patients receiving replacement doses of thyroid hormone.239  This steroid effect is rapid and may be seen within 24 hours.239,240 In rats, dexamethasone has been shown to decrease T4 to T3 conversion in liver homogenates. 329a

Earlier observations of cortisone-induced depression of uptake and clearance of iodide by the thyroid328,329 now are understood to be the result of steroid suppression of TSH secretion. Pharmacologic doses of glucocorticoids suppress the basal TSH level in euthyroid subjects and in patients with primary hypothyroidism, and decrease their TSH response to TRH.246,326,327,329b  The latter effect is less marked in the presence of hypothyroidism.327  Administration of as little as 34 mg. of hydrocortisone over 24 hours can be shown to reduce the pulse amplitude and mean TSH release the nocturnal rise of TSH and the T3 and TSH response to TRH.329b  Administration of the glucocorticoid antagonist, mifepristone, produces an increase in TSH that remains within the normal range accompanied by a transient decrease in total but not free T4.329c Normal adrenocortical secretion appears to have a suppressive influence on pituitary TSH secretion because patients with primary adrenal insufficiency have a significant elevation of TSH.330 In cultures from rat pituitary tumors, hydrocortisone increased the number of TRH receptors 331 Dexamethasone has also been shown to increase the transcription, translation and processing of TRH precursors. 331a,b   The mechanism of glucocorticoid action on the hypothalamic-pituitary axis is covered in Chapter 4.

No single change in thyroid function can be ascribed to a specific mode of action of glucocorticoids. For example, a diminished thyroidal RAIU may be due to the combined effects of TSH suppression and increased renal clearance of iodide. Similarly, a low serum TT4 level is the result of suppressed thyroidal secretion due to diminished TSH stimulation as well as the decreased serum level of TBG. One of the common problems in clinical practice is to separate the effect of glucocorticoid action on pituitary function from that of other agents and those caused by acute and chronic illness. This situation arises often since steroids are commonly used in a variety of autoimmune and allergic disorders as well as in the treatment of septic shock. The diagnosis of coexisting true hypothyroidism is difficult, if not impossible. Due to the suppressive effects of glucocorticoids on the hypothalamic-pituitary axis, the low levels of serum T4 and T3 may not be accompanied by an increase in the serum TSH concentration, which would otherwise be diagnostic of primary hypothyroidism. In such circumstances, a depressed rather than an elevated serum rT3 level may be helpful in the detection of coexistent primary thyroid failure.

Pharmacologic doses of glucocorticoids induce a prompt decline in serum T4 and T3 concentrations in thyrotoxic patients with Graves' disease.239  Amelioration of the symptoms and signs in such patients may also be accompanied by a decrease in the elevated thyroidal RAIU and a diminution of the TSH receptor antibody titer.325,332  This effect of glucocorticoids may be due in part to its immunosuppressive action since it has been shown that administration of dexamethasone to hypothyroid patients with Hashimoto's thyroiditis causes an increase in the serum concentration of both T4 and T3.333

 

 

Iodinated contrast It is estimated that in the US in the past year more than 80 million CT scans were performed and more than half of those utilized iodinated contrast.  Whether low or high ionic strength, low or high osmolality, all of these agents contain large amount of iodine ranging from 320 – 370 mg/ml.   In a prospective study, 2.6% of adults receiving contrast developed hyperthyroidism although many of theses cases were transient. 333a   In a study of hospitalized elderly patients with hyperthyroidism, 23% of them had a contrast CT performed in the preceeding months. 333b  When Alexander et al examined a data base of 4,500,000 patients, they found that the likelihood of developing hyperthyroidism within two years of being euthyroid was doubled by having a recent contrast CT. 333c    In a small study, pretreatment with thionamides reduced the incidence and severity of hyperthyroidism but did not always prevent it. 333d  Since many episodes of hyperthyroidism after iodinated contrast are transient, mild and asymptomatic, this approach may only be appropriate for patients who had more severe episodes.   Other options include avoidance of iodinated contrast and definitive treatment of any underlying thyroid disorder after the patient has recovered.   In a study of newly diagnosed hypothyroidism in children, the risk was increased nearly three fold by recent administration of iodinated contrast, 333e  while in adults, Kornelius et al found the risk was doubled. 333f

 

Ipodate and Iopanoic acid The principal effect of these iodine-containing radiologic contrast media is inhibition of T4 to T3 conversion by inhibiting both Type I and Type II 5’-deiodinase. In fact, they may be the most potent of all agents known to interfere with this step of iodothyronine metabolism. A triiodo-and a monoamino-benzene ring with a proprionic acid chain appear to be required because iodinated contrast agents without this chemical structure have little or no effect.334  Several of these agents, namely, ipodate (Oragrafin) and iopanoic acid (Telepaque), are used for oral cholecystography.

A decrease in the rate of deiodination of the outer ring of thyronines causes a profound decrease in the serum T3 concentration and an increase in the rT3 and T4 levels.334,335  The serum T4 concentration may reach values well within the thyrotoxic range.334  These changes are accompanied by an increase in serum TSH secretion.290  The latter is particularly notable, if not characteristic of these agents, probably because of their potent inhibitory effect on T3 generation in pituitary tissue.58  These agents have been used to study the regulation of thyroid hormone action via the process of iodothyronine deiodination.58,336  Changes persist for at least two to four weeks after their administration.334

Ipodate and iopanoic acid also decrease the hepatic uptake of T4337 and inhibit T3 binding to its nuclear receptors.338  These effects reduce both symptoms and thyroid hormone levels even when thyrotoxicosis occurs in settings where ongoing synthesis would be minimal such as thyrotoxicosis secondary to thyroid hormone ingestion338a , or sub-clinical hypothyroidism. 338b The antithyroidal effect of the iodine present in these agents is believed to be responsible for the falling T4 level and some of the amelioration of the symptoms and signs of thyrotoxicosis when they are administered to patients with Graves' disease338,338c,338d ,

 

Amiodarone. Most changes in thyroid function observed during the administration of this drug are similar to those seen with iodine-containing contrast agents. They include a marked decrease in serum T3, an increase in rT3, and a more modest elevation in the T4 concentration.241,339  Basal and TRH-stimulated TSH levels are increased. The principal mechanism of action is believed to be inhibition of both Type I and Type II 5’-deiodinase resulting in a marked reduction of T3 generation from T4. Amiodarone may reduce the entry of thyroid hormone into tissues339a, may reduce the binding of thyroid hormones to the receptor339b and may antagonize the effects of thyroid hormone at the cellular level.339c,339d The drug is used as an antianginal and antiarrhythmic agent and the bradycardia that almost invariably occurs when the drug is used in high doses, may suggest the presence of hypothyroidism.340

Amiodarone contains 37% iodine by weight. The major effects on thyroid function appear to be the result of its structural resemblance to thyroid hormone rather than its iodine content. In contrast to the typical alterations of thyroid hormone function, the more uncommon occurrence of frank hypothyroidism or thyrotoxicosis are products of the excess iodine released from the drug. The overall incidence of amiodarone induced thyroid disease is higher in areas of mild iodine deficiency340  as is the relative incidence of the thyrotoxic as compared to the hypothyroid form. 340 The iodine dependence of both of these diseases is confirmed by the improvement of both with the use of perchlorate to discharge iodine from the thyroid gland. 340a, 340b

Amoidarone induced thyrotoxicosis has been identified as having two main types; type 1 usually coccuring with underlying thyroid abnormalities and type 2 in normal glands with small goiters. 340a, 340b   Type 1 is more common in area of iodine deficiency. Early onset of thyrotoxicosisis is more typical for Type 1 and later onset with Type 2, but either form may present after amiodarone was discontinued. 340c, 340d Type 1 is associated with increased blood flow while hypervasculaity is absent in Type 2. Radioacitve iodine uptake may be low-normal or normal in Type 1 (especially in areas with iodine deficiency) and is low in Type 2. Type 1 is treated with thionamides but patients may be realtively resistant while patients with Type 2 respond to glucocorticoids. Some patient will present with a mixed form. Surgery may be used in cases refractory to medical therapy. 340e

Measurement of serum TSH, remains the most useful test in the differential diagnosis of hypothyroidism or thyrotoxicosis in amiodarone treated patients but the mild TSH elevation seen in euthyroid patients may make the diagnosis of mild to moderate hypothyroidism more difficult. If hypothyroidism is suspected, it is appropriate to obtain a measurement of the serum rT3 concentration. The absence of an elevated serum rT3 level in a patient receiving amiodarone suggests the patient is hypothyroid.

 

Diphenylhydantoin (Dilantin). Diphenylhydantoin (DPH) (Fig. 5-3) competes with thyroid hormone binding to TBG.228,229  This effect of DPH and diazepam, a related compound, has been exploited to study the conformational requirements for the interaction of thyroid hormone with its serum carrier protein229,341  It appears that the angle formed between the two phenyls and the hydantoin group of DPH is nearly identical to that formed between the two phenyls linked by an ether bond in T4.229  Although the affinity of DPH for TBG is far below that of T4, when used in therapeutic doses the serum concentration achieved is high enough to cause a significant occupancy of the hormone-binding sites on TBG. This effect of DPH is only partly responsible for the decrease in the total concentration of T4 and T3 in serum.

DPH accelerates the conjugation and clearance of T4 and T3 by the liver and probably enhances the conversion of T4 to T3.247,342  The net result is a decrease in the serum concentration of T4 and rT3 and, less consistently, that of T3343,344,344a,344b because the enhanced degradation of T3 is compensated for by an increase in its generation from T4. Yet, basal TSH- and TRH-stimulated values remain within the normal range343,344,344a,344b or slightly elevated.235,345  Calculated indices of FT4 are usually reduced, but the FT4 measured by dialysis is normal.247,343

Both DPH and diazepam are commonly used in clinical practice, the former most commonly as an anticonvulsant and the latter as an anxiolytic. Reduced serum levels of thyroid hormone in patients having therapeutic blood levels of DPH should not be viewed as indicative of thyroid dysfunction unless the TSH level is elevated. Treatment with T4 in such patients with a low T4 and normal TSH did not alter parameters of cardiac function or symptoms which might have been considered indicative of hypothyroidism.344b    DPH therapy may increase the required dose of thyroid hormone replacement in athyreotic individuals.346 .

 

Phenobarbital. Chronic administration of phenobarbital to animals induces increased binding of thyroid hormone to liver microsomes and increased deiodinating activity.248,249,347,347a  Phenobarbital administration reduces the biologic effectiveness of the hormone by diverting it to microsomal degradative pathways. In humans, phenobarbital augments fecal T4 clearance by nearly 100%,348 but serum T4 and FT4 levels remain near no rmal because of compensatory increases in T4 secretion. It is not apparent that barbiturates have an important effect on thyroid mediated metabolic action in normal humans, but it may potentate the effects of dilantin or carbamazapine. 348a  The augmented hepatic removal of T4 induced by phenobarbital lower the absolute T3 disposal by nearly 25%, increase T4 clearance, and lower T4 and FT4I in patients with Graves' disease but does not produce a clinical response.348

 

Propranolol. Propranolol, a ß-adrenergic blocker, is commonly used as an adjunct in the treatment of thyrotoxicosis Propranolol is usually used in the treatment of cardiac arrhythmias, angina and hypertension. Information regarding its effects on thyroid hormone action, and application in the symptomatic treatment of thyrotoxicosis is found in Chapters 3 and 11, respectively.

Propranolol does not affect the secretion or overall turnover rate of T4, nor TSH release or its regulatory mechanisms.349,350  A small to moderate lowering effect on serum T3 has been reported in euthyroid subjects as well as in patients with hyperthyroidism or with myxedema under L-T4 replacement therapy.243,245,351,352  Reciprocal increases in serum rT3 and 3',5'-T2 levels have also been reported.352  Such data, combined with the finding by some investigators of minimal increases in serum T4245 levels, suggest a mild blocking effect of this drug on the 5'-deiodination of iodothyronines. This effect does not appear to be related to the ß-adrenergic-blocking action of propranolol, since other ß-blocking agents do not share the deiodinase-blocking property and yet are effective in treating symptomatic thyrotoxicosis.353,354 The beneficial effects include the reduction of tachycardia, anxiety, and tremor 355-357 although the metabolic effects of thyrotoxicosis remain unaffected.

 

Reserpine. Reserpine formerly had wide use as an antihypertensive agent but has been replaced by more effective agents. Reserpine alters the manifestations of thyrotoxicosis by reducing anxiety, tachycardia, and tremulousness.358  This effect may arise from depression of autonomic centers or possibly from depletion of catecholamines in the peripheral tissues.359  Reserpine may depress the formation of iodotyrosines in thyroid tissue in vitro, but this action does not seem to be important clinically. Reserpine does not alter the results of thyroid function tests other than the BMR.358

 

Nitrophenols. 2,4-Dinitrophenol (Fig. 5-3) elevates the BMR, lowers the serum concentration of T4, accelerates the peripheral metabolism of T4, and depresses the thyroidal RAIU and secretion.275,360,361  The action is probably complex. The drug stimulates the metabolism by uncoupling oxidative phosphorylation in mitochondria.362  T4 in vitro also uncouples oxidative phosphorylation. Part of the effect of dinitrophenol may be to mimic the action of thyroid hormone on hypothalamic or pituitary receptor control centers; this effect would account for the diminished thyroid activity. Dinitrophenol also displaces thyroid hormone from T4-binding serum proteins.227  This action could lower the total hormone concentration in serum but should have no persistent effect on thyroid function. Dinitrophenol increases biliary and fecal excretion of T4, and this action largely accounts for the rapid removal of hormone from the circulation.363  Deiodination of T4 is also increased.364  Both of these effects may be related to displacement of hormone from TTR or to changes in metabolism of hormone in the liver.

2,4-Dinitrophenol does not share some of the most important properties of T4. It cannot initiate metamorphosis of tadpoles365 or provide a substitute for hormonal therapy in myxedema.

 

Dopaminergic Agents. It is generally accepted that endogenous brain dopamine plays a physiologic role in regulating TSH secretion via an effect on the hypothalamic-hypophyseal axis.252,366,367  Dopamine exerts a suppressive effect on TSH secretion and can be regarded as antagonistic to the stimulatory action of TRH at the pituitary level.284,287,367  Much of the information regarding the role of dopamine on the control of TSH secretion in humans has been derived from observations made during the administration of agents with dopamine-agonistic and -antagonistic activity (see Table 5-4 and Chapter 4).

Dopamine infusion is commonly used in the care of acutely ill hypotensive patients. It lowers the basal serum TSH level in both euthyroid and hypothyroid patients and blunts its response to the administration of TRH.252,284,287,368,368a

L-dopa, the precursor of dopamine, used in the treatment of Parkinson's disease and as a test agent in the diagnosis of pituitary diseases, also suppresses the basal and the TRH-stimulated serum TSH level in euthyroid subjects as well as in patients with primary hypothyroidism285,288,368b (Fig. 5-5). Metoclopropamide, a dopamine antagonist used as a diagnostic agent and in the treatment of motility disorders, increased TSH secretion. 368c

A similar effect has been observed during the administration of 2-brom-ergocryptine (bromocryptine), a dopamine agonist used in the treatment of some pituitary tumors and to suppress lactation during the puerperal period. Although the agent has been shown definitely to diminish the high serum TSH levels in patients with primary hypothyroidism,286 a significant inhibitory effect on TRH-induced TSH secretion has not been clearly demonstrated,369,370

The exact mechanism whereby dopaminergic drugs inhibit pituitary TSH secretion remains unknown, although a direct interaction with pituitary receptors has been suggested.371  While some authors have cautioned that prolonged infusion of dopamine may induce secondary hypothyroidism and worsen the prognosis of severely ill patients,372 there is no evidence that chronic treatment with dopaminergic drugs induces hypothyroidism in less critically ill patients.288 These drugs have been used with variable success in the treatment of some rare pituitary-induced thyrotoxicoses.373,374  When measurements of the basal or stimulated serum TSH levels are used in the differential diagnosis of primary and secondary hypothyroidism, the concomitant use of drugs with dopamine-agonistic or -antagonistic activity should be taken into account in the interpretation of results.

 

Alterations of Immunity

A number of drugs including interferon and lithium affected thyroid function either in part or completely by inducing thyroid immunity.  In the past few years a number of agents have been developed to treat cancer and multiple sclerosis by altering immune regulation.  Unintended side effects or these drugs has been the development of hyperthyroidism from Graves’ disease or thyroiditis and hypothyroidism as a consequence of autoimmune hypophyisitis or chronic thyroiditis.

 

 

Interferon and Interleukin These cytokines have been associated with the development of both hypothyroidism and thyrotoxicosis. 375-379 The overall rate of thyroid dysfunction induced by these agents is about 6%.379a They are used in the treatment of infectious diseases such as hepatitis, as well as malignancies including melanoma and renal cell carcinoma .   Acute administration has been used as a model of illness as the effects are similar; interferon-a leads to a decrease in T3 an increase in rT3 and a fall In TSH. 380 In a group of euthyroid HIV infected patients, however, short term administration of interleukin-2 was observed to lead to an increase in TSH, T3, T4 and free T4. 381

Cytokine induced thyroid disease appears to be immune mediated. The incidence is much greater in females and in patients with positive anti-peroxidase and anti-TPO antibodies prior to the initiation of therapy. 375-377 During therapy, patients who were antibody positive may have a rise in titer, while antibody positivity may develop in previously negative patients.375 In patients treated for hepatitis with interferon, the incidence of thyroid disease is much higher in those with Hepatitis C than those with Hepatitis B.375 The thyrotoxicosis often occurs as a manifestation of a destructive thyroiditis.376-377 In many patients, the thyroid disease resolves within several months after stopping the cytokine therapy. 375,377

 

Anti CD 52 Antibody  Alemtuzumab This monoclonal antibody reacts against CD52, a glycoprotein that is expressed on B cells and CD4+ T cells.  It is approved by the FDA for the treatment of multiple sclerosis.  It is initially dosed daily for five days and then a second course is given for three days, one year later.

Thyroid disease has been noted to occur in a third of treated patients with some cases seen within weeks of initiating treatment and most cases seen within the first three years but cases have been seen up to seven years after starting treatment. 376a,376b,377a

The most frequently observed disorder has been Graves’ disease which affects almost a quarter of all treated patients. 376a,376b,377a Most affect patients develop antibodies.  Most cases are overt and some patients have experienced significant ophthalmopathy.  While as many as 35% of these patients have been reported to have spontaneous resolution, most have been treated medically. Hypothyroidism develops in 5-7% of patients.  Most of these patients develop anti-thyroid antibodies and the deficits are usually permanent.  Less than 5 % of patients will develop typical painless thyroiditis with transient hyperthyroidism sometimes followed by hypothyroidism.  376a,376b,377a

 

 

Check Point Inhibitors

 

These agents target systems that normally act to limit activation of the immune system but are utilized by cancer cells to block immune mediated destruction.  Use of these agents allows the activated immune cells to kill tumor cells. . 377b- 377d   The same mechanism of action, however, leads to activation of immune mediated damage to other cells including skin, liver and thyroid cells. . 377b- 377d

 

Antibody against CTLA-4 ipilimumab Co-stimulation via HLA and B7 expressed on antigen presenting cells (and tumor cells) interacting respectively with T cell receptors and CD 28 expressed on T-cells leads to T-cell activation.  Activated T-cells then express CTLA-4 that competes with CD 28 for B7 binding and thus reduces T cell activation.   The monoclonal antibody, ipilmumab, specifically binds to CTLA-4 markedly enhancing T cell activation. 377b- 377d This promotes immune-mediated destruction of tumor cells.     This agent was approved by the FDA for the treatment of melanoma in 2011.  Usual dosing is every three weeks for four doses. The drug also remains under investigation for treating other tumors.

The most common associated autoimmune disorder affecting the thyroid has been the development of hypophyisitis. 377b- 377f   Prior to the introduction of this agent, autoimmune hypophyisitis was typically seen in women in late pregnancy or post-partum, while hypophysitis secondary to ipilimumab has been seen almost solely in men.   Onset has been within weeks of initiating therapy until almost two years later. 377b- 377f     Most patients present with systemic symptoms while some present with headaches or visual symptoms.  MRI abnormalities are common and include pituitary enlargement and stalk thickening.  Similar to post-partum hypophyisitis, the pituitary-adrenal axis is most commonly affected, but most deficits are permanent.  377b- 377f

Hypothyroidism occurs in 4-6% of treated patients from 2 months to 3 years after starting treatment.  Fatigue is the most common presenting symptom.  The hypothyroidism is usually permanent.  Some 1-3 % of treated patients will develop typical painless, thyroiditis with transient hyperthyroidism. Onset is usually two to four months after starting treatment. 377b- 377f  Hyperthyroidism is sometimes followed by hypothyroidism.  The package insert advises checking thyroid function prior to starting therapy, before each dose and as “clinically indicated’” but does not recommend checking cortisol or ACTH.

 

Antibodies Against Programmed Death Receptor Ligand (PDL-1)  Pembrolizumab and Nivolumab Recognition of tumor cells via MHC/T cell receptor interaction leads to T cell activation and interferon production which stimulates tumor cell production of the PD-1 ligand.  This then binds with the PD-1 receptor on the T cell and inhibits activation.   Dendritic cells also express this ligand to inhibit T-cell activation. Pembrolizumab and Nivolumab are monoclonal antibodies directed against the PD1 ligand that act to increase T-cell activity and thus promote immune-mediated destruction of tumor cells. 377d, 377g       These drugs were approved by the FDA in 2014 for the treatment of melanoma.   Pembrolizumab  is administered intravenously every 3 weeks.  Nivolumab is administered intravenously every 2 weeks.  These drugs also remains under investigation for treating other tumors.

Both agents may lead to the development of immune mediated hypothyroidism, thyroiditis and hypophysitis with hypothyroidism being the most common problem. 377g-377i   Hypothyroidism may occur within weeks of starting therapy or may not occur until after a year.   It is usually permanent. 377g-377i   In contrast, reported cases of hyperthyroidism from thyroiditis have occurred from 2 weeks to 5 months after initiating therapy and always resolves. 377g-377i

For pembrolizumab the occurrence of these complications has been reported as 8% for hypothyroidism, 2-3% for thyroiditis and 0.5% for hypophyisitis.   For nivolumab the rate of occurrence of these complications have been reported as 4-8% for hypothyroidism and 1-3% for thyroiditis.  For both drugs, it is recommended to check thyroid function tests prior to starting therapy and periodically afterwards.

Tyrosine Kinase Inhibitors Sunitinib maleate an oral tyrosine kinase inhibitor used in the treatment of renal cell carcinoma and gastrointestinal stromal tumors has also been associated with the development of hypothyroidism.381a In two studies, an elevated TSH has been seen in over 50% of patients treated with sunitinib. 382.383  In a prospective study, this was persistent in 36% and transient in 17%. 382  The mechanism remains unknown. 382a Antiperoxidase activity was demonstrated in vitro382, but other mechanisms include induction of destructive thyroiditis, 383  reduction of vascularity ot the gland, 383b and enhanced apoptosis. 383c

 

The thyroid effects are seen with other TK inhibitors as well although the frequency and severity of the effect may vary. Hypothyroidism has also been seen with Sorafenib, but the rate is about 1/3 of that seen with sunitinib.383d,383e In a few patients, transient thryotoxicosis has preceeded the hypothyroidism consistent with a destructive thyroiditis. 383d There is also evidence for enhanced thyroid hormone metabolism attributed to increased Type 3 deiodination. 383f   This would also explain the need for increased thyroid hormone doses in athyreotic thyroid cancer patients.

 

Imatinib mesylate is a selective tryrosine kinase inhibitor used in the treatment of chronic mylogenous leukemia (CML) and other malignancies. Thyroidectomized patients being treated with imatinib were noted to have a rise in TSH and a fall in serum T4 levels which responded to an increase in the T4 dose, suggesting enhanced metabolism of thyroid hormone 384 but changes have also been seen in euthyroid patients. 384a In most, cases, these changes were transient (74%) but have persisted in others. 384a   Even higher rates of thyroid dysfunction have been seen with the newer agents nilotinib (55%) and dasatinib (75%). 384a   As with the other TK inhibitors some patients have had thyrotoxicosis and some have developed antithyroid antibodies. 384a

 

Retinoids Bexarotene is a retinoid which is specific for the retinoid X- receptor and is used for the treatment of lymphomas and other malignancies. Therapy has been reported to produce central hypothyroidism 376a,376b,385, and a single dose, leads to a decrease in T3, T4 and TSH. 386  In addition to suppression of TSH synthesis and secretion, bexarotene also increases the peripheral metabolism of thyroid hormone by a nondeiodinase mediated pathway. 387

 

TSH Receptor Agonists and Antagonists A number of small molecules that interact with the TSH receptor were identified and characterized to select compounds which could behave as TSH receptor agonists or antagonists. 387 These were then further modified to increase their activity. Of note, these molecules do not bind to the TSH ligand binding region but rather to the serpentine trans-membrane region of the receptor. These compounds have multiple potential uses including use as imaging agents for thyroid cancer and Graves’ ophthalmopathy and as therapeutic agents for patients with Graves’ or thyroid cancer.

 

A TSH receptor agonist has been developed that when added to primary cultures of human thyrocytes increased messenger RNA expression for thyroglobulin, the sodium-iodide symporter, thyroid peroxidase (TPO), and deiodinase 2 similar to TSH. 388 When administered orally to mice, this compound increased radioactive iodine uptake in the thyroid and serum T4. 388 As an oral agent, this compound could potentially be used for imaging and treatment of thyroid cancer rather than parenteral rTSH or thyroid hormone withdrawal.

 

A TSH receptor antagonist has also been developed that has activity both in cells overexpressing the human TSH receptor and in primary cultures of human thyrocytes. 389 The compound reduces both basal and TSH stimulated cAMP production. Recently it was demonstrated in cultured human thyrocytes to reduce basal TPO mRNA expression and to antagonize the effect of sera from Graves’ patients to induce TPO mRNA expression. 389 As an oral agent, this compound could potentially be used for imaging, to treat Graves’ patients or to suppress thyroid cancer without requiring use of supraphysiologic T4 doses.

 

 

Thyronamines Thyronamines are small molecules identical to thyroxine, triiodothyronine and all of the deiodinated thyroid hormone metabolites except that they lack a carboxyl moiety at the amino terminus (ethylamine rather than alanine group). 390 Each compound is identified similar to the corresponding thyroid hormone or metabolite as TX(AM) where X is the number of iodine molecules and ranges from zero, T0(AM), to four, T4(AM).   Two of these compounds, [3-T1(AM) and T0(AM)], have been identified by liquid chromatography-tandem mass spectrometry as naturally present in small amounts in tissues and sera from hamsters, mice and rats. 390 They have been detected in both cardiac and brain tissue. 390    No published report has confirmed the presence of any of these compounds in humans. It has been speculated that some of these compounds could be directly produced by the decarboxylation of T4 or T3, but this has never been demonstrated. These compounds can be deiodinated, in vitro, by Deiodinases 1, 2 and 3. 390

 

These compounds can bind to a number of receptors and 3-T1(AM) binds strongly to APO B 100 in serum. Despite the structural similarities to thyroid hormones, the thyronamines do not bind to nuclear thyroid hormone receptors and they do not alter T3 binding to these receptors. 390  Several thyronamines bind to the beta-adrenergic receptor, but any effects on cAMP signaling remain unclear. 390  There is conflicting evidence regarding the ability of these compounds to signal via the trace amine associated receptor 1 (TAAR-1) or via the Alpha2A adrenergic receptor. 390 There is also conflicting evidence about the ability of these compounds to alter intracellular signaling via the cAMP or tyrosine phosphorylation or dephosphorylation pathways. 390

 

There are no known physiologic actions of any of these compounds.   In animals, several of these compounds have been found to have pharmacologic activity both in vitro and after intraperitoneal or intraventicular injection. These observations include a reduction in cardiac contractility and rate, a reduction in the metabolic rate, a reduction in fat mass and the development of hypothermia, ketonuria and hyperglycemia. 390 Many of these activities are noted within minutes after injection and resolve after a few hours but the development of ketonuria and the reduction of fat mass occur later and persist longer. 390  Potential therapeutic uses of these compounds are being evaluated in animal models. The ability of these compounds to induce hypothermia, has been shown to decrease infarct size when they were administered 2 days before or 1 hour after the induction of a stroke in an animal model. 390

 

Metformin Metformin is a biguanide used in the treatment of diabetes mellitus as well as insulin resistance and polycystic ovary syndrome. In four patients with hypothyroidism on stable thyroxine therapy, TSH levels became markedly reduced with either no change in serum thyroid hormone levels or despite a reduction in the T4 dose and serum thyroid hormone levels suggesting a direct suppression of TSH release. 388 Subsequent studies have reported mixed effects, but a meta-analysis conluded that TSH alterations are seen in both overt and sub-clinical hypothryoidism but not in euthyroid patients suggesting an effect to suppress TSH that is not seen when the thyroid gland is able to respond to any change in TSH. 388b  

 

Biotin Biotin is a B vitamin that acts as a cofactor for carboxylase enzymes involved in gluconeogenesis and fatty acid synthesis.  It is produced by gut bacrteria and normal daily intake is 35-350 mcg daily.  It is used in the treatment of biotinidase deficiency and proprionic acidemia and as a supplement for TPN.  It is frequently used by individuals in doses of 5,000 to 10,000 mcg daily as a supplement to improve hair and nail growth and to treat hair and nail disorders

 

Many laboratory platforms for the measurement of fT4, fT3 TSH and thyroglobulin depend on the strong binding of biotin and strep avidin.  If patients ingest biotin in doses of 5,000 to 10,000 mcg prior to blood being drawn for these analytes, measurements of fT4 and fT3 will be falsely high and thyroglobulin and TSH will be falsely low as biotin interferes in the assays. . 397,398  The combination of a high fT4 and low TSH mimics hyperthyroidism. 397-400   These effects correspond to the blood level of biotin with a peak effect seen several hours after ingestion and potentially even lasting until the next day.  398,400   Variable times between ingestion and blood measurements can results in confusing variations in these measurement not corresponding to patents clinical status.  Confirmation of this effect can be made by measuring several hours after ingestion and after abstaining for 48 hours or by re-measuring in an assay not utilizing biotin. 397-400    This effect is not limited to thyroid hormone measurements but have also been reported for PTH, DHEA-sulfate, estradiol and ferritin. 398

 

 

SUMMARY

This chapter considers the effects of various environmental factors, drugs and chemicals, and nonthyroidal diseases on thyroid function.

In animals, cold exposure causes a prompt increase in TSH secretion, which gives rise to thyroid hormone release and leads to thyroid gland hyperplasia. Part of this effect is due to an apparent increase in the need for thyroid hormone by peripheral tissues and to an excessive rate of hormone degradation and excretion. In humans, hypothermia causes a dramatic TSH secretion in the newborn, but this response is lost after the first few years of life. Exposure to heat has an opposite effect, although of lesser magnitude. A small seasonal variation in serum thyroid hormone levels that follow this general pattern has been reported.

Simulated altitude and anoxia depress thyroid hormone formation in rats, but in humans serum T4 and T3 concentrations, T4 degradation, and oxygen consumption are at least temporarily augmented by high altitude.

Starvation has a profound effect on thyroid function, causing a decrease in serum T3 concentration and a reciprocal increase in rT3 level. These changes are due to a selective inhibition of the 5'-monodeiodination of iodothyronines by peripheral tissues. Reduction in carbohydrate intake rather than total calorie deprivation appears to be the determinant factor. These alterations in thyroid function are believed to reduce the catabolic activity of the organism and thus to conserve energy in the face of decreased calorie intake. Chronic malnutrition is accompanied by similar changes. Overfeeding has opposite although transient effects.

Physical and emotional stresses can have variable and opposite effects. Increased thyroid hormone secretion and serum levels have been observed in stressed animals and in acute psychiatric patients on admission. The physical stress of surgery causes a prompt decrease in the serum T3 concentration, probably as a consequence of decreased T3 neogenesis. This effect of surgery cannot be fully explained on the basis of increased adrenocortical activity or calorie deprivation.

Many minerals alter the synthesis of thyroid hormone, mainly through their interference with iodide concentration and binding by the thyroid gland. The action of iodine is only briefly covered here since it is discussed in Chapters 2 and 13. Calcium, nitrate, bromine, rubidium, and fluorine are allegedly goitrogenic. Lithium carbonate, used in the usual doses for the treatment of affective disorders, can produce goiter in susceptible persons. It inhibits iodide binding and hormonal release from the thyroid gland, probably through a synergistic action with iodide.

Numerous dietary goitrogens, including cyanogenic glucosides, thioglucosides, thiocyanate, and goitrin, are present in a wide variety of foods, and are believed to contribute to the occurrence of endemic goiter in some areas of the world. Monovalent anions such as thiocyanate and perchlorate inhibit iodide transport into the thyroid and cause goiter.

Thionamide drugs such as PTU and the related compound, methimazole, inhibit thyroid peroxidase and thus prevent thyroid hormone synthesis. In addition, PTU but not methimazole inhibits the conversion of T4 to T3 in peripheral tissues. Under appropriate circumstances, sulfonamides, sulfonylureas, salicylamides, resorsinol, amphenone, aminoglutethamide, antipyrine, aminotriazole, amphenidone, 2,3-dimercaptopropanolol, and phenylbutazone have antithyroid action.

A growing list of drugs and diagnostic agents have been found to affect thyroid economy by modulating the regulation of the hypothalamic-pituitary-thyroid axis, as well as by interfering with thyroid hormone transport, metabolism, excretion, and action. Some drugs, such as salicylates, diphenylhydantoin, and glucocorticoids, act at several levels. Several compounds, most notably estrogens, diphenylhydantoin, diazepam, heparin, halophenate, fenclofenac, and some biologically inactive thyroid hormone analogs compete with binding of thyroid hormone to its carrier proteins in serum. The only consequence of drugs affecting hormone transport is a decrease or increase in the concentration of total but not free hormone in serum.

Glucocorticoids, drugs such as propranolol, and amiodarone and some iodinated contrast media inhibit the extrathyroidal generation of T3. The result is a decrease in serum T3 and an increase in rT3 concentrations, with a slight increase or no change in T4 values. Thyroid hormone disposal is accelerated by diphenylhydantoin and phenobarbital, which increase several of the pathways of hormone degradation, and by hypolipemic resins, which increase the fecal loss of hormone. Homeostasis is usually maintained by a compensatory increase in thyroid hormone secretion.

Some drugs act through inhibition or stimulation of TSH secretion. Most notable of the former effect are dopamine agonists such as L-dopa and bromocryptine, as well as some -adrenergic blockers, glucocorticoids, acetylsalicylic acid, and opiates. A variety of dopamine antagonists as well as cimetidine, clomifene, and spirolactone appear to increase TSH secretion. These compounds seem to interfere with the normal dopaminergic suppression of the hypothalamic-pituitary axis. Observed changes in TSH secretion are not associated with significant metabolic alterations. Some of the drugs have an apparent effect on TSH secretion through changes induced at the levels of the free and active forms of the thyroid hormone. A handful of drugs appear to block or antagonize the action of thyroid hormone on tissues. These drugs include guanethidine, propranolol, and dinitrophenol. Some drugs may induce autoimmune thyroid disease. Notably among these are lithium, interferon, interleukin, alemtuzumab. prembrolizumab and nivolumab.

The clinician should be thoroughly familiar with the effects of drugs, nonthyroidal illnesses, and other extraneous factors on thyroid function. These factors should all be taken into account in the differential diagnosis of primary thyroid disease.

 

 

REFERENCES

  1. Bernstein G, Oppenheimer JH: Factors influencing the concentration of free and total thyroxine in patients with nonthyroidal disease. J Clin Endocrinol Metab 26:195, 1966.

1a.          Silva JE, Larsen PR: Potential of brown adipose tissue type II thyroxine 5’-deiodinase as a local and systemic source of triiodothyronine in rats. J Clin Invest 76:2296-,1985.

1b.          Hackney AC, Feith S, Pozos R, Seale J: Effects of high altitude and cold exposure on resting thyroid hormone concentrations. Aviat Space Environ Med 66:325-9,1995.

1c.          Hackney AC, Hogdon JA, Hesslink R Jr, Trygg K: Thyroid hormone responses to military winter exercise in the Arctic region. Arctic Med Res 54:82-90,1995.

  1. Wilber JF, Baum D: Elevation of plasma TSH during surgical hypothermia. J Clin Endocrinol Metab 31:372-375, 1970.
  2. Fisher DA, Odell WD: Acute release of thyrotropin in the newborn. J Clin Invest 48:1670, 1969.
  3. Fisher DA, Oddie TH: Neonatal thyroidal hyperactivity. Response to cooling. Am J Dis Child 107:574, 1964.
  4. Hershman JM, Read DG, Bailey AL, Norman VD, Gibson TB: Effect of cold exposure on serum thyrotropin. J Clin Endocrinol Metab 30:430, 1970.
  5. Nagata H, Izumiyama T, Kamata K, et al: An increase of plasma triiodothyronine concentration in man in a cold environment. J Clin Endocrinol Metab 43:1153, 1976.
  6. Golstein-Golaire J, Van Haelst L, Bruno OD, Leclercq R, Copinschi G: Acute effects of cold on blood levels of growth hormone, cortisol, and thyrotropin in man. J Appl Physiol 29:622, 1970.

7a.          Reed HL: Environmental Influences on thyroid hormone regulation. pp259-265 in Werner and Ingbar’s The Thyroid , Seventh Edition   Braverman LE, Utiger RD eds. JB Lippincot Philadelphia 1996.

7b.          McCormack PD, Reed HL, Thomas JR, Malik MJ: Increase in rT3 levels observed during extended Alaskan field operations of Naval personnel. Alaska Med 38:89-97,1996.

7c.          Van Do N, Mino L., Merriam G, LeMAr H, Case HS, Palinkas LA, Reedy K, Reed HL: Elevation in serum thyroglbulin during prolonged Antarctic residence: Effect of thyroxine supplement in the polar 3,5,3’-triidothyronine syndrome.   J Clin Endocrinol Metab   89:1529-1533,2004

7d           Reed HL, Silverman ED, Shakir KM et al: Changes in serum triodothyronine (T3) kinetics after prolonged antarctic residence: the polar T3 syndrome. J Clin Endocrinol Metab 70:965-,1990

  1. Balsam A, Sexton FC: Increased metabolism of iodothyronines in the rat after short-term cold adaptation. Endocrinology 97:385, 1975.
  2. Bernal J, Escobar del Rey F: Effect of the exposure to cold on the extrathyroidal conversion of L-thyroxine to triiodo-L-thyronine, and on intramitochondrial -glycerophosphate dehydrogenase activity in thyroidectomized rats on L-thyroxine. Acta Endocrinol 78:481, 1975.

9a.          Tsukahara F, Uchida Y, Ohba K, Nomoto T, Muraki T: Defective stimulation of thyroxine 5’-deiodinase activity by cold exposure and norepinephrine in brown adipose tissue of monosodium glutamate-obese mice. Horm Metab Res 29:496-500,1997.

9b.     Margarity M, Valcana T Effect of cold exposure on thyroid hormone metabolism and nuclear bindng in rat brain. Neurochem Res 24:423-6, 1999

  1. DuRuisseau JP: Seasonal variation of PBI in healthy Montrealers. J Clin Endocrinol Metab 25:1513, 1965.
  2. Smals AGH, Ross HA, Kloppenborg PWC: Seasonal variation in serum T3 and T4 levels in man. J Clin Endocrinol Metab 44:998, 1977.
  3. Panda JN, Turner CW: Effect of thyroidectomy and low environmental temperature (4.4C) upon plasma and pituitary thyrotrophin in the rat. Acta Endocrinol 54:485, 1975.
  4. Emerson CH, Utiger RD: Plasma thyrotropin-releasing hormone concentrations in the rat. J Clin Invest 56:1564, 1975.
  5. Andersson B: Hypothalamic temperature and thyroid action. C. F. S. G. 18 (eds), Brain-THyroid Relationships, pp. 35-50,1964.
  6. Montoya E, Seibel MJ, Wilber JF: Thyrotropin-releasing hormone secretory physiology: studies by radioimmunoassay and affinity chromatography. Endocrinology 96:1413, 1975.
  7. Szabo M, Frohman LA: Suppression of cold-stimulated thyrotropin secretion by antiserum to thyrotropin-releasing hormone. Endocrinology 101:1023, 1977.
  8. Hefco E, Krulich L, Illner P, Larsen PR: Effect of acute exposure to cold on the activity of the hypothalamic-pituitary-thyroid system. Endocrinology 97:1185, 1975.
  9. Jobin M, Ferland L, Coté J, Labrie F: Effect of exposure to cold on hypothalamic TRH activity and plasma levels of TSH and prolactin in the rat. Neuroendocrinology 18:204, 1975.
  10. Melander A, Rerup C: Studies on e thyroid activity in the mouse. Acta Endocrinol 58:202, 1968.
  11. Yamada T, Kajihara A, Onaya T, Kobayashi I, Takemura Y, Shichijo K: Studies on acute stimulatory effect of cold on thyroid activity and its mechanism in the guinea pig. Endocrinology 77:968, 1965.
  12. Balsam A Leppo, L.E.: Augmentation of the peripheral metabolism of L-triiodothyronine and L-thyroxine after acclimation to cold. Multifocal stimulation of the binding of iodothyronines by tissues. J Clin Invest 53:980, 1974.
  13. Galton VA, Nisula BC: Thyroxine metabolism and thyroid function in the cold-adapted rat. Endocrinology 85:79-, 1969.

22a.        Puigserver p, Wu Z, Park CW, Graves R, Wright M, Spiegelman BM. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 92:892-39,1998.

  1. Epstein Y, Udassin R, Sack J: Serum 3,5,3'-triiodothyronine and 3,3',5'-triiodothyronine concentrations during acute heat load. J Clin Endocrinol Metab 49:677, 1979.
  2. Ljunggren JG, Klalner G, Tryselius M: The effect of body temperature on thyroid hormone levels in patients with nonthyroidal illness. Acta Med Scand 202:459, 1977.
  3. O'Malley BP, Davies TJ, Rosenthal FD: TSH responses to temperature in primary hypothyroidism. Clin Endocrinol 13:87, 1980.
  4. Rastogi GK, Malhotra MS, Srivastava MC, et al: Study of the pituitary-thyroid functions at high altitude in man. J Clin Endocrinol Metab 44:447, 1977.
  5. Moncloa F, Guerra-Garcia R, Subauste C, Sobrevilla LA, Donayre J: Endocrine studies at high altitude. I. Thyroid function in seal level natives exposed for two weeks to an altitude of 4,300 meters. J Clin Endocrinol Metab 26:1237, 1966.
  6. Surks MI, Beckwitt HJ, Chidsey CA: Changes in plasma thyroxine concentration and metabolism, catecholamine excretion, and basal oxygen consumption in man during acute exposure to high altitude. J Clin Endocrinol Metab 27:789, 1967.

28a.        Mordes JB, Blume FD, Boyer et al: High-altitude pituitary-thyroid dysfunction on Mount everest. N Engl J Med 308:1135-,1983.

28b.        Ramirez G, Herrera R, Pineda D, Bittle PA, Rabb HA, Bercu BB: The effects of high altitude on hypothalamic-pituitary secretory dynamics in man. Clin Endocrinol (oxf) 43:11-18,1995.

  1. Mulvey PF, Macaione JMR: Thyroidal dysfunciton during simulated altitude conditions. Fed Proc 23:1243, 1969.
  2. Surks MI: Effect of hypoxia and high altitude on thyroidal iodine metaoblism in the rat. Endocrinology 78:307, 1966.
  3. Surks MI: Effect of thyrotropin on thyroidal iodine metabolism during hypoxia. Am J Physiol 216:436, 1969.
  4. Pazo JH, Houssay AB, Davison TA, Chait RJ: On the mechanism of the thyroid hypertrophy in pinealectomized rats. Acta Physiol Lat Am 18:332, 1968.
  5. Singh DV, Turner CW: Effect of light and darkness upon thyroid secretion rate and on the endocrine glands of female rats. Proc Soc Exp Biol Med 131:1296, 1969.

33a.        Shavali SS, Haldar C: Effects of continuous light, continuous darkness and pinealectomy on pineal-thyroid-gonadal axis of the female Indian palm squirrel, Funambulus pennati.   J Neurol Transm 105:407-13,1998.

  1. Singh DV, Narang GD, Turner CW: Effect of melatonin and its withdrawal on thyroid hormone secretion rate of female rats. J Endocrinol 43:489, 1969.

34a.        Uchiyama M, Ishibashi K, Enomoto T, Nakajima T, Shibui K, Hirokawa G, Okawa M: Twenty-four hour profiles of four hormones under constant routine. Psychiatry Clin Neurosci 52:241-3,1998.

  1. Martino E, Seo H, Lernmark A, Refetoff S: Ontogenetic pattern of thyrotropin-releasing hormone-like material in rat hypothalamus, pancreas and retina: Selective effect of light deprivation. Proc Natl Acad Sci 77:4345, 1980.

35a.        Allan JS, Czeisler CA: Persistence of the circadian thyrotropin rhythmn under constant conditions and after light-induced shifts of circadian phase. J Clin Endocrionol Metab 79:508-,1994.

  1. Portnay GI, O'Brian JT, Bush J, al: The effect of starvation on the concentration and binding of thyroxine and triiodothyronine in serum and on the response to TRH. J Clin Endocrinol Metab 39:191-194, 1974.
  2. Merimee TJ, Fineberg ES: Starvation-induced alterations of circulating thyroid hormone concentrations in man. Metabolism 25:79, 1976.
  3. Carlson HE, Drenick EJ, Chopra IJ, Hershman JM: Alterations in basal and TRH-stimulated serum levels of thyrotropin, prolactin, and thyroid hormones in starved obese men. J Clin Endocrinol Metab 45:707, 1977.
  4. Azizi F: Effect of dietary composition on fasting-induced changes in serum thyroid hormones and thyrotropin. 27:935-942, 1978.
  5. Scriba PC, Bauer M, Emmert D, et al: Effects of obesity, total fasting and re-alimentation of L-thyroxine (T4), 3,5,3'-L-triiodothyronine (T3), 3,3',5'-L-triiodothyronine (rT3), thyroxine binding globulin (TBG), cortisol, thyrotrophin, cortisol binding globulin (CBG), transferrin, 2-haptoglobin and complement C'3 in serum. 91:629-643, 1979.

40a.        Alvero R, Kimzey L, Sebring N, Reynolds J, Loughran M, Nieman L, Olson BR: Effects of fasting on neuroendocrine function and follicle development in lean women. JCEM 83:76-80,1998.

40b.        Byerly LO, Heber D: Metabolic effects of triiodothyronine replacement during fasting in obese subjects. JCEM 81:968-76,1996.

  1. Vagenakis AG, Portnay GI, O'Brian JT, et al: Effect of starvation on e production and metabolism of thyroxine and triiodothyronine in euthyroid obese patients. J Clin Endocrinol Metab 45:1305, 1977.
  2. Suda AK, Pittman CS, Shimizu T, Chambers JB Jr.: The production and metabolism of 3,5,3'-triiodothyronine and 3,3',5'-triiodothyronine in normal and fasting subjects. J Clin Endocrinol Metab 47:1311, 1978.
  3. Stokholm KH: Decrease in serum free triiodothyronine, thyroxine-binding globulin and thyroxine-binding prealbumin whilst taking a very-low-calorie diet. Int J Obest 4:133, 1980.
  4. Balsam A, Ingbar SH: The influence of fasting, diabetes, and several pharmacological agents on the pathways of thyroxine metabolism in rat liver. J Clin Invest 62:415, 1978.
  5. Chopra IJ, Geola F, Solomon DH, Maciel RMB: 3',5'-diiodothyroxine in health and disease: Studies by a radioimmunoassay. J Clin Endocrinol Metab 47:1198-1207, 1978.
  6. Chopra IJ: A radioimmunoassay for measurement of 3'-monoiodothyronine. J Clin Endocrinol Metab 51:117-123, 1980.
  7. Pangaro L, Burman KD, Wartofsky L, et al: Radioimmunoassay for 3,5-diiodothyronine and evidence for dependence on conversion from 3,5,3'-triiodothyronine. J Clin Endocrinol Metab 50:1075-1081, 1980.

47a.   Van der Geyten S, Van Rompaey E, Sanders JP, Visser TJ, Kuhn ER, Darras VM : Regulation of thyroid hormone metabolism during fasting and refeeding in chicken. Gen Comp Endocrinol 116:272-80,1999

  1. Pittman CS, Shimizu T, Burger A, Chambers JB Jr.: The nondeiodinative pathways of thyroxine metabolism: 3,5,3',5'-tetraiodothyroacetic acid turnover in normal and fasting human subjects. J Clin Endocrinol Metab 50:712-716, 1980.
  2. Balsam A, Ingbar SH: Observations on the factors that control the generation of triiodothyronine from thyroxine in rat liver and the nature of the defect induced by fasting. J Clin Invest 63:1156, 1979.
  3. Chopra IJ: Alterations in monodeiodination of iodothyronines in the fasting rat: Effects of reduced nonprotein sulfhydryl groups and hypothyroidism. Metabolism 29:161, 1980.
  4. Gavin LA, McMahon FA, Moeller M: Dietary modification of thyroxine deiodination in rat liver is notmediated by hepatic sulfhydryls. J Clin Invest 65:943, 1980.
  5. Burman KD, Dimond RC, Harvey GS, et al: Glucose modulation of alterations in serum iodothyronine concentrations induced by fasting. Metabolism 28:291, 1979.
  6. Danforth E, Sims EAH, Horton ES, Goldman RF: Correlation of serum triiodothyronine concentrations with dietary composition. Diabetes 24:406, 1975.
  7. Harris ARC, Fang SL, Vagenakis AG, Braverman LE: Effect of starvation, nutriment replacement, and hypothyroidism on in vitro hepatic T4 to T3 conversion in the rat. Metabolism 27:1680, 1978.
  8. Burger AG, Berger M, Wimpfheimer K, Danforth E: Interrelationships between energy metabolism and thyroid hormone metabolism during starvation in the rat. Acta Endocrinol 93:322, 1980.
  9. Gardner DF, Kaplan MM, Stanley CA, Utiger RD: Effect of triiodothyronine replacement on the metabolic and pituitary responses to starvation. N Engl J Med 300:579, 1979.
  10. Silva JE, Dick TE, Larsen PR: The contribution of local tissue thyroxine monodeiodination to nuclear 3,5,3'-triiodothyronine in pituitary, liver and kidney of euthyroid rats. Endocrinology 103:1196, 1978.
  11. Cheron RG, Kaplan MM, Larsen PR: Physiological and pharmacological influences on thyroxine to 3,5,3'-triiodothyronine conversion and nuclear 3,5,5'-triiodothyronine binding in rat anterior pituitary. J Clin Invest 64:1402, 1979.

58a.        Diano S, Naftolin F, Goglia F, Horvath TL: Fating-induced increase in Type II iodothyronine deiodinase activity and messenger ribonucleic acid levels is not reversed by thyroxine in the rat hypothalamus. Edocrinology 139:2879-84,1998.

  1. Harris ARC, Fang SL, Azizi F, Lipworth L, Vagenakis AG, Braverman LE: Effect of starvation on hypothalamic-pituitary-thyroid function in the rat. Metabolism 27:1074, 1978.
  2. Morley JE, Russell RM, Reed A, Carney EA, Hershman JM: The interrelationship of thyroid hormones with vitamin A and zinc nutritional status in patients with chronic hepatic and gastrointestinal disorders. Am J Clin Nutr 34:1489, 1981.

60a.        von Haasteren GA, Linkels E, van Toor H, Klootwijk W, Kaptein E, de Jong FH, Reymond MJ, Visser TJ, de Greef WJ: Effects of long term food reduction on the hypothalamic-pituitary-thyroid axis in male and female rats. J Endocrinol 150:169-78,1996.

60b.        von Haasteren GA, Linkels E, Klootwijk W, van Toor H, Rondeel JM, Themmen AP, de Jong FH, Valentijn K, Vaudry H, Bauer K, et al: Starvation induced changes in the hypothalamic content of prothyrotropin-releasing hormone (proTRH) mRNA and the hypothalmic release of proTRH derived peptides:role of the adrenal gland. J Endocrinol 145:143-53,1995.

  1. Shambaugh GE III, Wilber JF: The effect of caloric deprivation upon thyroid function in the neonatal rat. Endocrinology 94:1145, 1974.
  2. DeGroot LJ, Coleoni AH, Rue PA, Seo H, Martino E, Refetoff S: Reduced nuclear triiodothyronine receptors in starvation-induced hypothyroidism. Biochem Bioophys Res Commun 79:173, 1977.

62a.        Tagami T, Nakamura H, Sasaki S, Miyoshi Y, Nakao K: Starvation-induced decrease in the maximal binding capacity for triiodothyronine of the thyroid hormone receptor due to a decrease on the receptor protein. Metabolism 45:970-3,1996.

  1. Schussler GC, Orlando J: Fasting decreases triiodothyronine receptor capacity. Science 199:686, 1978.
  2. Buergi V, Larsen PN: Nuclear triiodothyronine binding in mononuclear leukocytes in normal subjects and obese patients before and after fasting. J Clin Endocrinol Metab 54:1199, 1982.
  3. Jung RT, Shetty PS, James WPT: Nutritional effects on thyroid and catecholamine metabolism. Clin Sci 58:183, 1980.
  4. Huang HS, Pittman CS: Effects of thyroid hormone evaluated by cardiac systolic time interval in fasted subjects. J Formosan Med Ass 83:1087-93, 1994.
  5. Vignati L, Finley RJ, Haag S, Aoki TT: Protein conservation during prolonged fast: a function of triiodothyronine levels. Trans Assoc Am Physicians 16:169, 1978.
  6. Carter WJ, Shakir KM, Hodges S, Faas FH, Wynn JO: Effect of thyroid hormone on metabolic adaptation to fasting. Metabolism 24:1177, 1975.
  7. Wartofsky L, Burman D: Alterations in thyroid function in patients with systemic illness: The "euthyroid sick syndrome". Endocr Rev 3:164, 1982.
  8. Ingenbleek Y, Malvaux P: Peripheral turnover of thyroxine and related parameters in infant protein-calorie malnutrtion. Am J Clin Nutr 33:609, 1980.
  9. van der Westhuyzen JM: Plasma-T3 assay in Kwashiorkor. Lancet 2:965, 1973.
  10. Chopra IJ, Smith SR: Circulating thyroid hormones and thyrotropin in adult patients with protein-calorie malnutrition. J Clin Endocrinol Metab 40:221, 1975.

72a.        Orbak Z, Akin Y, Varoglu E, Tan H: Serum thyroid hormone and thyroid gland weight measurements in protein-energy malnutrition. J Pediatr Endocrinol Metab 11:719-24,1998.

72b.        Turkay S, Kus S, Gokalp A, Baskin E, Onal A: Effects of protein energy malnutrition on circulating thyroid hormones. Indian Pediatr   32:193-7,1995.

  1. Ingenbleek Y, Beckers C: Thyroidal iodide clearance and radioiodide uptake in protein-calorie malnutriton. Am J Clin Nutr 31:408, 1978.
  2. Pimstone B, Becker D, Hendricks S: TSH response to synthetic thyrotropin-releasing hormone in human protein-calorie malnutrition. J Clin Endocrinol Metab 36:779, 1973.
  3. Tulp OL, Krupp PP, Danforth E Jr., Horton ES: Characteristics of thyroid function in experimental protein malnutrion. J Nutr 109:1321, 1979.
  4. Falconer IR, Marchant B: Thyroxine utilization in lambs in natural and controlled environments. J Endocrinol 46:363, 1970.
  5. Danforth E Jr.,, Horton ES, O'Connell M, et al: Dietary-induced alterations in thyroid hormone metabolism during overnutrition. J Clin Invest 64:1336, 1979.
  6. Bray GA, Fisher DA, Chopra IJ: Relation of thyroid hormones to bodyweight. Lancet 1:1206, 1976.
  7. Glass AR, Burman KD, Dahms WT, Boehm TM: Endocrine function in human obesity. Metabolism 30:89, 1981.
  8. Robison LM, Sylvester PW, Birkenfeld P, Lang JP, Bull RJ Comparison of the effects of iodine and iodide on thyroid functioinn in humans. J Toxicol Environ Health 55:93-106,1998.

80a.        Uyttersprot N, Pelgrims N, Carrasco N, Gervy C, Maenhaut C, Dumont JF, Miot F: Moderate doses of iodid in vivo inhibit cell proliferation and the expression of thyroperoxidase and the Na+/I- symporter mRNAs in dog thyroid. Moll Cell Endocrinol 131:195-203,1997.

80b.        Pregliasco L, Bocanera L, Krawiec L, Siberschmidt D, Pisarev M, Juvenal G: Effects of iodid on thyroglobulin biosynthesis in FRTL-5 cells. Thyroid 6:319-23,1996

  1. Vagenakis AG, Downs P, Braverman LE, Burger A, Ingbar SH: Control of thyroid hormone secretion in normal subjects receiving iodides. J Clin Invest 52:528, 1973.
  2. Vagenakis AG, Rapoport B, Azizi F, et al: Hyper-response to thyrotropin-releasing hormone accompanying small decreases in serum thyroid hormone concentration. J Clin Invest 54:913-918, 1974.

82a.        Vitale M, DiMatola T, D”Ascoli F, Salzano S, Bogazzi F, Frnzi G, Martino E, Rossi G.   Iodide excess induces apoptosis through a p53 independent mechanism involving oxidative stress. Endocrinology 141:598-605, 2000

82b.        Burikhanov RB, Matsuzaki S. Excess iodine induces apoptosis in the thyroid of goitrogen-pretreated rats in vivo. Thyroid 10:123-9,2000

  1. Braverman LE, Ingbar SH, Vagenakis AG, Adams L, Maloof F: Enhanced susceptibility to iodide myxedema in patients with Hashimoto's disease. J Clin Endocrinol Metab 32:515, 1971.
  2. 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 281:816, 1969.
  3. Azizi F, Bentley D, Vagenakis A, et al: Abnormal thyroid function and response to iodides in patients with cystic fibrosis. Trans Assoc Am Physicians 87:111, 1974.
  4. Begg TB, Hall R: Iodide goitre and hypothyroidism. Q J Med 32:351, 1963.
  5. Pasternak DP, Socolow EL, Ingbar SH: Synergistic interaction of phenazone and iodide on thyroid hormone biosynthesis in the rat. Endocrinology 84:769, 1969.
  6. Shopsin B, Shenkman L, Blum M, Hollander CS: Iodine and lithium-induced hypothyroidism. Documentation of synergism. Am J Med 55:695, 1973.
  7. Milne K, Greer MA: Comparison of the effects of propylthiouracil and sulfadiazine on thyroidal biosynthesis and the manner by which they are influenced by supplemental iodide. Endocrinology 71:580, 1962.
  8. Vagenakis AG, Ingbar SH, Braverman LE: The relationship between thyroglobulin synthesis and intrathyroid iodine metabolism as indicated by the effects of cycloheximde in the rat. Endocrinology 94:1669, 1974.

90a.        Stanbury JB, Ermans AE, Bourdoux T, Todd C, Oken E, Tonglet R,Vidor G, Braverman LE, Medeiros-Neto G: Iodine induced hyperthyroidism: occurrence and epidemiology. Thyroid 8:83-100,1998.

  1. Jackson AS: Iodine hyperthyroidism: An analysis of fifty cases. Boston Med Surg J 193:1138, 1925.
  2. Vidor GI, Stewart JD, Wall JR, Wangel A, Hetzel BS: Pathogenesis of iodide induced thyrotoxicosis: Studies in northern Tasmania. J Clin Endocrinol Metab 37:901, 1973.
  3. Ermans AM, Camus M: Modifications of thyroid function induced by chronic administration of iodide in the presence of "autonomous" thyroid tissue. Acta Endocrinol 70:463, 1972.
  4. Vagenakis AG, Wang CA, Burger A, et al: Iodide-induced thyrotoxicosis in Boston. N Engl J Med 287:523-527, 1972.
  5. Suzuki H, Higuchi T, Sawa K, Ohtaki S, Horiuchi Y: "Endemic coast goitre" in Hokkaido, Japan. Acta Endocrinol 50:161, 1965.
  6. Wartofsky L: Low remission after therapy for Graves' disease: Possible relation of dietary iodine with antithyroid therapy results. JAMA 226:1083, 1973.
  7. Boyle JA, Greig WR, Fulton S, Dalakos TG: Excess dietary calcium and human thyroid function. J Endocrinol 34:532, 1966.

97a. Singh N, Singh PN, Hershman JM: Effect of calcium carbonate on the absorption of levothyroxine. JAMA 283:2822-5,2000

97b.        Singh N, Weisler, SL, Hershman JM The acute effect of calcium carbonate on the intestinal absoprption of levothyroxine. Thyroid 11:967-971; 2001

  1. Bloomfield RA, Welsch CW, Garner GB, Muhrer ME: Effect of dietary nitrate on thyroid function. Science 134:1690, 1961.
  2. Clode W, Sobral JM, Baptista AM: Bromine interference in iodine metabolism and its goitrogenic action. R. Pitt-Rivers (eds), Advances in Thyroid Research, Pergamon Press, New York, pp. 65,1961.

99a.        Vobecky M, Babicky A, Lerner J, Svandova E: Interaction of bromine with iodine in the rat thyroid gland at enhanced bromide intake. Biol Trace elem Res 54:207-12,1996.

99b.        Velicky J, Titlbach M, Duskova J, Vobecky M, Strbak V, Raska I: Potassium bromide and the thyroid gland of the rat:L morphology and immunohistochemistry, RIA and INAA analysis. Anat Anz 179:421-31,1997.

  1. Bach I, Braun S, Gati T, Kertai P, Sós J, Udvardy A: Effect of rubidium on the thyroid. R. Pitt-Rivers (eds), Advances in Thyroid Research, Pergamon Press, New York, pp. 505,1961.
  2. Galletti PM, Joyet G: Effect of fluorine on thyroidal iodine metaoblism in hyperthyroidism. J Clin Endocrinol Metab 18:1102, 1958.
  3. Gedalia I, Brand N: The relationship of fluoride and iodine in drinking water in the occurrence of goiter. Arch Int Pharmacodyn Ther 142:312, 1963.
  4. Siddiqui AH: Incidence of simple goitre in areas of endemic fluorosis. J Endocrinol 20:201, 1960.
  5. Day TK, Powell-Jackson PR: Fluoride water hardness, and endemic goiter. Lancet 1:1135, 1972.
  6. Paley KR, Sobel ES, Yalow RS: Effect of oral and intravenous cobaltous chloride on thyroid function. J Clin Endocrinol Metab 18:850, 1958.

105a.      Stangl GI, Schwartz FJ, Kirchgessner M. Cobalt deficiency effects on trace elements, hormones and enzymes involved in energy metabolsim in cattle. Int J Vitam Nutr Res 69:120-6,1999

105b.      Barceloux, DG. Cobalt. J Toxicol Clin Toxicol 37:201-6,1999

  1. Pimentel-Malaussera E, Roche M, Lavrisse M: Treatment of eight cases of hyperthyroidism with cobaltous chloride. JAMA 167:1719, 1958.

106a.      Gupta P, Kar A: Role of ascorbic acid in cadmium-induced thyroid dysfunction and lipid peroxidation. J Appl Toxicol 18:317-20,1998.

106b.      Paier B, Pavia MA Jr, Hansi C, Noli MI, Hagmuller K, Zaninovich AA: Cadmium inhibits the in vitro conversion of thyroxine to triiodothyronine in rat brown adipose tissue. Bull Environ Contam Toxicol 59:164-70,1997.

106c.      Gupta P, Chaurasia SS, Maiti PK, Kar A: Cadmium induces alterations in extrathyroidal conversion of thyroxine to triiodothyronine 5” monodeiodinase in male mouse. Horm Metab Res 29:151-2,1997

  1. Pousset GB J., Berthezene F, Tourniare J, Devic M: Myxoedeme au lithium. Ann Endocrinol 34:549, 1973.
  2. Berens SC, Bernstein RS, Robbins J, Wolff J: Antithyroid effects of lithium. J Clin Invest 49:1357, 1970.
  3. Spaulding SW, Burrow GN, Bermudez F, Himmelhoch JM: The inhibitory effect of lithium on thyroid hormone release in both euthyroid and thyrotoxic patients. J Clin Endocrinol Metab 35:905, 1972.
  4. Carlson HE, Temple R, Robbins J: Effect of lithium on thyroxine disappearance in man. J Clin Endocrinol Metab 36:1251, 1973.

110a.      Lazarus, JH. The effects of lithium therapy on thyroid and thyrotropin-releasing hormone. Thyroid 8:909-13,1998.

110b.      Tasevski V, Been D, King M, Luttrell B, Simpson A: Mitogenic effects in FRTL-5 cells can be reversed by blocking de novo cholesterol synthesis and subsequent signal transduction. Thyroid 2000:305-11, 2000.

  1. Burman KD, Diamond RC, Earll JM, Wright FD, Wartofsky L: Sensitivity to lithium in treated Graves' disease: Effects on serum T4, T3 and reverse T3. J Clin Endocrinol Metab 43:606, 1976.
  2. Blomqvist N, Lindstedt G, Lundberg PA, Walinder J: No inhibition by Li+ of thyroxine monodeiodination to 3,5,3'-triiodothyronine and 3,3',5'-triiodothyronine (reverse triiodothyronine). Clin Chim Acta 79:457, 1977.
  3. Linquette M, Lefebre J, Van Parys C, Wemeau JL: Le lithium dans le traitement des thyrotoxicoses. Ann d'Endocrinol 39:15, 1978.
  4. Andersen BF: Iodide perchlorate discharge test in lithium-treated patients. Acta Endocrinol 73:35, 1973.
  5. Berens SC, Wolff J, Murphy DL: Lithium concentration by the thyroid. Endocrinology 87:1085, 1970.
  6. Wolff J, Berens SC, Jones AB: Inhibition of thyrotropin-stimulated adenyl cyclase activity of beef thyroid members by low concentration of lithium ion. Biochem Biophys Res Commun 39:77, 1970.
  7. Bhattacharya B, Wolff J: Stabilization of microtubules by lithium ion. Biochem Biophys Res Commun 73:383, 1976.

117a.      Baumgartner A, Pinna G, Hiedra L, Gaio U, Hessenius C, Campos-Barros A, Eravci M, Prengel H, Thoma R, Meinhold H. Effects of lithium and carbamazapine on thyroid hormone metabolism in rat brain. Neuropsychopharmacology 16:25-41,1997.

117b.   Hahn CG, Pawlyk AC, Whybrow, PC Gyulai L, Tejani-Butt SM. Lithium administration affects gene expression of thyroid hormone receptors in rat brain. Life Sci 64:1793-802, 1999

  1. Lazarus JH, Joh R, Bennie EH, et al: Lithium therapy and thyroid function: A long term study. Psych Med 11:85-92, 1981.

118a.      Kirov G. Thyroid disorders in lithium-treated patients. J Affect Disord 50:33-40,1998

118b.      Johnston AM, Eagles JM. Lithium-associated hypothyroidism. Prevalence and risk factors. Br J Psychiatry 175:336-9,1999

118c.      Kusalic M, Engelsmann F. Effect of lithium maintenance therapy on thyropid and parathyroid function. J Psychatry Neurosci 24:227-33,1999

  1. Segal RL, Rosenblatt S, Eliasoph I: Endocrine exophthalmos during lithium therapy of manic-depressive disease. N Engl J Med 289:136, 1973.

119a.      Berry MJ, Banu L, Larsen PR: Type I iodothyronine deiodinase is a selenocystein-containing enzyme. Nature 349:438-,1991.

119b.      Olivieri O, Girelli D, Stanzial AM, Rossi L, Bassi A, Corrocher R: Selenium, zinc and thyroid hormones in healthy subjects: low T3/T4 ratio in the elderly is related to selenium status. Biol Trace Elem Res 51:31-41,1996.

119c          Duntas LH: Selenium and the thyroid: A close knit connection.   JCEM   95:5180-5188, 2010

119d.      Hotz CS, Fitzpatrick DW, Trick KD, L’Abbe MR: Dietary iodine and selenium interact to affect thyroid hormone metabolism. J Nutr 127:1214-8,1997

119e.      Mitchell JH, Nicol F, Beckett GJ, Arthur JR: Selenium and iodine deficiencies: effects on brain and brown adipose tisse seleneoenzyme activity and expression. J Endocrinol 155:255-63,1997

119f.       Zimerman MB, Adou P, Torresani T, Zeder C, Hurrell RF. Effect of iodized oil on thyroid size and thyroid hormone metabolism with concurrent selenium and iodine deficiency. Eur J Clin Nutr. 54:209-13, 2000

119f.       Wu O, Rayman MP, Lv H, Schomburg L, Cui B, Gao C er al: Low population selenium status is associated with increased prevalence of thyroid disease. J Clin Endocrinol Metab 100:4037-4047, 2015

  1. Kracht J: Fright-thyrotoxicosis in the wild rabbit, a model of thyrotrophic alarm-reaction. Acta Endocrinol 15:355, 1954.
  2. Falconer IR, Hetzel BS: Effect of emotional stress and TSH on thyroid vein hormone level in sheep with exteriorized thyroids. Endocrinology 75:42, 1964.
  3. Haibach H, McKenzie JM: Increased free thyroxine postoperatively in the rat. Endocrinology 81:435, 1967.
  4. Hagenfeldt I, Melander A, Thorell J, Tibblin S, Westgren U: Active and inactive thyroid hormone levels in elective and acute surgery. Acta Chir Scand 145:77, 1979.
  5. Chan V, Wang C, Yeung RTT: Pituitary-thyroid responses to surgical stress. Acta Endocrinol 88:490, 1978.
  6. Socolow EL, Woeber KA, Purdy RH, Holloway MT, Ingbar SH: Preparation of I131 labeled human serum prealbumin and its metabolism in normal and sick patients. J Clin Invest 44:1600, 1976.
  7. Brandt MR, Skovsted L, Kehlet H, Hansen JM: Rapid decrease in plasma-triiodothyronine during surgery and epidural analgesia independent of afferent neurogenic stimuli and of cortisol. Lancet 2:1333, 1976.
  8. Tingley JO, Morris AW, Hill SR, Pittman JA: The acute thyroid response to emotional stress. Ala J Med Sci 2:297, 1965.
  9. Cohen KL, Swigar ME: Thyroid function screening in psychiatric patients. JAMA 242:254, 1979.
  10. Levy RP, Jensen JB, Laus VG, Agle DP, Engel IM: Serum thyroid hormone abnormalities in psychiatric disease. Metabolism 30:1060, 1981.
  11. Spratt DI, Pont A, Miller MB, McDougall IR, Bayer MF, McLaughlin WT: Hyperthyroxinemia in patients with acute psychiatric disorders. Am J Med 73:41, 1982.
  12. Chesney AM, Clawson TA, Webster B: Endemic goitre in rabbits. I. Incidence and characteristics. Johns Hospkins Hosp Bull 43:261, 1928.
  13. Ermans AM, Delange F, Van Der Velden M, Kinthaert J: Possible role of cyanide and thiocyanate in the etiology of endemic cretinism. J. B. Stanbury and R. L. Kroc (eds), Human Development and the Thyroid Gland. Relation to Endemic Cretinism, Plenum Press, New York, pp. 455,1972.
  14. Monekasso GL, Wilson J: Plasma thiocyanate and vitamin B12 in Nigerian patients with degenerative neurological disease. Lancet 1:1062, 1971.
  15. Delange F, Ermans AM: Role of a dietary goitrogen in the etiology of endemic goiter on Idjwi Island. Am J Clin Nutr 24:1354, 1971.
  16. Delange F, Thilly C, Ermans AM: Iodine deficiency, a permissive condition in the development of endemic goiter. J Clin Endocrinol Metab 28:114, 1968.
  17. Langer P, Greer MA: Antithyroid activity of some naturally occurring isothiocyanates in vitro. Metabolism 17:596, 1968.
  18. Langer P: Antithyroid action in rats of small doses of some naturally occurring compounds. Endocrinology 79:1117, 1966.

137b   Chu M, Selzer TF: Myxedema coma induced by ingestion of raw bok choy. New Engl J Med 362:1945-1946, 2010

  1. Clements FW, Wishart JW: A thyroid-blocking agent in the etiology of endemic goiter. Metabolism 5:623, 1956.
  2. Peltola P: The goitrogenic effect of milk obtained from the region of endemic goitre in Finland. R. Pitt-RIvers (eds), Advances in Thyroid Research, Pergamon Press, New York, pp. 10,1961.
  3. Peltola P, Krusius FE: Effect of cow's milk from the goitre endemic district of Finland on thyroid function. Acta Endocrinol 33:603, 1960.
  4. Kilpatrick R, Broadhead GD, Edmonds CJ, Munro DS, Wilson GM: Studies on goitre in the Sheffield region. R. Pitt-Rivers (eds), Advances in Thyroid Research, Pergamon Press, New York, pp. 273,1961.

141a.      Laurberg P, Andersen S, Knudsen N, Ovesen L, Nohr SB, Pedersen IB: THiocyanate in food and iodine in milk: From domestic animal feeding to improved understanding of cretinism. Thyroid 12:897-902; 2002

141b.      Charatchaoenwitthaya N, Ongphiphadhanakul B, Pearce EC, Charintip S, Chanthasenanont A, He X, Chailurkit L and Braverman LE: The association between perchlorate and thiocyanate exposure and thyroid function in first trimester pregnant Thai women J Clin Endocrinol Metab 99:2365-2371, 2014

  1. Astwood EB, Greer MA, Ettlinger MG: L-5-Vinyl-2-thiooaxazolidone, an antithyroid compound from yellow turnip and from bassica seeds. J Biol Chem 181:121, 1949.
  2. Greer MA: The isolation and identification of progoitrin from bassica seed. Arch Biochem Biophys 99:369, 1962.
  3. Langer P, Michajlovskij N: Studies on the antithyroid activity of naturally occurring L-5-vinyl-2-thiooxazolidone and its urinary metabolite in rats. Acta Endocrinol 62:21, 1969.
  4. Krusius FE, Peltola P: The goitrogenic effect of naturally occurring L-5-vinyl- and L-5-phenyl-2-thio-oxazolidone in rats. Acta Endocrinol 53:342, 1966.
  5. Arstila A, Krusius FE, Peltola P: Studies on the transfer of thio-oxazolidone-type goitrogens into cow's milk in goiter endemic districts of Finland and in experimental conditions. Acta Endocrinol 60:712, 1969.
  6. Barzelatto J, Beckers C, Stevenson C, et al: Endemic goiter in Pedgregoso (Chile). I. Description and fuction studies. Acta Endocrinol 54:577, 1967.
  7. Linazasoro JM, Sanchez-Martin JA, Jiminez-Diaz C: Goitrogenic effect of walnuts. Lancet 2:501, 1966.
  8. Gaitan E, Wahner HW, Correa P, et al: Endemic goiter in the Cauca Valley: I. Results and limitations of twelve years of iodine prophylaxis. J Clin Endocrinol Metab 28:1730, 1968.

149a.  Abel Gadir WS, Adam SE: Development of goitre and enterohepatonephropathy in Nubian Goats fed with pearl millet (pennisetum typhoides) Vet J 157: 178-85,1999

  1. McCarrison R: The goitrogenic action of soybean and ground-nut. Indian J Med Res 21:179, 1933.
  2. Van Wyk JJ, Arnold MB, Wynn J, Pepper F: The effects of a soybean product on thyroid functions in humans. Pediatrics 24:752-760, 1959.
  3. Pinchera A, MacGillivray MH, Crawford JD, Freeman AG: Thyroid refractoriness in an athyrotic cretin fed soybean formula. N Engl J Med 273:83-86, 1965.
  4. Yamada T: Effect of fecal loss of thyroxine on pituitary-thyroid feedback control in the rat. Endocrinology 82:327, 1968.
  5. Bray GA: Increased sensitivity of the thyroid in iodine-depleted rats to the goitrogenic effects of thyrotropin. J Clin Invest 47:1640, 1968.
  6. Bull GM, Fraser R: Myxedema from resorcinol ointment applied to leg ulcers. Lancet 1:851, 1950.
  7. Selenkow HA, Rivera A, Thorn GW: The effects of amphenone on thyroid function in man. J Clin Endocrinol Metab 17:1131, 1957.
  8. Pittman JA, Brown RW: Antithyroid and antiadrenocortical activity of aminoglutethimide. J Clin Endocrinol Metab 26:1014, 1966.
  9. Rallison ML, Kumagai LF, Tyler FH: Goitrous hypothyroidism induced by aminoglutethmide, anticonvulsant drug. J Clin Endocrinol Metab 27:265, 1967.
  10. Jukes TH, Shaffer CB: Antithyroid effects of aminotriazole. Science 132:296, 1960.
  11. Pittman JA, Brown RW: Antithyroid action of amphenidone. J Clin Endocrinol Metab 22:100, 1962.
  12. Current JV, Hales IB, Dobyns BM: The effect of 2,3-dimercaptopropanol (BAL) on thyroid function. J Clin Endocrinol Metab 20:13, 1960.
  13. Sharpe AR Jr.: Inhibition of thyroidal 131I uptake by parabromdylamine maleate. J Clin Endocrinol Metab 21:739, 1961.
  14. Linsk JA, Paton BC, Persky M, Isaacs M, Kupperman HS: The effect of phenylbutazone and a related analogue (G25671) upon thyroid function. J Clin Endocrinol Metab 17:416, 1957.
  15. Wyngaarden JB, Stanbury JB, Rapp B: The effects of iodide, perchlorate, thiocyanate, and nitrate administration upon the iodide concentrating mechanism of the rat thyroid. Endocrinology 52:568, 1953.
  16. Ermans AM, Goossens F: Influence du perchlorate et du methimazol sur l'excretion urinaire de l'iode chez l'homme. Arch Int Pharmacodyn Ther 132:487, 1961.
  17. Stewart RDH, Murray IPC: An evaluation of the perchlorate discharge test. J Clin Endocrinol Metab 26:1050, 1966.
  18. Scranton JR, Nissen WM, Halmi NS: The kinetics of the inhibition of thyroidal iodide accumulation by thiocyanate: A reexamination. Endocrinology 85:603, 1969.
  19. Frohman LA, Klocke FJ: Recurrent thiocyanate intoxication, with pancytopenia, hypothyroidism, and psychosis. N Engl J Med 268:701, 1963.
  20. Taurog A, Potter GD, Chaikoff IL: Conversion of inorganic 131I to organic 131I by cell free preparations of thyroid tissue. J Biol Chem 213:119, 1955.
  21. Anbar M, Guttman S, Lewitus Z: Effect of monofluorosulphanate, difluorophosphate, and F borate ions on the iodine uptake of the thyroid gland. Nature 183:1517, 1959.
  22. Anbar M, Guttman S, Lewitus Z: The accumulation of fluoroborate ions in thyroid glands of rats. Endocrinology 66:888, 1960.
  23. Chow SY, Chang LR, Yen MS: A comparison between the uptakes of radioactive perchlorate and iodide by rat and guinea-pig thyroid glands. J Endocrinol 45:1, 1969.
  24. Crooks J, Wayne EJ: A comparison of potassium perchlorate, methylthiouracil, and carbimazole in the treatment of thyrotoxicosis. Lancet 1:401, 1960.
  25. Michajlovskij N, Langer P: Increase of serum free thyroxine following the administration of thiocyanate and other anions in vivo and in vitro. Acta Endocrinol 75:707-716, 1974.

174a.      Li FX, Squartsoff L, Lamm SH: Prevalence of thyroid disease in Nevada counties with respect to perchlorate in drinking water. J Occup Environ Med 43:630-4;2001

174b    Pearce EN, Lazarus JH, Smyth PPA, He X, Dall’amico D, et al: Perchlorate and thiocyanate exposure and thyroid function in first-trimester pregnant women. JCEM 945:3207-3215, 2010

174c.      Morgan JW, Cassady RE: Community cancer assessment in response to long-time exposure to perchlorate and trichloroethylene in drinking water. J Occup Environ Med 44:616-21;2002

174d.      Kelsh MA, Buffler PA, Daaboul JJ, Rutherford GW, Lau EC, Barnard JC, Exuzides AK, Madl AK, Palmer LG, Lorey FW: Primary congenital hypothyroidism, newborn thyroid function, and environmental perchlorate exposure among residents of a Southern California community. J Occup Environ Med 45:1116-27;2003

  1. Rosenberg IN: The antithyroid activity of some compounds that inhibit peroxidase. Science 116:503, 1952.
  2. DeGroot LJ, Davis AM: Studies on the biosynthesis of iodotyrosines: A soluble thyroidal iodide-peroxidase tyrosine-iodinase system. Endocrinology 70:492, 1962.
  3. Yamazaki E, Noguchi A, Slingerland DW: Effect of methylthiouracil and iodide on the iodinated constitutents of thyroid tissue in Graves' disease. J Clin Endocrinol Metab 20:889, 1960.
  4. Iino S, Yamada T, Greer MA: Effect of graded doses of propylthiouracil on biosynthesis of thyroid hormones. Endocrinology 68:582, 1961.
  5. Mulvey PF Jr., Slingerland DW: The in vitro stimulation of thyroidal activity by propylthiouracil. Endocrinology 70:7, 1962.
  6. Selenkow HA, Collaco FM: Clinical pharmacology of antithyroid compounds. Clin Pharmacol Ther 2:191, 1961.
  7. Astwood EB: Mechanisms of action of various antithyroid compounds. Ann NY Acad Sci 50:419, 1949.
  8. Maloff F, Spector L: The desulfuration of thiourea by thyroid cytoplasmic particulate fractions. J Biol Chem 234:949, 1959.
  9. Maloof F, Soodak M: Cleavage of disulfide bonds in thyroid tissue by thiourea. J Biol Chem 236:1689, 1961.
  10. Mitchell ML, Sanchez-Martin JA, Harden AB, O'Rourke ME: Failure of thiourea to prevent hormone synthesis by the thyroid gland of man and animals treated with TSH. J Clin Endocrinol Metab 21:157, 1961.
  11. Mayberry WE, Astwood EB: The effect ofpropylthiouracil on the intrathyroid metaoblism of iodine in rats. J Biol Chem 235:2977, 1960.
  12. Escobar del Rey F, Morreale de Escobar G: The effect of propylthiouracil, methylthiouracil and thiouracil on the peripheral metabolism of L-thyroxine in thyroidectomized L-thyroxine maintained rats. Endocrinology 69:456-465, 1961.
  13. Van Middlesworth L, Jones SL: Interference with deiodination of some thyroxine analogues in the rat. Endocrinology 69:1085, 1961.
  14. Escobar del Rey F, Morreale de Escobar G, Garcia-Garcia MD, Mouriz Garcia J: Increased secretion of thyrotrophic hormone in rats with a depressed peripheral deiodination of thyroid hormone and a normal or high plasma PBI. Endocrinology 71:859, 1962.
  15. Slingerland DW, Burrows BA: Inhibition by propylthiouracil of the peripheral metabolism of radiothyroxine. J Clin Endocrinol Metab 22:511, 1962.
  16. Furth ED, Rives K, Becker DV: Nonthyroidal action of propylthiouracil in euthyroid, hypothyroid, and hyperthyroid man. J Clin Endocrinol Metab 26:239-246, 1966.
  17. Oppenheimer JH, Schwartz HL, Surks MI: Propylthiouracil inhibits the conversion of L-thyroxine to L-triiodothyronine. An explanation of the antithyroxine effect of propylthiouracil and evidence supporting the concept that triiodothyronine is the active hormone. J Clin Invest 51:2493-2497, 1972.
  18. Stasilli NR, Kroc RL, Edlin R: Selective inhibition of the calorigenic activities of certain thyroxine analogues with chronic thiouracil treatment in rats. Endocrinology 66:872, 1960.
  19. Bray GA, Hildreth S: Effect of propylthiouracil and methimazole on the oxygen consumption of hypothyroid rats receiving thyroxine or triiodothyronine. Endocrinology 81:1018, 1967.
  20. Ruegamer WR, Warren JS, Barstow M, Beck W: Effects of thiouracil on rat liver alpha-glycerophosphate dehydrogenase and serum PBI responses to L-thyroxine. Endocrinology 81:277, 1967.
  21. Chopra IJ, Solomon DH, Chopra U, Wu SY, Fisher DA, Nakamura Y: Pathways of metabolism of thyroid hormones. Recent Prog Horm Res 34:521, 1978.
  22. Pittman JA, Beschi RJ, Smitherman TC: Methimazole: Its absorption and excretion in man and tissue distribution in rats. J Clin Endocrinol Metab 33:182, 1971.
  23. Marchant B, Alexander WD, Lazarus JH, Lees J, Clark DH: The accumulation of 35S antithyroid drugs by the thyroid gland. J Clin Endocrinol Metab 34:847, 1972.
  24. Krieger DT, Moses A, Ziffer H, Gabrilove JL, Soffer LJ: Effect of acetazoleamide on thyroid metabolism. Am J Physiol 196:291, 1959.
  25. Gabrilove JL, Alvarez AA, Soffer LJ: Effect of acetazoleamide (Diamox) on thyroid function. J Appl Physiol 13:491, 1958.
  26. Brown J, Solomon DH: Mechanism of antithyroid effects of a sulfonylurea in the rat. Endocrinology 63:473, 1958.
  27. Tranquade RE, Solomon DH, Brown J, Greene R: The effect of ora hypoglycemic agents on thyroid function in the rat. Endocrinology 67:293, 1960.
  28. Nikkilä EA, Jakobson T, Josipii SG, Karlsson K: Thyroid function in diabetic patients under long-term sulfonylurea treatment. Acta Endocrinol 33:623, 1960.
  29. Skinner NS Jr., Hayes RL, Hill SR Jr.: Studies on the use of chlorpropamide in patients with diabetes mellitus. Ann NY Acad Sci 74:830, 1959.
  30. Hunton RB, Wells MV, Skipper EW: Hypothyroidism in diabetics treated with sulphonylurea. Lancet 2:449, 1965.
  31. Hershman JM, Konerding K: Effects of sulfonylurae drugs on the thyroid and serum protein binding of thyroxine in the rat. Endocrinology 83:74, 1968.

205a.      Hagmer L: Polychlorinated biphenyls and thyroid status in humans: a review. Thyroid 13:1021-1028;2003

  1. Azizi F, Vagenakis AG, Portnay GI, et al: Thyroxine transport and metabolism in methadone and heroin addicts. Ann Intern Med 80:194-199, 1974.
  2. McKerron CG, Scott RL, Asper SP, Levy RI: Effects of clofibrate (Atromid S) on the thyroxine-binding capacity of thyroxine-binding globulin and free thyroxine. J Clin Endocrinol Metab 29:957-961, 1969.
  3. Beex L, Ross A, Smals P, Kloppenborg P: 5-Fluorouracil-induced increase of total thyroxine and triiodothyronine. Cancer Treat Rep 61:1291-1295, 1977.
  4. Oltman JE, Friedman S: Protein-bound iodine in patients receiving perphenazine. JAMA 185:726-727, 1963.
  5. Garnick MB, Larsen PR: Acute deficiency of thyroxine-binding globulin during L-asparaginase therapy. N Engl J Med 301:252-253, 1979.

210a.      Cashin-Hemphill L, Spencer CA, Nocoloff JT, et al: Alterations in serum thyroid hormonal indices with colestipol-niacin therapy. Ann Intern Med 107:324-329, 1987.

210b.      O'Brien T, Silverberg JD, Nguyen TT: Nicotinic-acid-induced toxicity associated with cytopenia and decreased levels of thyroxine-binding globulin. Mayo Clin Proc 67:465-468, 1992.

  1. Stockigt JR, Lim CF, Barlow JW, et al: Interaction of furosemide with serum thyroxine-binding sites: In vivo and in vitro studies and comparison with other inhibitors. J Clin Endocrinol Metab 60:1025-1031, 1985.
  2. Hollander CS, Scott RL, Burgess JA, et al: Free fatty acids: A possible regulator of free thyroid hormone levels in man. J Clin Endocrinol Metab 27:1219-1223, 1967.
  3. Tabachnick M, Hao YL, Korcek L: Effect of oleate, diphenylhydantoin, and heparin on the binding of 125I-thyroxine to purified thyroxine-binding globulin. J Clin Endocrinol Metab 36:392-394, 1973.
  4. Marshall JS, Tompkins LS: Effect of o,p'-DDD and similar compounds on thyroxine binding globulin. J Clin Endocrinol Metab 28:386-392, 1968.
  5. Abiodun MO, Bird R, Havard CW, Sood NK: The effects of phenylbutazone on thyroid function. Acta Endocrinol 72:257-264, 1973.
  6. Davis PJ, Hsu TH, Bianchine JR, Morgan JP: Effects of a new hypolipidemic agent, MK-185, on serum thyroxine-binding globulin (TBG) and dialysable fraction thyroxine. J Clin Endocrinol Metab 34:200-208, 1972.
  7. Taylor R, Clark F, Griffiths ID, Weeke J: Prospective study of effect of fenclofenac on thyroid function tests. Br J Med 281:911-912, 1980.
  8. Wiersinga WM, Fabius AJ, Touber JL: Orphenadrine, serum thyroxine and thyroid function. Acta Endocrinol 86:522-532, 1977.
  9. Pages RA, Robbins J, Edelhoch H: Binding of thyroxine and thyroxine analogs to human serum prealbumin. Biochemistry 12:2773-2779, 1973.
  10. Oppenheimer JH: Role of plasma proteins in the binding, distribution, and metabolism of the thyroid hormones. N Engl J Med 278:1153-1162, 1968.
  11. Man EB, Reid WA, Hellegers AE, Jones WS: Thyroid function in human pregnancy. III. Serum thyroxine-binding prealbumin (TBPA) and thyroxine-binding globulin (TBG) of pregnant women aged 14 through 43 years. Am J Obstet Gynecol 103:338, 1969.
  12. Glinoer D, Fernandez-Deville M, Ermans AM: Use of direct thyroxine-binding gloublin measurement in the evaluation of thyroid function. J Endocrinol Invest 1:329-335, 1978.
  13. Braverman LE, Ingbar SH: Effects of norethandrolone on the transport in serum and peripheral turnover of thyroxine. J Clin Endocrinol Metab 27:389-396, 1967.
  14. Graham RL, Gambrell RD: Changes in thyroid function tests during danazol therapy. Obstet Gyneocl 55:395-397, 1980.

224a.      Draper MW, Flowers, DE, Neild JA, Huster WJ, Zerbe RL: Antiestrogenic properties of raloxifene. Pharmacology 50:209-17,1995

224b.      Groonroos PE, Irjala KM, Selen GP, Forsstrom JJ: Computerized monitoring of potentially interfering medication in thyroid function diagnostics. Int j Clin Monit Comput 14:255-9,1997

224c.      Amberson J, Drinka PJ: Medication and low serum thyroxine values in nursing home residents. South Med J 91:437-40,1998

  1. Austen FK, Rubini ME, Meroney WH, Wolff J: Salicylates and thyroid function. I. Depression of thyroid function. J Clin Invest 37:1131-1143, 1958.
  2. Wolff J, Austen FK: Salicylates and thyroid function. II. The effect on the thyroid-pituitary interrelation. J Clin Invest 37:1144-1165, 1958.
  3. Christensen LK: Thyroxine-releasing effect of salicylate and of 2,4-dinitrophenol. Nature 183:1189-1190, 1959.
  4. Larsen PR: Salicylate-induced increases in free triiodothyronine in human serum: Evidence of inhibition of triiodothyronine binding to thyroxine-binding globulin and thyroxine-binding prealbumin. J Clin Invest 51:1125-1134, 1972.
  5. Oppenheimer JH, Tavernetti RR: Displacement of thyroxine from human thyroxine-binding globulin by analogues of hydantoin. Steric aspects of the thyroxine-binding site. J Clin Invest 41:2213-2220, 1962.
  6. Schatz DL, Sheppard RH, Steiner G, et al: Influence of heparin on serum free thyroxine. J Clin Endocrinol Metab 29:1015-1022, 1969.
  7. Hershman JM, Jones CM, Bailey AL: Reciprocal changes in serum thyrotropin and free thyroxine produced by heparin. J Clin Endocrinol Metab 34:574, 1972.

231a.      Jaume JC, Mendel CM, Frost PH, Greenspan FS, Laughton CW: Extremely low doses of heparin release lipase activity into the plasms and can thereby cause artifactual elevations in the serum-free thyroxine concentration as measured by equilibrium dialysis. Thyroid 6;79-83,1996

231b.      Stevenson HP, Archbold GP, Johnston P, Young IS, Sheridan B: Misleading serum free thyroxine results during low molecular weight heparin treatment. Clin Chem 44:1002-7,1998

  1. Dowling JT, Frienkel N, Ingbar SH: The effect of estrogens upon the peripheral metabolism of thyroxine. J Clin Invest 39:1119-1130, 1974.
  2. Refetoff S, Fang VS, Marshall JS, Robin NI: Metabolism of thyroxine-binding globulin (TBG) in man: Abnormal rate of synthesis in inherited TBG deficiency and excess. J Clin Invest 57:485-495, 1976.
  3. Schlienger JL, Kapfer MT, Singer L, Stephan F: The action of clomipramine on thyroid function. Horm Metab Res 12:481-482, 1980.
  4. Rootwelt K, Ganes T, Johannessen SI: Effect of carbamazapine, phenytoin and phenobarbitone on serum levels of thyroid hormones and thyrotropin in humans. Scand J Clin Lab Invest 38:731-736, 1978.
  5. Northcutt RC, Stiel MN, Nollifield JW, Stant EG Jr.: The influence of cholestyramine on thyroxine absorption. JAMA 208:1857-1861, 1969.
  6. Witztum JL, Jacobs LS, Schonfeld G: Thyroid hormone and thyrotropin levels in patients placed on colestipol hydrochloride. J Clin Endocrinol Metab 46:838-840, 1978.
  7. Isley WL: Effect of rifampin therapy on thyroid function tests in a hypothyroid patient on replacement L-thyroxine. Ann Int Med 107:517-518, 1987.

238a.      Campbell NR, Hasinoff BB, Stalts H, Rao B, Wong, NC: Ferrous sulfate reduces thyroxine efficacy in patietns with hypothyroidism. Ann Intern Med 117:1010-3,1992

238b.      Liel Y, Sperber AD, Shany S: Nonspecific intestinal adsorption of levothyroxine by aluminum hydroxide. AM J Med 97:363-5,1994.

238c.      Sherman SI, Tielens ET, Ladenson PW: Sucralfate causes malabsorption of L-thyroxine. Am J Med 96:531-5,1994.

  1. Chopra IJ, Williams DE, Orgiazzi J, Solomon DH: Opposite effects of dexamethasone on serum concentrations of 3,3',5'-triiodothyronine (reverse T3) and 3,3',5-triiodothyronine (T3). J Clin Endocrinol Metab 41:911-920, 1975.
  2. Duick DS, Warren DW, Nicoloff JT, et al: Effect of a single dose of dexamethasone on the concentration of serum triiodothyronine in man. J Clin Endocrinol Metab 39:1151-1154, 1974.
  3. Burger A, Dinichert D, Nicod P, et al: Effects of amiodarone on serum triiodothyronine, reverse triiothyronine, thyroxine and thyrotropin. J Clin Invest 58:255-259, 1976.
  4. Savoie JC, Massin JP, Thomopoulos P, Leger F: Iodine-induced thyrotoxicosis in apparently normal thyroid glands. J Clin Endocrinol Metab 41:685-691, 1975.
  5. Lotti G, Delitala G, Devilla L, Alagna S, Masala A: Reduction of plasma triiodothyronine induced by propranolol. Clin Endocrinol 6:405, 1977.
  6. Faber J, Kirkegaard C, Lumholtz IB, et al: Measurements of serum 3',5'-diiodothyronine and 3,3'-diiodothyronine concentrations in normal subjects and in patients with thyroid and nonthyroid disease: Studies of 3',5'-diiodothyronine metabolism. J Clin Endocrinol Metab 48:611-617, 1979.
  7. Wiersinga WM, Touber JL: The influence of ß-adrenoreceptor blocking agents on plasma thyroxine and triiodothyronine. J Clin Endocrinol Metab 45:293-298, 1977.
  8. Re RN, Kourides IA, Ridgway EC, et al: The effect of glucocorticoid administration on human pituitary secretion of thyrotropin and prolactin. J Clin Endocrinol Metab 43:338-346, 1976.
  9. Larsen PR, Atkinson AJ, Wellman HN, Goldsmith RE: The effect of diphenylhydantoin on thyroxine metabolism in man. J Clin Invest 49:1266-,1279, 1970.
  10. Schwartz HL, Kozyreff V, Surks MI, Oppenheimer JH: Increased deiodination of L-thyroxine and L-triiodothyronine by liver microsomes from rats treated with phenobarbital. Nature 221:1262-1263, 1969.
  11. Schwartz HL, Bernstein G, Oppenheimer JH: Effect of phenobarbital administration on the subcellular distribution of 125I-thyroxine in rat liver: Importance of microsomal binding. Endocrinology 84:270, 1969.
  12. Blum C, Corvette C, Beckers C: Effect of insulin induced hypoglycemia on thyroid function and thyroxine turnover. Eur J Clin Invest 3:124, 1973.
  13. Johnstone RE, Kennel EM, Brummond W Jr., Shaw LM, Ebersole RC: Effect of halothane anesthesia on muscle, liver, thyroid, and adrenal-function tests in man. Clin Chem 22:217, 1976.
  14. Scanlon MF, Weightman DR, Shale DJ, et al: Dopamine is a physiological regulator of thyrotropin (TSH) secretion in normal man. Clin Endocrinol 10:7-15, 1979.
  15. Scanlon MF, Rodriguez-Arnao MD, Pourmand M, et al: Catecholaminergic interactions in the regulation of thyrotropin (TSH) secretion in man. J Endocrinol Invest 3:125-129, 1980.
  16. Delitala G, Devilla L, Lotti G: Domperidone, an extracerebral inhibitor of dopamine receptors, stimulates thyrotropin and prolactin release in man. J Clin Endocrinol Metab 50:1127-1130, 1980.
  17. Massara F, Camanni F, Belforte L, et al: Increased thyrotropin secretion induced by sulpiride in man. Clin Endocrinol 9:419-428, 1978.
  18. Delitala G, Devilla L, Lotti G: TSH and prolactin stimulation by the decarboxylase inhibitor benserazide in primary hypothyroidism. Clin Endocrinol 12:313-316, 1980.
  19. Kirkegaard C, Bjoerum CN, Cohn D, et al: Studies of the influence of biogenic amines and psychoactive drugs on the prognostic value of the TRH stimulation test in endogeneous depression. Psychoneuroendocrinology 2:131-136, 1977.
  20. Kirkegaard C, Bjoerum N, Cohn D, Lauridsen UB: TRH stimulation test in manic-depressive illness. Arch Gen Psychiatry 35:1017-1021, 1978.
  21. Nelis GF, Van DeMeene JG: The effect of oral cimetidine on the basal and stimulated values of prolactin, thyroid stimulating hormone, follicle stimulating hormone and luteinizing hormone. Postgrad Med J 56:26-29, 1980.
  22. Feldt-Rasmussen U, Lange AP, Date J, Kern-Hansen M: Effect of clomifen on thyroid function in normal men. Acta Endocrinol 90:43-51, 1979.
  23. Smals AG, Kloppenborg PW, Hoefnagesl WH, Drayer JM: Pituitary-thyroid function in spirolactone treated hypertensive women. Acta Endocrinol 90:577-584, 1979.
  24. Morley JE, Shafer RB, Elson MK, et al: Amphetamine-induced hyperthyroxinemia. Ann Int Med 93:707-709, 1980.
  25. Gloebel B, Weinheimer B: TRH-test during D-T4 application. Nuc-Compact 8:44, 1977.
  26. Medeiros-Neto G, Kallas WG, Knobel M, et al: Triac (3,5,3'-triiodothyroacetic acid) partially inhibits the thyrotropin response to thyrotropin-releasing hormone in normal and thyroidectomized hypothyroid patients. J Clin Endocrinol Metab 50:223-225, 1980.
  27. Emrich D: Untersuchungen zum einfluss von Etiroxat-HCL auf den Jodstoffwechsel beim menschen. Arzneim Forsch 27:422-426, 1977.
  28. Tamagna EI, Hershman JM, Jorgensen EC: Thyrotropin suppression by 3,5-dimethyl-3'-isopropyl-L-thyronine in man. J Clin Endocrinol Metab 48:196-200, 1979.
  29. Yoshimura M, Hachiya T, Ochi Y, et al: Suppression of elevated serum TSH levels in hypothyroidism by fusaric acid. J Clin Endocrinol Metab 45:95-98, 1977.
  30. Delitala G, Rovasio P, Lotti G: Suppression of thyrotropin (TSH) and prolactin (PRL) release by pyridoxine in chronic primary hypothyroidism. J Clin Endocrinol Metab 45:1019-1022, 1977.
  31. Masala A, Delitala G, Devilla L, et al: Effect of apomorphine and peribedil on the secretion of thyrotropin and prolactin in patients with primary hypothyroidism. Metabolism 27:1608-1612, 1978.
  32. Delitala G, Wass JAH, Stubbs WA, et al: The effect of lisurgide hydrogen maleate, an ergot derivative on anterior pituitary hormone secretion in man. Clin Endocrinol 11:1-9, 1979.
  33. Nilsson KO, Thorell JI, Hökfelt B: The effect of thyrotrophin releasing hormone on the release of thyrotrophin and other pituitary hormones in man under basal conditions and following adrenergic blocking agents. Acta Endocrinol 76:24-34, 1974.
  34. Lamberg BA, Linnoila M, Fogelholm R, et al: The effect of psychotropic drugs on the SH-response to thyroliberin (TRH). Neuroendocrinology 24:90-97, 1977.
  35. Delitala G, Rovasio PP, Masala A, et al: Metergoline inhibition of thyrotropin and prolactin secretion in primary hypothyroidism. Clin Endocrinol 8:69-73, 1978.
  36. Ferrari C, Paracchi A, Rondena M, et al: Effect of two serotonin antagonists on prolactin and thyrotropin secretion in man. Clin Endocrinol 5:575-578, 1976.
  37. Collu R: The effect of TRH on the release of TSH, PRL and GH in man under basal conditions and following methysergide. J Endocrinol Invest 2:121-124, 1978.
  38. Yoshimura M, Ochi Y, Miyazaki T, et al : Effect of intravenous and oral administration of L-DOPA on HGH and TSH release. Endocrinol Jpn 19:543-548, 1972.
  39. Porter BA, Refetoff S, Rosenfield RL, et al: Abnormal thyroxine metabolism in hyposomatotrophic dwarfism and inhibition of responsiveness to TRH during GH therapy. Pediatrics 51:668-674, 1973.
  40. Siler TM, Yen SS, Guillemin R: Inhibition by somatostatin on the release of TSH induced in man by thyrotropin-releasing factor. J Clin Endocrinol Metab 38:742, 1974.
  41. Weeke J, Hansen AP, Lundbaek K: Inhibition by somatostatin of basal levels of serum thyrotropin (TSH) in normal men. J Clin Endocrinol Metab 41:168-171, 1975.

279a.      Colao A, Merola B, Ferone D, Marzullo P, Cerbone G, Longbardi S, Di Somma C, Lombardi G: Acute and chronic effects of octreotide on thyroid axis in growth hormone-secreting and clinically non-functional pituitary adenomas. Eur J Endocrinol 133:189-94,1995

  1. Thomas JA, Shahid-Salles KS, Donovan MP: Effects of narcotics on the reproduction system. Avd Sex Horm Res 3:169-195, 1977.
  2. May P, Mittler J, Manougian A, Erte N: TSH release-inhibiting activity of leucine-enkephaline. Horm Metab Res 11:30-33, 1979.
  3. Chan V, Wang C, Yeung RT: Effects of heroin addiction on thyrotropin, thyroid hormones and prolactin secretion in men. Clin Endocrinol 10:557-565, 1979.
  4. Kobayashi I, Shimomura Y, Maruta S, et al: Clofibrate and a related compound suppress TSH secretion in primary hypothyroidism. Acta Endocrinol 94:53-57, 1980.
  5. Delitala G: Dopamine and TSH secretion in man. Lancet 2:760-761, 1977.
  6. Refetoff S, Fang VS, Rapoport B, Friesen HG: Interrelationships in the regulation of TSH and prolactin secretion in man: Effects of L-DOPA, TRH and thyroid hormone in various combinations. J Clin Endocrinol Metab 38:450-457, 1974.
  7. Miyai K, Onishi T, Hosokawa M, et al: Inhibition of thyrotropin and prolactin secretions in primary hypothyroidism by 2-Br--ergocryptine. J Clin Endocrinol Metab 39:391-394, 1974.
  8. Burrow GN, May PB, Spaulding SW, Donabedian RK: TRH and dopamine interactions affecting pituitary hormone secretion. J Clin Endocrinol Metab 45:65, 1977.
  9. Spaulding SW, Burrow GN, Donabedian RK, Van Woert M: L-dopa supression of thyrotropin releasing hormone response in man. J Clin Endocrinol Metab 35:182, 1977.
  10. Collu R, Jéquier JC, Leboeuf G, et al: Endocrine effects of pimozide, a specific dopaminergic blocker. J Clin Endocrinol Metab 41:981-984, 1975.
  11. Kleinman RE, Vagenakis AG, Braverman LE: The effect of iopanoic acid on the regulation of thyrotropin secretion in euthyroid subjects. J Clin Endocrinol Metab 51:399-403, 1980.
  12. Faglia G, Ambrosi B, Beck-Peccoz P, et al: The effect of theophylline on plasma thyrotropin response (HTSH) to thyrotropin releasing factor (TRF) in man. J Clin Endocrinol Metab 34:906-909, 1972.
  13. Wolff J, Varrone S: The methyll xanthines - A new class of goitrogens. Endocrinology 85:410-414, 1969.
  14. Oyama T, Potsaid MS, Slingerland DW: Effect of diethyl ether anesthesia on thyroid function of rats: Pituitary, adrenal and thyroid relationship. Endocrinology 65:459, 1959.
  15. Fore W, Kohler P, Wynn J: Rapid redistribution of serum thyroxine during ether anesthesia. J Clin Endocrinol Metab 26:821, 1966.
  16. Cass R, Kuntzman R, Brodie BB: Norepinephrine depletion as possible mechanism of action of guanethidine (SU 5864), a new hypotensive drug. Proc Soc Exp Biol Med 103:871, 1960.
  17. Gaffney TE, Braunwald E, Kahler RL: Effects of guanethidine on triiodothyronine induced hyperthyroidism in man. N Engl J Med 265:16-20, 1961.
  18. Lee WY, Bronsky D, Waldstein SS: Studies of thyroid and sympathetic nervous system interrelationships. II. Effect of guanethidine on manifestations of hyperthyroidism. J Clin Endocrinol Metab 22:879-885, 1962.
  19. Ramey ER, Bernstein H, Goldstein MS: Effect of sympathetic blocking agents on the increased oxygen consumption following administration of thyroxine. Fed Proc 14:118, 1955.
  20. Surtskin A, Cordonnier JK, Lang S: Lack of influence of the sympathetic nervous system on the calorigenic response to thyroxine. Am J Physiol 188:503, 1957.
  21. Schwartz NB, Hammond GE, Gronert GA: Interaction between thyroxine and dibenzyline on metabolic rate. Am J Physiol 191:573, 1957.

300a      Ladenson PW, McCarren M, Morkin E, Edson RG, Shihs MC et al: Effects of the thyromimetic agent diiodothyropropanoic acid on body weight, body mass index and serum lipoproteins: A pilot prospective, randomized, controlled trial. JCEM 95:1349-1354, 2010

300b    Ladenson PW, Kristensen JD, Ridgway EC, Olsson AG, Carlsson B et al: Use of the thyroid hormone analogue eprotirome in statin-treated dyslipidemia. NEJM 362:906-916, 2010

300c       Sjouke B, Langslet G, Ceska R, Nicholis S, Nissen S, Ohlander M, Ladenson P, Olsson A, Hovingh G, Kastelen J     Eprotirome in patients with familial hypercholesterolemia (the AKKA trial); a randomized, double-blind, placebo-controlled phase 3 study. The Lancet Diabetes & Endocrinology 2:455-463, 2014

300d       Yehuda-Schnaidman E, Kalderon B and Bar-Tana J: Thyroid hormone, thryomimetics and metabolic efficiency. Endocrine Reviews 35:35-58, 2014

  1. Cutting WC, Rytand DA, Tainter ML: Relationship between blood cholesterol and increased metabolism from dinitrophenol and thyroid. J Clin Invest 13:547-552, 1934.

301a.      Anker GB, Lonning PE, Aakvaag A, Lien EA: Thyroid function in postmenopausal breast cancer patients treated with tamoxifen. Scand J Lab Clin Invest 58:103-7,1998.

301b.      Hsu SH, Cheng WC, Jang MW, Tsai KS: Effects of long term use of raloxifene, a selective estrogen receptor modulator, on thyroid function test profiles. Clin Chem 47:1865-1867;2001

301c.      Ceresini G, Morganti S, Rebecchi I, Bertone L, Ceda GP, Bacchi-Modena A, Sgarabotto M, Baldini M, Ablondi F, Valenti G, Braverman LE: A one-year follow-up on the effects of raloxifene on thyroid function in post-menopausal women. Menopause 11:176-9;2004

301d       Siraj ES, Gupta MK, Reddy SS Raloxifene causing malabsorption of levothyroxine. Arch Intern Med 9:1367-70;2003

  1. Ain KB, Mori Y, Refetoff S: Reduced clearance of thyroxine-binding globulin (TBG) with increased sialylation: A mechanism for estrogen induced elevation of serum TBG concentration. J Clin Endocrinol Metab 65:689-696, 1987.
  2. Doe RP, Mellinger GT, Swaim WR, Seal JS: Estrogen dosage effects on serum proteins: A longitudinal study. J Clin Endocrinol Metab 27:1081-1086, 1967.
  3. Ramey JN, Burrow GN, Polackwich RK, Donabedian RK: The effect of oral contraceptive steroids on the response of thyroid-stimulating hormone to thyrotropin-releasing hormone. J Clin Endocrinol Metab 40:712, 1975.
  4. Gross HA, Appleman MD, Nicoloff JT: Effect of biologically active steroids on thyroid function in man. J Clin Endocrinol Metab 33:242-248, 1971.
  5. Lemarchand-Beraud T, Rappoport G, Magrini G, Berthier C, Reymond M: Influences of different physiological conditions on the gonadotropins and thyrotropin responses to LHRH and TRH. Horm Metab Res 5 (suppl):170, 1974.

306a.      Moreira RM, Lisboa PC, Curty FH, Pazos-Moura CC: Dose-dependent effects of 17-beta-estriol on pituitary thyrotropin content and secretion in vitro. Braz. J. Med. Biol. Res. 30:1129-34,1997.

306b.      Zidan J, Rubenstein W: Effect of adjuvant tamoxifen therapy on thyroid function in postmenopausal women with breast cancer. Oncology 56:43-45, 1999.

  1. Haigler ED Jr., Hershman JM, Pittman JA Jr., Blaugh CM: Direct evaluation of pituitary thyrotropin reserve utilizing thyrotropin releasing hormone. J Clin Endocrinol Metab 33:573-581, 1971.
  2. Snyder PJ, Utiger RD: Response to thyrotropin releasing hormone (TRH) in normal man. J Clin Endocrinol Metab 34:380-385, 1972.
  3. Carlson HE, Jacobs LS, Daughaday WH: Growth hormone, thyrothyropin and prolactin responses to thyrotropin-releasing hormone following diethylstilbestrol pretreatment. J Clin Endocrinol Metab 37:488, 1973.
  4. Rutlin E, Haug E, Torjesen PA: Serum thyrotrophin, prolactin and growth hormone, response to TRH during oestrogen treatment. Acta Endocrinol 84:23-35, 1977.

310a   Poppe K, Glinoer D, Tournaye H, Schiettecatte J, Devorey P et al: Impact of ovarian hyperstimulation on thyroid function in women with and without thyroid autoimmunity. JCEM 89:3808-3812, 2004

310 b    Poppe K, Glinoer D, Tounaye H, Devroey P, Velkeniers B Impact of the ovarian hyperstimulation syndrome on thyroid function   Thyroid 18:801-802, 2008

  1. Federman DD, Robbins J, Rall JE: Effects of methyl testosterone on thyroid function, thyroxine metabolism, and thyroxine-binding protein. J Clin Invest 37:1024, 1958.
  2. Woeber KA, Barakat RM, Ingbar SH: Effects of salicylate and its noncalorigenic congeners on the thyroidal release of 131I in patients with thyrotoxicosis. J Clin Endocrinol Metab 224:1163-1168, 1964.
  3. Dussault JH, Turcotte R, Guyda H: The effect of acetylsalicylic acid on TSH and PRL secretion after TRH stimulation in the human. J Clin Endocrinol Metab 43:232-235, 1976.
  4. Langer P, Földes O, Michajlovskij N, et al: Short-term effect of acethylsalicylic acid on pituitary-thyroid axis and plasma cortisol level in healthy human volunteers. Acta Endocrinol 88:698, 1978.
  5. Chopra IJ, Solomon DH, Chua Teco GN, Nguyen AH: Inhibition of hepatic outer ring monodeiodination of thyroxine and 3,3',5'-triiodothyronine by sodium salicylate. Endocrinology 106:1728-1734, 1980.
  6. Alexander WD, Johnson KWM: A comparison of the effects of acetylsalicylic acid and DL-triiodothyronine in patients with myxoedema. Clin Sci 15:593-600, 1956.
  7. Yamamoto T, Woeber KA, Ingbar SH: The influence of salicylate on serum TSH concentration in patients with primary hypothyroidism. J Clin Endocrinol Metab 34:423-426, 1972.
  8. Woeber KA, Ingbar SH: The effects of noncalorigenic congeners of salicylate on the peripheral metabolism of thyroxine. J Clin Invest 43:931-942, 1964.
  9. Christensen K: The metabolic effect of p-aminosalicylic acid. Acta Endocrinol 31:608-610, 1959.
  10. MacGregor AG, Somner AR: The antithyroid action of para-amino salicylic acid. Lancet 2:931-936, 1954.
  11. Christensen LK: The metabolic effect of salicylate and other hydroxybenzoates. Acta Pharmacol Toxicol 16:129, 1959.
  12. McConnell RJ: Abnormal thyroid function in patients taking salsalate. JAMA 267:1242-1243, 1992.
  13. Gamstedt A, Jarnerot A, Kagedal B, Soderholm B: Corticosteroids and thyroid function. Acta Med Scand 205:379, 1979.
  14. Oppenheimer JH, Werner SC: Effect of prednisone on thyroxine-binding proteins. J Clin Endocrinol Metab 26:715-721, 1966.
  15. Werner SC, Platman SR: Remission of hyperthyroidism (Graves' disease) and altered pattern of serum-thyroxine binding induced by prednisone. Lancet 2:751, 1965.
  16. Otsuki M, Dakoda M, Baba S: Influence of glucocorticoids on TRF-induced TSH response in man. J Clin Endocrinol Metab 36:95, 1973.
  17. Dussault JH: The effect of dexamethasone on TSH and prolactin secretion after TRH stimulation. Can Med Assoc J 111:1195-1197, 1974.
  18. Berson SA, Yalow RS: The effect of cortisone on the iodine accumulating functions of the thyroid gland in euthyroid subjects. J Clin Endocrinol Metab 12:407, 1952.
  19. Ingbar SH: The effect of cortisone on the thyroidal and renal metabolism of iodine. Endocrinology 53:171-181, 1953.

329a.      Kaplan MM, Utiger RD: Iodothyronine metabolism in rat liver homogenates. J Clin Invest 61:459,1978.

329b.      Samuels MH, McDaniel PA: Thyrotrophin levels during hydrocortisone infusions that mimic fasting-induced cortisol elevations: a clinical research study. J Clin Endocrinol Metab 82:3700-4,1997.

329c.      Heikinheimo O, Ranta S, Grunberg S, Lahteenmaki P, Spitz IM: Alterations in pituitary-thyroid and pituitary-adrenal axes—consequences of long-term mifepristone treatment. Metabolism 46:292-6,1997.

  1. Topliss DJ, White EL, Stockigt JR: Significance of thyrotropin excess in untreated primary adrenal insufficiency. J Clin Endocrinol Metab 50:52-56, 1980.
  2. Tashjian AH, Osborne R, Maina D, Knaian A: Hydrocortisone increases the number of receptors for thyrotropin-releasing hormone on pituitary cells in culture. Biochem Biophys Res Commun 79:333, 1977.

331a.      Bruhn TO, Huang SS, Vaslet C, Nillni EA: Glucocorticoids modulate the biosynthesis and processing of prothyrotropin releasing hormone (proTRH). Endocrine 9:143-52.1998.

331b.      Jackson IM, Luo LG: Antidepressants inhibit the glucocorticoid stimulation of thyrotropin releasing hormone expression in cultured hypothalamic neurons. J Investig Med 46:470-4,1998.

  1. Benoit FL, Greenspan FS: Corticoid therapy for pretibial myxedema. Observations on the long-acting thyroid stimulator. Ann Intern Med 66:711-720, 1967.
  2. Yamada T, Ikejiri K, Kotani M, Kusakabe T: An increase of plasma triiodothyronine and thyroxine after administration of dexamethasone to hypothyroid patients with Hashimoto's thyroiditis. J Clin Endocrinol Metab 46:784-790, 1978.

333a.      Conn J, Sebiastain M, Deam D   A prospective study of the effect of non-ionic contast media on thryoid function. Thyroid 6:107-110;1996

333b.      Martin FI, Dream DR Hyperthyroidism in elderly hospitalized patients. Clinical features and treatment outcomes. Med j Aust 164:200-203,1996

333c.      Rhee CM, Bhan I, Alexnader EK, Brunelli SM Association between iodinated contrast media exposure and incident hyperthyroidism and hypothyroidism Arch Intern Med 172:153-159:2012

333d.      Nolte W, Muller, R, Siggelkow H, Emrich D, Hufner M   Prophylactic application of thyrostatic drugs during excessive iodine exposure in euthyroid patients with thyroid autonomy: a randomized study. Eur J of Endocrinol 134:337-41: 1996

333e.      Barr ML,Chiu HK, Li N, Yeh MW, Rhee CM, Casillas J, Iskander PJ, Leung AM Thyroid dysfunction in children exposed to iodinated contrast material J Clin Endocrinol Metab 2016 101:2366-2370

333f.       Kornelius E, Chiou JY, Yang YS, Peng CH, Lai YR and Huang CN:   Iodinated contrast media increased the risk of thyroid dysfunction: A 6-year retrospective cohort study.   J Clin Endocrinol Metab 100:3372-3379, 2015

  1. Burgi H, Wimpfheimer C, Burger A, Zaunbauer W, Rosler H, Lamarchand-Beraud T: Changes of circulating thyroxine, triiodothyronine and reverse triiodothyronine after radiographic contrast agents. J Clin Endocrinol Metab 43:1203, 1976.
  2. Wu SY, Chopra IJ, Solomon DH, Bennett LR: Changes in circulating iodothyronines in euthyroid and hyperthyroid subjects given Ipodate (Oragrafin), an agent for oral cholescystography. J Clin Endocrinol Metab 46:691-697, 1978.
  3. Larsen PR, Dick TE, Markovitz BP, Kaplan MM, Gard TG: Inhibition of intrapituitary thyroxine to 3,5,3'-triiodothyronine conversion prevents the acute suppression of thyrotropin release by thyroxine in hypothyroid rats. J Clin Invest 64:117, 1979.
  4. Felicetta JV, Green WL, Nelp WB: Inhibition of hepatic binding of thyroxine by cholecystographic agents. J Clin Invest 65:1032-1040, 1980.
  5. DeGroot LJ, Rue PA: Roentgenographic contrast agents inhibit triiodothyronine binding to nuclear receptors in vitro. J Clin Endocrinol Metab 49:538-542, 1979.

338a.      Brown RS, Cohen JH, Braverman LE: Successful treatment of massive acute thyroid hormone poisoning with iopanoic acid. J Pediatr 132:903-05,1998.

338b.      Arem R, Munipalli B: Ipodate therapy in patients with severe destruction induces thyrotoxicosis. Arch Intern Med 156:1752-7,1996.

338c.      Chopra IJ, van Herle AJ, Korenman SG, Viosca S, Younai S: Use of sodium ipodate in management of hyperthyroidism in subacute thyroiditis. J Clin Endocrinol Metab 80:2178-80,1995.

338d.      Fontanilla JC, Schnedier AB, Sarne DH: The use of oral radiographic Contrast agents in the management of hyperthyroidism. Thyroid 11:561-567: 2001

  1. Nademanee K, Piwonka RW, Singh BN, Hershman JM: Amiodarone and thyroid function. Prog Cardiovascul Dis 31:427-437, 1989.

339a.      deJong M, Docter R, Van der Hoek H, et al: Different effects of amiodarone on transport of T4 and T3 into the perfused rat liver. Am J Physiol 266:E44, 1994.

339b.      Vorperian VR, Havighurst TC, Miller S, January CT: Adverse effects of low dose amiodarone: a meta-analysis. J Am Coll Cardiol 30:791-8,1997.

339c.      Hudig F, Bakker O, Wiersinga WM: Tri-iodothyronine prevent the amiodarone-induced decrease in the expression of the liver low-density lipoprotein receptor gene. J Endocrinol 152:413-21,1997

339d.      Guo W, Kamiya K, Toyama J: Evidences of antagonism between amiodarone and triiodothyronine on the K+ channel activities of cultured rat cardiomyocytes. J Mol Cell Cardiol 29:617-27,1997.

  1. Martino E, Safran M, Aghini-Lombardi F, et al: Environmental iodine intake and thyroid dysfunction during chronic amiodarone therapy. Ann Intern Med 101:28-34, 1984.

340a.      Bartlena L, Brogioni S, Grasso L, Bogazzi F, Burelli A, Martino E: Treatment of amiodarone-induced thyrotoxicosis, a difficult challenge: results of a prospective study. J Clin Endocrinol Metab 81:2930-3,1996.

340b   Bogazzi F, Bartalena L, Martino E: Approach to the patient with amiodarone-induced thyrotoxicosis. JCEM 96:2529-2535, 2010

340c.         Tomisits L, Rossi G, Baratalena L, Martino E Bigazzi F The onset of amiodzrone-induced thyrotoxicosis (AIT) depends on AIT type. Eur J Endocrinol 171:363-368, 2014

340d. Yagishita A, Hachiya H, Kawabata M, Nakamura T, Sugiyama K, Tanaka Y, Sasano T, Iasobe M, Hirao K Amiodarone-induced thyrotoxicosis late after amiodarone withdrawal Circ J 77:2898-2903, 2013

340e.      Houghton SG, Farley DR, Brennan MD, van Heerden JA, Thompson GB, Grant CS Surgical management of amiodarone-associated thyrotoxicosis: Mayo Clinic experience World J Surg 28:1083-1087, 2004

  1. Schussler GC: Diazepam competes for thyroxine binding. J Pharmacol Exp Ther 178:204-209, 1971.
  2. Molholm Hansen J, Skovsted L, Birk Lauridsen U, Kirkegaard C, Stersbaek-Nielsen K: The effect of diphenylhydantoin on thyroid function. J Clin Endocrinol Metab 39:785, 1974.
  3. Heyma P, Larkins RG, Perry-Kenne D, Peter CT, Ross D, Sloman JG: Thyroid hormone levels and protein binding in patients on long term diphenylhydantoin treatment. Clin Endocrinol 6:369, 1977.
  4. Cavalieri RR, Gavin LA, Wallace A, et al: Serum thyroxine, free T4, triiodothyronine, and reverse-T3 in diphenylhydantoin treated patients. Metabolism 28:1161-1165, 1979.

344a.      Surks MI, DeFesi CR: Normal serum free thyroid hormone concentrations in patients treated with phenytoin or carbamazapine, a paradox resolved. JAMA 275:1495-8,1996

344b.      Tiihonen M, Liewendahl K, Waltimo O, Ojala M, Valimaki M: Thyroid status of patients receiving long-term anticonvulsant therapy assessed by peripheral parameters: a placebo-controlled thyroxine therapy trial. Epilepsia 36:1118-25,1995

  1. Hansen JM, Skovsted L, Lauridsen UB, Kirkegaard C, Siersbaek-Nielsen K: The effect of diphenylhydantoin on thyroid function. J Clin Endocrinol Metab 39:785-, 1974.
  2. Blackshear JL, Schultz AL, Napier JS, Stuart DD: Thyroxine replacement requirements in hypothyroid patients receiving phenytoin. Ann Intern Med 99:341-359, 1983.
  3. Oppenheimer JH, Shapiro HC, Schwartz HL, Surks MI: Dissociation between thyroxine metabolism and hormonal action in phenobarbital-treated rats. Endocrinology 88:115, 1971.

347a.      Brookstaff RC, Murphy VA, Skare JA, Minnema D, Sangiri U, Parkinson A: Effects of doxylamine succinate on thyroid hormone balance and enzyme induction in mice. Toxicol Appl Pharmacol 141:584-94,1996.

  1. Cavalieri RR, Sung LC, Becker CE: Effects of phenobarbital on thyroxine and triiodothyronine kinetics in Graves' disease. J Clin Endocrinol Metab 37:308-316:1973.

348a.      Rootwelt K, Ganes T, Johannessen SI: Effects of carbamazapine, phenytoin and phenobarbitone on serum levels of thyroid hormones and thyrotropin in humans. Scan J Clin Lab Invest   38:731,1978.

  1. Wartofsky L, Dimond RC, Noel GL, et al: Failure of propranolol to alter thyroid iodine release, thyroxine turnover, or the TSH and PRL responses to thyrotropin-releasing hormone in patients with thyrotoxicosis. J Clin Endocrinol Metab 41:485-490, 1975.
  2. Woolf PD, Lee LA, Schalch DS: Adrenergic manipulation and thyrotropin releasing hormone (TRH)-induced thyrotropin (TSH) release. J Clin Endocrinol Metab 35:616-, 1972.
  3. Faber J, Friis T, Kirkegaard C, et al: Serum T4, T3, and reverse T3 during treatment with propranolol in hyperthyroidism, L-T4 treated myxedema and normal man. Horm Metab Res 11:34-36, 1979.
  4. Faber J, Kirkegaard C, Lumholtz IB, Siersbaek-Nielsen K, Friis T: Variations in serum T3, rT3, 3,3'-diiodothyronine and 3',5'-diiodothyronine induced by acute myocardial infarction and propranolol. Acta Endocrinol 94:341, 1980.
  5. Murchison LE, How J, Bewsher PD: Comparison of propranolol and metoprolol in the management of hyperthyroidism. Br J Clin Pharmacol 8:581, 1979.
  6. How ASM, Khir AN, Bewsher PD: The effect of atenolol on serum thyroid hormones in hyperthyroid patients. Clin Endocrinol 13:299-302, 1980.
  7. Hadden DR, Bell TK, McDevitt DG, Shanks RG, Montgomery DAD, Weaver JA: Propranolol and the utilization of radioiodine by the human thyroid gland. Acta Endocrinol 61:393, 1969.
  8. Wilson WR, Theilen ED, Fletcher FW: Propranolol and its effects in thyrotoxicosis on heart at rest or exercise. J Clin Invest 43:1697, 1964.
  9. Das G, Krieger M: Treatment of thyrotoxic storm with intravenous administration of propranolol. Ann Intern Med 70:985, 1969.
  10. Canary JJ, Shaaf M, Duffy BJ Jr., Kyle LH: Effects of oral and intramuscular administration of reserpine in thyrotoxicosis. N Engl J Med 257:435, 1957.
  11. Waud DR, Kattegoda SR, Krayer O: Threshold dose and time course of norepinephrine depletion of mammalian heart by reserpine. J Pharmacol Exp Ther 124:340, 1958.
  12. Goldberg RC, Wolff J, Greep RO: Studies on the nature of the thyroid-pituitary interrelationship. Endocrinology 60:38, 1957.
  13. Goldberg RC, Wolff J, Greep RO: The mechanism of depression of plasma protein-bound iodine by 2,4-dinitrophenol. Endocrinology 56:560-566, 1955.
  14. Lardy HA, Wellman H: Oxidative phosphorylations: Role of inorganic phosphate and acceptor systems in control of metabolic rates. J Biol Chem 195:215-224, 1952.
  15. Escobar del Rey F, Morreale de Escobar G: Studies on the peripheral disappearance of thyroid hormone. IV. The effect of 2,4-dinitrophenol on the 131I distribution in thyroidectomized, L-thyroxine maintained rats, 24 hours after the injection of 131I-labeled L-thyroxine. Acta Endocrinol 29:161-175, 1958.
  16. Escobar del Rey F, Morreale de Escobar G: Studies on the peripheral disappearance of thyroid hormone. V. The effect of 2,4-dinitrophenol on the variations of the 131I distribution pattern with time, after the injection of 131I-labeled L-thyroxine into thyroidectomized, L-thyroxine maintained rats. Acta Endocrinol 29:176, 1958.
  17. Cutting CC, Tainter ML: Comparative effects of dinitrophenol and thyroxin on tadpole metamorphosis. Proc Soc Exp Biol Med 31:97-100, 1933.
  18. Reichlin S: Regulation of the hypophysiotropin secretion of the brain. Arch Intern Med 135:1350, 1975.
  19. Morley JE: Neuroendocrine control of thyrotropin secretion. Endocr Rev 2:396-436, 1981.
  20. Kaptein EM, Spencer CA, Kamiel MB, Nicoloff JT: Prolonged dopamine administration and thyroid hormone economy in normal and critically ill subjects. J Clin Endocrinol Metab 51:387-393, 1980.

368a.      deZegher F, Van den Bershe G, Dumoulin M, Gewillig M, Daenen W, Deviliger H: Dopamine suppresses thyroid-stimulating hormone secretion in neonatal hypothyroidism. Acta Paediatr 84:213-4,1995

368b.      Martignoni E, Horowski R, Liuzzi A, Costa A, Dallabonzana D, Cozzi R, Attanasio R, Rainer E, Nappi G. Clin Neuropharmacol 19:72-80, 1996

368c.      Samuels MH, Kramer P: Effects of metoclopramide on fasting-induced TSH suppression. Thyroid 6:85-9,1996

  1. Benker G, Zäh W, Hackenberg K, Hamburger B, Gunnewig H, Reinwein D: Long-term treatment of acromegaly with bromocryptine: Postprandial HGH levels and response to TRH and glucose administration. Horm Metab Res 8:291, 1976.
  2. Köbberling J, Darrach A, Del Pozo E: Chronic dopamine receptor stimulation using bromocryptine: Failure to modify thyroid function. Clin Endocrinol 11:367-370, 1979.
  3. Foord SM, Peters J, Scanlon MF, Rees-Smith B, Hall R: Dopaminergic control of TSH secretion in isolated pituitary cells. FEBS Lett 121:257, 1980.
  4. Heinen E, Herrmann J, Konigshausen T, Kruskemper HL: Secondary hypothyroidism in severe non-thyroidal illness? Horm Metab Res 13:284-288, 1981.
  5. Weintraub BD, Gershengorn MC, Kourides IA, Fein H: Inappropriate secretion of thyroid stimulating hormone. Ann Intern Med 95:339-351, 1981.
  6. Chanson P, Weintraub BD, Harris AG: Octreotide therapy for thyroid hormone-stimulating hormone-secreting pituitary adenomas. Ann Int Med 119:236-240, 1994.
  7. Fernandez-Soto L, Gonzalez A, Escobar_jimenez F, Vazquez R, Ocete E, Olea N, Salmeron J: Increased risk of autoimmune thyroid disease in hepatitis C vs hepatitis B before, during and after discontinuing interferon therapy. Arch Intern Med 158:1445-8, 1998.
  8. Amenomori M, Mori T, Fukuda Y, Sugawa H, Nishida N, Furukawa M, Kita R, Sando T, Komeda T, Nakao K: Incidence and characteristics of thyroid dysfunction following interferon therapy in patients with chronic hepatitis C. Intern Med 37:246-52, 1998.

376a.      Hamnvik OR, Larsen PR, Marquses E Thyroid dysfunction from antineoplastic agents J Natl Cancer Inst 103:1572-1587, 2011

376b.      Torino F, Barnabel A, Paragliola R, Baldelli R, Appetecchuia M, Corsello SM   Thyroid dysfunction as an unintended side effect of anticancer drugs Thyroid 231345-1366, 2013

  1. Koh LK, Greenspan FS, Yeo PP: Interferon-alpha induced thyroid dysfunction: three clinical presentations and a review of the literature. Thyroid 7:891-6,1997.

377a.      Daniels GH, Vladic A, Brinar V, Zavalishin, Valente W, Oyuela P, Plamer J, Margolin DH   Alemtuzumab-related thyroid dysfunction in a phase 2 trial of patients with relapsing-remitting multiple sclerosis.   J Clin Endocrinol Metab   99:80-89, 2014

377b Ribas A Releasing the brakes on cancer immunotherapy NEJM 373:1490-1492

377c       Tarhini A Immune-mediated adverse events associated with Ipuilmumab CTLA-4 blockade therapy: The underlying mechanisms and clinical management Scientifica Artcile 857519 2013

377d Corsello SM, Barnabel A, Marchetti P, Vecchis LD, Salvatori R, Torino F   Endocrine side effects induced by immune checkpoint inhibitors J Clin Endocrinol Metab 98:1361-1375, 2013

377e Faje AT, Sullivan R, Lawrence D, Tritos NA, Fadden R, Klibanski A and Nachitgall L: Ipilumumab-induced hypophyisitis: a detailed longitudinal analysis in a large cohort of patients with metastatic melanoma.   J Clin Endocrinol Metab 99:4078-4085, 2014

377f       Ryder M, Callahan M, Postow MA, Wolchok J, Fagin JA   Endocrine-related adverse events following ipilimumab in patients with advanced melanoma: a comprehensive retrospective review from a single institution   Endocrine-Related Cancer   21: 371-381, 2014

377g      Ansell SM, Leshokin AM, Borrello I, Halwani A, Scott EC, Gutierrez, M, Schuster SJ, Millenosn MM, Cattry D, Freeman GJ, Rodig SJ, Chapuy, B, Ligon AH, Zhu, L, Grosso JF, Kim SY, Timmerman JM, Shipp MA, Armand P PD-1 blockade with nivolumab in relapsed or refractory Hodgkin’s lymphoma N Engl J Med 372:311-319, 2014

377h   Orlov S, Salari F, Kashat L and Walfish PG: Case reports: Induction of painless thyroiditis in patients receiving programmed death 1 receptor immunotherapy for metastatic malignancies.   J Clin Endocrinol Metab 100:1738-1741, 2015

377i       Naidoo J, Page DB, Connell LC, Schindler L|KJ, Lacouture ME, Postow MA, Wolchok JD Toxicities of the anti-PD-1 and anti-PD-L1 immune checkpoint antibodies   Ann Onc 26:2375-2391, 2015

  1. Schuppert F, Rambusch E, Kirchner H, Atzpodien J, Kohn LD, von zur Muhllen A: Patients treated with interferon-alpha, interferon-beta, and interleukin-2 have a different autoantibody pattern than patients suffering from endogenous thyroid disease.   Thyroid 7:837-42,1997.
  2. Krouse RS, Yoral RE, Heywood G, Weintraub BD, White DE, Steinberg SM, Rosenberg SA, Schwartzentruber DJ: Thyroid dysfunction in 281 patients with metastatic melanoma or renal carcinoma treated with interleukin-2 alone. J Immunother Emphasis Tumor Immunol 18:272-8,1995.

379a.      Prummel MF, Laurberg P: Interferon- and autoimmuine thyroid disease. Thyroid 13:547-551, 2003

  1. Corssmit EP, Heyligenberg R, Endert E, Sauerwein HP, Romijn JA: Acute effects of interferon-alpha on thyroid hormone metabolism in healthy men. J Clin Endocrinol Metab 80:3140-4, 1995.
  2. Witske O, Winterhagen T, Saller, B, Rogenbuck U, Lehr I, Phillipp T, Mann K, Reinhardt W: Transient stimulatory effect on pituitary-thyroid axis in patients treated with interleukin-2 Thyroid 11:665-670, 2001
  3. Desai J, Yassa L, Marqusee E, George S, Frates MC, Chen MH, Morgan JA, Dychter SS, Larsen PR, Demetri GD, Alexander EK: Hypothyroidism after suntinib treatment for patients with gastrointestinal stromal tumors. Ann Intern Med 145:660-4, 2006

382a   Rogiers A, Wolter P, de Beeck KO, Thijs M Decallonne B, Schoffski P:   Shrinkage of thyroid volume in sunitinib-treated patients with renal cell carcinoma: A potential marker of irreversible thyroid dysfunction? Thyroid 20:317-322, 2010

  1. Wong E, Rosen LS, Mulay M, Vanvugt A, Dinolfo M, Tomoda C, Sugawara M, Hershman JM:   Sunitinib induces hypothyroidism in advanced cancer patients and may inhibit thyroid peroxidase activity.  Thyroid   17:351-5,2007

383b     Makita N, Megumi M, Fujita T, IIri T: Sunitinib induces hypothyroidism with a markedly reduced vascularity. Thyroid 20:323-326, 2010

383c     Hershman JM, Liwanpo L: How does sunitinib cause hypothyroidism? Thyroid 20:243-244 2010

383d     Iavarone M, Perrino M, Vigano M, Beck-Peccoz P, Fugazzola L: Sorafenib-induced destructive thyroiditis. Thyroid 20:1043-1044, 2010

383e     Tamaskar I, Bukowski R, Elson P, Ioachimescu AG, Wood L et al Thyroid function test abnormalities in patients with metastatic renal cell carcinoma treated with sorafenib.   Ann Oncol 19:265-268, 2008

383f     Abdulrahman RM, Verloop H, Hoftijzer H, Verburg E, Hovens GC et al: Sorafenib-induced hypothyroidism is associated with increased type 3 deiodination. JCEM 95:3758-3762, 2010

  1. DeGroot JW, Zonnenberg BA, Plukker JT, van DerGraaf WT, Links TP: Imatinib induces hypothyroidism in patients receiving levothyroxine. Clin Phamacol Ther 78:4334-8, 2005

384a     Kim TD, Schwarz M, Nogai H, Grille P, Westermann J et al: Thyroid dysfunction caused by second generation tyrosine kinase inhibitors in Philadelphia chromosome-positive chronic myeloid leukemia.   Thyroid 20:1209-1214, 2010

  1. Sherman SI, Gopal J, Haugen BR, Chiu AC, Whaley K, Nowlakha P, Duvic M: Central hypothyroidism associated with retinoid X receptor selective ligands. N Engl J Med 340:1075-9, 1999
  2. Golden WM, Weber KB, Hernandez TL, Sherman SI, Woodmansee WW, Haugen BR: Single-dose rexinoid rapidly and specifically suppresses serum thryotropin in normal subjects. J Clin Endocrinol Metab 92:124-30, 2007
  3. Smit JWA, Stokkel MPM, Pereira AM, Romijn JA, Visser TJ: Bexarotene-induced hypothyroidism: Bexarotene stimulates the peripheral metabolism of thyroid hormones.   J Clin Endocrinol Metab 92:2496-99, 2007

388         Neumann S, Raaka BM, Gershengon MC: Human TSH receptor ligands as pharmacologic probes with potential clinical application. Expert Rev Endocrinol Metab 4:669-671, 2009

388b.   Lupoli R, Di Minno A, Tottora A, Amborsino P, Lupoli GA, Di Minno MND   Effects of treatment with metformin on TSH Levels: A meta-analysis of literature studies. J Clin Endo Metab 99: E143-E148, 2014

389         Neumann S, Huang W, Titus S, Krause G, Kleinau G et al: Small-molecule agonists for the thyrotropin receptor stimulate thyroid function in human thyrocytes and mice. PNAS 106:12471-12476, 2009

390         Neumann S, Eliseeva E, McCoy JG, Napolitano G, Giuliani C et al: A new small-molecule antagonist inhibits Graves’ disease antibody activation of the TSH receptor. J Clin Endocrinol Metab 96:548-554, 2011

391         Piehl S, Hoefig CS, Scanlan TS, Kohrle J: Thyronamines – Past, present and future. Endocrien Reviews 32:64-80, 2011

392         Vigersky RA, Filmore-Nassar A, Glass AR: Thyrotropin suppression by metformin.   J Clin Endocrinol Metab 91:225-7, 2006

393       John-Kalarickal J, Pearlman G, Carlson HE: New medications which decrease levothyroxine absorption. Thyroid 17:763-765, 2007

394     Weitzman SP, Ginsburg KC, Carlson HE:   Coleselvam hydrochloride and lanthanum carbonate interfere with the absorption of levothyroxine. Thyroid 19:77-79, 2009

395       Benvenega S, Bartolone L, Pappalardo MA, Russo A, Lapa D Giogianni G et al: Altered intestinal absorption of l-thyroxine caused by coffee.   Thyroid 18:293-301, 2008

396       Badros AZ, Siegel E, Bodenner D, Zangarui M, Zeldis J, Barlogie B, Tricot G: Hypothyroidism in patients with multiple myeloma following treatment with thalidomide. AM J Med 112:412-413, 2002

  1. Kwok JS, Chan IH, Chan MH Biotin interference of TSH and free thyroid hormone measurement. Pathology 44:278-285, 2012
  2. Wijeratne NG, Doery JCG, Lu ZX Positive and negative interference in immunoassays following biotin ingestion: a pharmacokinetic study. Pathology 44:674-675, 2012
  3. Kummer S, Hermsen D, Distelmaier F Biotin treatment mimicking Graves’ Disease   New Engl J Med   375:704-706, 2016
  4. Elston MS, Seghal S, Du Toit S, Yarndley T Conaglen JV Factitious Graves’ disease due to biotin immunoassay interference – A case and review of the literature. J Clin Endocrinol Metab 101:3251-3255,2016

 

 

Multinodular Goiter

ABSTRACT

Multinodular goiter (MNG) is the most common of all the disorders of the thyroid gland. MNG is the result of the genetic heterogeneity of follicular cells and apparent acquisition of new cellular qualities that become inheritable. Nodular goiter is most often detected simply as a mass in the neck, but sometimes an enlarging gland produces pressure symptoms. Hyperthyroidism develops in a large proportion of MNGs after a few decades, frequently after iodine excess. Diagnosis is based on the physical examination. Thyroid function test results are normal, or indicate subclinical or overt hyperthyroidism. Imaging procedures are useful to detect details such as distortion of the trachea, and to provide an estimation of the volume before and after therapy. From 4 to 17% of MNGs fulfill the criteria of malignant change, however, the majority of these lesions are not lethal. If a clinical and biochemically euthyroid MNG is small and produces no symptoms, treatment is controversial. T4 given to shrink the gland or to prevent further growth is effective in about one third of patients. If the clinically euthyroid goiter is unsightly, shows subclinical hyperthyroidism or is causing pressure symptoms, treatment with ¹³¹I preceded by recombinant human TSH is successful but causes hypothyroidism in varying degrees. This treatment can lead to 45-65% shrinkage of the MNG, even if in an intrathoracic position, with a relatively low cost, thus it is considered a good alternative to surgery. However, surgery is an acceptable option. The efficacy of T4 treatment after surgery, to prevent regrowth, is debatable although frequently usedt. For complete coverage of all related aeas of Endocrinology, please visit our on-line FREE web-text, WWW.ENDOTEXT.ORG.

INTRODUCTION

The normal thyroid gland is a fairly homogenous structure, but nodules often

form within its substance. These nodules may be only the growth and fusion of localized colloid-filled follicles, or more or less discrete adenomas, or cysts. Nodules larger than 1 cm may be detected clinically by palpation. Careful examination discloses their presence in at least 4% of the general population. Nodules less than 1 cm in diameter not clinically detectable unless located on the surface of the gland, are much more frequent. The terms adenomatous goiter, nontoxic nodular goiter, and colloid nodular goiter are used interchangeably as descriptive terms when a multinodular goiter is found.

INCIDENCE

The incidence of goiter, diffuse and nodular, is very much dependent on the status of iodine intake of the population. In areas of iodine deficiency, goiter prevalence may be very high and especially in goiters of longstanding, multinodularity develops frequently (Figure 17-1). The incidence of multinodular goiter in areas with sufficient iodine intake has been documented in several reports (1-10). In a comprehensive population survey of 2,749 persons in northern England, Tunbridge et al (1) found obvious goiters in 5.9% with a female/male ratio of 13:1. Single and multiple thyroid nodules were found in 0.8% of men and 5.3% of women, with an increased frequency in women over 45 years of age. Routine autopsy surveys and the use of sensitive imaging techniques produce a much higher incidence. In three reports nodularity was found in 30% to 50% of subjects in autopsy studies, and in 16% to 67% in prospective studies of randomly selected subjects on ultrasound (2). In Framingham the prevalence of multinodular goiter as found in a population study of 5234 persons over 60 years was 1% (3). Results from Singapore show a prevalence of 2.8% (4). In an evaluation in 2,829 subjects, living in southwestern Utah and Nevada (USA, between 31 and 38 years) of age, 23% had non-toxic goiter, including 18 single nodules, 3 cysts, 38 colloid goiters and 7 without a histological diagnosis. No mention was made of multinodular goiters, although some might have been present in the colloid and unidentified group (5). In general, in iodine sufficient countries the prevalence of multinodular goiter is not higher than 4% (6). In countries with previous deficiency that was corrected by universal salt iodination, elderly subjects may have an incidence of, approximately, 10% of nodular and multinodular goiter, attributed to lack of nutritional iodine in early adult life (7).

ETIOLOGY

The first comprehensive theory about the development of multinodular goiter was proposed by David Marine (8) and studied further by Selwyn Taylor (9), and can be considered one of the classics in this field. Nodular goiter may be the result of any chronic low-grade, intermittent stimulus to thyroid hyperplasia. Supporting evidence for this view is circumstantial. David Marine first developed the concept, that in response to iodide deficiency, the thyroid first goes through a period of hyperplasia as a consequence of the resulting TSH stimulation, but eventually, possibly because of iodide repletion or a decreased requirement for thyroid hormone, enters a resting phase characterized by colloid storage and the histologic picture of a colloid goiter. Marine believed that repetition of these two phases of the cycle would eventually result in the formation of nontoxic multinodular goiter (8). Studies by Taylor of thyroid glands removed at surgery led him to believe that the initial lesion is diffuse hyperplasia, but that with time discrete nodules develop (9).

By the time the goiter is well developed, serum TSH levels and TSH production rates are usually normal or even suppressed (10). For example, Dige-Petersen and Hummer evaluated basal and TRH-stimulated serum TSH levels in 15 patients with diffuse goiter and 47 patients with nodular goiter (11). They found impairment of TRH-induced TSH release in 27% of the patients with nodular goiter, suggesting thyroid autonomy, but in only 1 of the 15 with diffuse goiter. Smeulers et al (12), studied clinically euthyroid women with multinodular goiter and found that there was an inverse relationship between the increment of TSH after administration of TRH, and size of the thyroid gland (Figure 17-1). It was also found that, while being still within the normal range, the mean serum T3 concentration of the group with impaired TSH secretion was significantly higher than the normal mean, whereas the mean value of serum T4 levels was not elevated (12). These and other results (13) are consistent with the hypothesis that a diffuse goiter may precede the development of nodules. They are also consistent with the clinical observation that, with time, autonomy may occur, with suppression of TSH release, even though such goiters were originally TSH dependent.

Figure 17-1. Relationship of TSH (after 400 mg TRH i.v.) and thyroid weight (g) in 22 women with clinically euthyroid multinodular goiter (with permission ref. 12)

Figure 17-1. Relationship of TSH (after 400 mg TRH i.v.) and thyroid weight (g) in 22 women with clinically euthyroid multinodular goiter (with permission ref. 12)

Comprehensive reviews about insights into the evolution of multinodular goiter have been published by Studer and co-workers (14-16). An adapted summary of the major factors that are discussed is presented in Table 17-1 and will be referred to in the discussion that follows.

Table 17-1. Factors that may be involved in the evolution of multinodular goiter.

PRIMARY FACTORS

  • Functional heterogeneity of normal follicular cells, most probably due to genetic and acquisition of new inheritable qualities by replicating epithelial cells. Gender (women) is an important factor.
  • Subsequent functional and structural abnormalities in growing goiters.

SECONDARY FACTORS

  • Elevated TSH       (induced by iodine deficiency, natural goitrogens, inborn errors of thyroid hormone synthesis)
  • Smoking, stress, certain drugs
  • Other thyroid-stimulating factors (IGF-1 and others)
  • Endogenous factor (gender)

PRIMARY FACTORS

Genetic heterogeneity of normal follicular cells and acquisition of new inheritable qualities by replicating epithelial cells. (Figure 17-2).It has been shown cells of many organs, including, the thyroid gland, are often polyclonal, rather than monoclonal of origin. Also from a functional aspect it appears that through developmental processes the thyroid epithelial cells forming a follicle are functionally polyclonal and possess widely differing qualities regarding the different biochemical steps leading to growth and to thyroid hormone synthesis like e.g. iodine uptake (and transport), thyroglobulin production and iodination, iodotyrosine coupling, endocytosis and dehalogenation. As a consequence there is some heterogeneity of growth and function within a thyroid and even within a follicle Studer et al (14-16) demonstrated the existence of monoclonal and polyclonal nodules in the same multinodular gland. They analyzed 25 nodules from 9 multinodular goiters and found 9 to be polyclonal and 16 monoclonal. Three goiters contained only polyclonal nodules and 3 contained only monoclonal nodules. In 3 goiters poly- and monoclonal nodules coexisted in the same gland (17).

Figure 17-2. Heterogeneity of morphology and function in a human multinodular goiter. Autoradiographs of two different areas of typical multindular euthyroid human goiter excised after administration of radioiodine tracer to the patient. There are enormous differences of size, shape and function among the individual follicles of the same goiter. Note also that there is no correlation between the size or any other morphological hallmark of a single follicle and its iodine uptake. (with permission ref.15).

Figure 17-2. Heterogeneity of morphology and function in a human multinodular goiter. Autoradiographs of two different areas of typical multindular euthyroid human goiter excised after administration of radioiodine tracer to the patient. There are enormous differences of size, shape and function among the individual follicles of the same goiter. Note also that there is no correlation between the size or any other morphological hallmark of a single follicle and its iodine uptake. (with permission ref.15).

Newly generated cells may acquire qualities not previously present in mother cells. These qualities could subsequently be passed on to further generations of cells. A possible example of this process is the acquired abnormal growth pattern that is reproduced when a tissue sample is transplanted into a nude mouse (16). Other examples are acquired variable responsiveness to TSH (13). These changes may be related to mutations in oncogenes which do not produce malignancy per se, but that can alter growth and function. An example of acquisition of genetic qualities is the identification in the last few years of constitutively activating somatic mutations not only in solitary toxic adenoma, but also in hyperfunctioning nodules of toxic multinodular goiters (18). So far these mutations in MNG have only been found in the TSH-receptor (TSHR) gene, and not in the Gs-alpha gene. Different somatic mutations are found in exon 9 and 10 of the TSHR gene and the majority of mutations that are present in toxic adenomas are also found in toxic nodules in multinodular goiter (19-21).

Genes associated with multinodular goiter

In contrast to sporadic goiters, caused by spontaneous recessive genomic variation, most cases of familial goiter present an autosomal dominant pattern of inheritance, indicating predominant genetic defects. Gene-gene interactions or various polygenic mechanisms (i.e. synergistic effects of several variants or polymorphisms) could increase the complexity of the pathogenesis of nontoxic goiter and offer an explanation for its genetic heterogeneity (22-26). A strong genetic predisposition is indicated by family and twin studies (27-29). Thus, children of parents with goiter have a significantly higher risk of developing goiter compared with children of nongoitrous parents (24). The high incidence in females and the higher concordance in monozygotic than in dizygotic twins also suggested a genetic predisposition (24). Moreover, there is preliminary evidence of a positive family history for thyroid diseases in those who have postoperative relapse of goiter, which can occur from months to years after surgery.

Defects in genes that play an important role in thyroid physiology and thyroid hormone synthesis could predispose to the development of goiter, especially in case of borderline or overt iodine deficiency. Such defects could lead to dyshormonogenesis as an immediate response, thereby indirectly explaining the nodular transformation of the thyroid as late consequences of dyshormonogenesis, as a form of maladaptation (12). The genes that encode the proteins involved in thyroid hormone synthesis, such as the thyroglobulin-gene (TG-gene), the thyroid peroxidase-gene (TPO-gene), the sodium – iodide – symporter-gene (SLC5A5), the Pendred syndrome-gene (SLC26A4), the TSH receptor-gene (TSH-R-gene), the iodotyrosine deiodinase (DEHAL 1) and the thyroid oxidase 2 gene3 (DUOX2) are convincing candidate genes in familial euthyroid goiter (30). Originally, several mutations in these genes were identified in patients with congenital hypothyroidism (30). However, in cases of less severe functional impairment, with can still be compensated, a contribution of variants of these genes in the etiology of nontoxic goiter is possible. 

Linkage studies

A genome-wide linkage analysis has identified a candidate locus, MNG1 on chromosome 14q31, in a large Canadian family with 18 affected individuals (31). This locus was confirmed in a German family with recurrent euthyroid goiters (32). A dominant pattern of inheritance with high penetrance was assumed in both investigations. Moreover, a region on 14q31 between MNG1 and the TSH-R-gene was identified as a potential positional candidate region for nontoxic goiter (33). However, in an earlier study the TSH-R-gene was clearly excluded (31). Furthermore, an X-linked autosomal dominant pattern and linkage to a second locus MNG2 (Xp22) was identified in an Italian pedigree with nontoxic familial goiter (34). To identify further candidate regions, the first extended genome-wide linkage analysis was performed to detect susceptibility loci in 18 Danish, German and Slovakian euthyroid goiter families (35). Assuming genetic heterogeneity and a dominant pattern of inheritance, four novel candidate loci on chromosomes 2q, 3p, 7q and 8p (36) were identified . An individual contribution was attributable to four families for the 3p locus and to 1 family to each of the other loci, respectively. On the basis of the previously identified candidate regions and the established environmental factors, nontoxic goiter can consequently be defined as a complex disease. However, for this first time a more prevalent putative locus, present in 20% of the families investigated, was identified (35).

The candidate region on 3p (37) suggests a dominant pattern of inheritance for goiter. However, whereas linkage studies are suitable for the detection of candidate genes with a strong effect it is possible to miss weak genetic defects of first-line candidate gene-variants or of novel genes by linkage studies. Moreover, it is conceivable that the sum of several weak genetic variations in different genomic regions could lead to goiter predisposition. Therefore, the widely accepted risk factors such as iodine deficiency, smoking, old age, and female gender are likely to interact with and / or trigger the genetic susceptibility (22).

Mutagenesis leading to multinodular goiter

Most goiters become nodular with time. (Figure 17-3) From animal models of hyperplasia caused by iodine depletion (38) we have learned that besides an increase in functional activity a tremendous increase in thyroid cell number occurs. These two events likely induce a number of mutation events. It is known that thyroid hormone synthesis goes along with increased H2O2 production and free radical formation with may damage genomic DNA and cause mutations. Together with a higher spontaneous mutation rate, a higher replication rate will more often prevent mutation repair and increase the mutation load of the thyroid, thereby also randomly affecting genes essential for thyrocyte physiology. Mutations that confer a growth advantage (e.g. TSH-R mutations) very likely initiate focal growth. Hence, autonomously functioning thyroid nodules (AFTNs) are likely to develop from small cell clones that contain advantageous mutation as shown for the TSH-R in “hot” microscopic regions of euthyroid MNG (18).

Epidemiologic studies, animal models and molecular/genetic data outline a general theory of nodular transformation. Based on the identification of somatic mutations and the predominant clonal origine of AFTNs and cold thyroid nodules (CTNs) the following sequence of events could lead to thyroid nodular transformation in three steps. First, iodine deficiency, nutritional goitrogens or autoimmunity cause diffuse thyroid hyperplasia (39-41). Secondly, at this stage of thyroid hyperplasia, increased proliferation together with a possible DNA damage due to H2O2 action causes a higher mutation load, i.e. a higher number of cells bearing mutations. Some of these spontaneous mutations confer constitutive activation of the cAMP cascade (e.g. TSH-R mutations) which stimulates growth and function. Finally, in a proliferating thyroid, growth factor expression (e.g. insulin-like growth factor 1   [IGF-1], transforming growth factor ß [TGF-ß], or epidermal growth factor [EGF]) is increased (42-51). As a result of growth factor co-stimulation most cells divide and form small clones. After increased growth factor expression ceases, small clones with activating mutations will further proliferate if they can achieve self-stimulation. They could thus form small foci, which could develop into thyroid nodules. This mechanism could explain AFTNs by advantageous mutations that both initiate growth and function of the affected thyroid cells as well as CTNs by mutations that stimulate proliferation only. Moreover, nodular transformation of thyroid tissue due to TSH secreting pituitary adenomas, nodular transformation of thyroid tissue in Graves´ disease and in goiters of patients with acromegaly could follow a similar mechanism, because thyroid pathology in these patients is characterized by early thyroid hyperplasia.

As an alternative to the increase of cells mass, and as illustrated by those individuals who do not develop a goiter when exposed to iodine deficiency, the thyroid might also adapt to iodine deficiency without extended hyperplasia. Although the mechanism that allows this adaptation is poorly understood, data from a mouse model suggests an increase of mRNA expression of TSH-R, NIS and TPO in response to iodine deficiency, which might be a sign of increased iodine turnover in the thyroid cell in iodine deficiency. Moreover, expansion of the thyroid microvasculature, caused by up regulation of vascular endothelial growth factor and other proangiogenic factors, could be an additional mechanism that might help the thyroid to adapt to iodine deficiency (52).

SECONDARY FACTORS

The secondary factors discussed below stimulate thyroid cell growth and / or function and, because of differences in cellular responsiveness that are presumed to exist, aggravate the expression of heterogeneity which leads to further growth and focal autonomic function of the thyroid gland. Local necrosis, cyst formation sometimes with bleeding and fibrosis may be the anatomical end stage of such processes (Figure 17-3).

Figure 17-3: Mild iodine deficiency associated or not with smoking, presence of natural goitrogenic, drugs, familial goiter, genetic markers and gender (women) will decrease the inhibition of serum T4 on the pituitary thyrotrophs. Increased TSH production will cause diffuse goiter followed by nodule formation. Finally, after decades of life, a large multinodular goiter is present with cystic areas, hemorrhage, fibrosis and calcium deposits.

Figure 17-3: Mild iodine deficiency associated or not with smoking, presence of natural goitrogenic, drugs, familial goiter, genetic markers and gender (women) will decrease the inhibition of serum T4 on the pituitary thyrotrophs. Increased TSH production will cause diffuse goiter followed by nodule formation. Finally, after decades of life, a large multinodular goiter is present with cystic areas, hemorrhage, fibrosis and calcium deposits.

Iodine Deficiency

Stimulation of new follicle generation seems to be necessary in the formation of simple goiter. (Figure 17-3) Evidence accumulated from many studies indicates that iodine deficiency or impairment of iodine metabolism by the thyroid gland, perhaps due to congenital biochemical defects, may be an important mechanism leading to increases in TSH secretion (30,53). Since in experimental animals the level of iodine per se may modulate the response of thyroid cells to TSH, this is an additional mechanism by which relatively small increases in serum TSH level may cause substantial effects on thyroid growth in iodine-deficient areas (53). It was found that the thyroidal iodine clearance of patients with nontoxic nodular goiter was, on overage, higher than that in normal persons (Fig. 17-3). This finding was interpreted as a reflection of a suboptimal iodine intake by such patients. When data published from various major cities in Western Europe, regarding thyroid volume and iodine excretion are put together (54) and inverse relation is found between urinary iodine excretion and thyroid volume (Fig. 17-4). Physiologic stresses, such as pregnancy, may increase the need for iodine and require thyroid hypertrophy to increase iodine uptake that might otherwise satisfy minimal needs. An elevated renal clearance   of iodine   occurs during normal pregnancy (24). It has been suggested that in some patients with endemic goiter there are similar increases in renal iodine losses (53). Increased need for thyroxin during pregnancy may also lead to thyroid hypertrophy when iodine intake as limited. Iodide need in pregnancy is increased by increased iodide loss through the kidneys, but also because of significant transfer of thyroid hormone from the mother to the fetus (24). In areas of moderate iodine intake, thyroid volume increase is predominantly affected by a higher HCG serum concentration during the first trimester of pregnancy, and by a slightly elevated serum TSH level present at delivery (24). Finally mutations in the thyroglobulin gene may impair the efficiency of thyroid hormone synthesis and release, leading to a decreased rate of inhibition of TSH at pituitary level. The relatively high TSH released from the thyrotrophs will continuously stimulate the thyroid gland growth (55).

Figure 17-4.Relationship between nontoxic goiter and thyroidal iodine clearance.

Figure 17-4. Relationship between nontoxic goiter and thyroidal iodine clearance.

Figure 17-5. Correlation between thyroid volume and urinary iodine excretion in normal population from various areas.

Figure 17-5. Correlation between thyroid volume and urinary iodine excretion in normal population from various areas.

Natural occurring goitrogens

Patients occasionally have thyroid enlargement either because of goitrogenic substances in their diet or because of drugs that have been given for other conditions (53). Feeding rats with minute doses of a natural goitrogen over many months will result in the same kind of response. Similar results have been obtained using combinations of the three most prevalent goitrogens contained in cabbage. The explanation for the effect of such substances is that the goitrogen is much more effective at the level of iodothyronine synthesis than at earlier steps in hormone production such as iodide trapping. Thus, the RAIU may be high, but with a block in hormone synthesis the stage would be set for the production of a goiter. This possibility remains to be proved in humans, but one might surmise that, if true, it would operate most effectively in a situation of borderline iodine supply. The goitrogen thiocyanate potentiates the effect of severe iodine deficiency in endemic areas of Africa (53).

Several natural occurring goitrogens are listed in Table 17-2. Note that excessive

Nutritional use of seaweed (rich in iodine) may induce goiter. Moreover malnutrition (protein-caloric malnutrition) iron deficiency, selenium deficiency when associated with marginally low nutritional iodine may impair thyroid hormone synthesis and induce thyroid enlargement.

Table 17-2. Natural goitrogens associated with Multinodular Goiter

Goitrogens Agent Action
Millet, soy beans Flavonoids Impairs thyroperoxidase
Cassava, sweet potato, sorghum Cyanogenic glucosides metabolized to thiocyanates Inhibits iodine thyroidal uptake
Babassu coconut Flavoniods Inhibits thyroperoxidase
Cruciferous vegetables: Cabbage, cauliflower, Broccoli, turnips Glucosinolates Impairs iodine thyroidal uptake
Seaweed (kelp) Iodine excess Inhibits release of thyroidal Hormones
     
Malnutrition, Iron deficiency

Vitamin A deficiency

Iron deficiency

Increases TSH stimulation

Reduces heme-dependent thyroperoxidase thyroidal activity

Selenium Selenium deficiency Accumulates peroxidase and cause deiodinase deficiency; impairs thyroid hormone synthesis

 

Modified and adapted from Medeiros-Neto & Knobel, ref. 53

Inherited defects in thyroid hormone synthesis and resistance to thyroid hormone action

Inherited goiter and congenital hypothyroidism were first described by Stanbury and associated (30) in two goitrous siblings with defective thyroperoxidase action resulting in impaired iodine organification. Both siblings were mentally retarded and had enormous multinodular goiters. In the next fifty years a number of genetic defects in every step of thyroid hormone synthesis have been described in detail. If not diagnosed at birth the impaired thyroid hormone synthesis would result in an elevated TSH secretion and diffuse goiter could progressively appears. Other factors might be of importance regarding goiter formation. The level of nutritional iodine seems to be quite important in patients with the defective sodium iodine symporter (NIS), thyroglobulin gene mutations and the defective dehalogenase system (DEHAL gene). If a relatively high intake of iodine is provided goiter formation may be slowed down to a certain extent. On the contrary in marginally low nutritional iodine intake goiter will progress to a very large size and nodules will appear (multinodular goiter). It has been proposed that mutations of certain genes involved in thyroid hormone synthesis that do not entirely affect the physiological action of the translated protein may cause goiter later on life and more frequently in women (55). Thus the variable phenotype resulting from genetically documented mutations may be quite variable depending on environmental factors (iodine). Individual adaptation to the defective protein, rapid hydrolysis of defective TG, serum level of TSH and response of the thyroid epithelial cells to the growth-promoting effect of TSH are other factors to be considered.

It is conceivable that multinodular goiter could result from a defect in any step of thyroid hormone synthesis, and to resistance to thyroid hormone action. In both groups of defects in the thyroid hormone system serum TSH would be elevated and goiter would be the logical consequence of a prolonged stimulation to growth. In the context of other factor that might induce multinodular goiters the defective thyroid hormone system and resistance to thyroid hormone action are relatively rare conditions as compared to other factors.

etx-lipidpeds-ch17-table3

IOD Iodine organification defect; PIOD: Partial IOD;TIOD: Total IOD; RAI: radioactive iodine. Source: modified from “Genetic causes of dyshormonogenesis. Grasberger and Refetoff, 2011.

Other Thyroid-Stimulating Factors

Other substances that could be involved in stimulating thyroid enlargement are epidermal growth factor (EGF) and insulin-like growth factors (IGF). EGF stimulates the proliferation of thyrocytes from sheep, dogs, pigs, calves, and humans (42-51). While stimulating growth, EGF reduces trapping and organification of iodide, TSH receptor binding, and release of thyroglobulin, T3 and T4. On the other hand TSH may modulate EGF binding, to thyroid cell membranes and thyroid hormone may stimulate EGF production and EGF receptor number. In a study on adenomatous tissue, obtained from patients with multinodular goiter, it was found, by immunohistochemistry, that expression of EGF was increased (43). IGF-2 interacts with trophic hormones to stimulate cell proliferation and differentiation in a variety of cell types. The interaction between TSH and IGF-2 is synergestic (44). Increased IGF-I expression may contribute to goiter formation. A similar synergistic effect may exist between IGF-I and TSH. This synergism on DNA synthesis is mediated by complex interactions including the secretion of one or more autocrine amplification factors. Non-functioning nodules in patients with multinodular goiter contain the same IGF-1 receptors that are present in the normal adjacent extra-nodular follicles but are expressed in higher concentrations. Fibroblast growth factor (FGF)-1, stimulates colloid accumulation in thyroids of rat s but only in the presence of TSH (43). Expression of FGF-1 and -2 and FGF-receptor-1 will be followed by thyroid hyperplasia and may play a role in development of multinodular goiter (49). Fancia et al (50) found that in goiters with aneuploid components growth rate was higher than when euploid components were present (51). Other factors promoting cell growth and differentiation have been identified in the past. These include cytokines, acetylcholine, norepinephrine, prostaglandins, substances of neural origin like vasoactive intestinal peptide, and substances of C-cell origin. It is however not known to what extent these compounds play a role in the genesis of multinodular goiter.

The hypothesis that the development of thyroid autonomy is due to a gradual increase in the numbers of cells having relatively autonomous thyroid hormone synthesis is supported by the 27% prevalence of impaired TSH responses to TRH in patients with nodular goiter as opposed to such responses in only 1 of 15 patients with diffuse goiter (11). Such partial autonomy may appear only with time and could possibly be prevented by TSH-suppressive therapy. The fact that it is possible to induce hyperthyroidism in some patients with multinodular goiters by administration of iodide suggests that certain of the nodules in the multinodular gland are autonomous but unable under normal iodine intake to concentrate sufficient quantities of iodide to cause hyperthyroidism (53). Presumably iodide administration provides sufficient substrate for generation of excessive amounts of hormone, although it does not readily account for the long persistence of the hyperthyroidism in some of those cases.

Thus, there may be several etiologic factors in simple and nodular goiter, and some of these factors may act synergistically. The end result is a collection of heterogeneously functioning thyroid follicles, some of which may be autonomous and produce sufficient amounts of thyroid hormone to cause hyperthyroidism.

PATHOLOGY

Although it is rare to obtain pathological examination of thyroid glands in the early phase of development of multinodular goiters, such glands should show areas of hyperplasia with considerable variation in follicle size. The more typical specimen coming to pathologists is the goiter that has developed a nodular consistency. Such goiters characteristically present a variegated appearance, with the normal homogeneous parenchymal structure deformed by the presence of nodules     (Figure 17-6). The nodules may vary considerably in size (from a few millimeters to several centimeters); in outline (from sharp encapsulation in adenomas to poorly defined margination for ordinary nodules); and in architecture (from the solid follicular adenomas to the gelatinous, colloid-rich nodules or degenerative cystic structures). The graphic term “Puddingstone goiter” has been applied. Frequently the nodules have degenerated and a cyst has formed, with evidence of old or recent hemorrhage, and the cyst wall may have become calcified. Often there is extensive fibrosis, and calcium may also be deposited in these septae. Scattered between the nodules are areas of normal thyroid tissue, and often-focal areas of lymphocytic infiltration. Radioautography shows a variegated appearance, with RAI localized sometimes in the adenomas and sometimes in the paranodular tissue. Occasionally, most of the radioactivity is confined to a few nodules that seem to dominate the metabolic activity of the gland.

Figure 17-6.(A) Cross section of multinodular goiter. (B) Cross radioautograph of The thyroid in part a. Observe the variation in 131I uptake indifferent areas.

Figure 17-6. (A) Cross section of multinodular goiter. (B) Cross radioautograph of
The thyroid in part a. Observe the variation in 131I uptake indifferent areas.

If careful sections are made of numerous areas, 4-17% of these glands removed at surgery will be found to harbor microscopic papillary carcinoma (56-60). The variable incidence can most likely be attributed to the different criteria used by the pathologists and the basis of selection of the patients for operation by their physicians. These factors are discussed below.

NATURAL HISTORY OF THE DISEASE

Multinodular goiter is probably a lifelong condition that has its inception in adolescence or at puberty. Minimal diffuse enlargement of the thyroid gland is found in many teenage boys and girls, and is almost a physiologic response to the complex structural and hormonal changes occurring at this time. It usually regresses, but occasionally (much more commonly in girls) it persists and undergoes further growth during pregnancy. This course of events has not been documented as well as might be desired in sporadic nodular goiter, but it is the usual evolution in areas where mild endemic goiter is found.

Patients with multinodular goiter seek medical attention for many reasons. Perhaps most commonly they consult a physician because a lump has been discovered in the neck, or because a growth spurt has been observed in a goiter known to be present for a long time. Sometimes the increase in the size of the goiter will cause pressure symptoms, such as difficulty in swallowing, cough, respiratory distress, or the feeling of a lump in the throat. Rarely, an area of particularly asymmetrical enlargement may impinge upon or stretch the recurrent laryngeal nerve. Commonly the goiter is discovered by a physician in the course of an examination for some other condition. An important scenario is for the patient to seek medical attention because of cardiac irregularities or congestive heart failure, which proves to be the result of slowly developing thyrotoxicosis. (The issue is discussed more fully later in this chapter). Many times the goiter grows gradually for a period of a few too many years, and then becomes stable with little tendency for further growth. It is rare for any noteworthy spontaneous reduction in the size of the thyroid gland to occur, but patients often describe fluctuation in the size of the goiters and the symptoms they give. These are usually subjective occurrences, and more often than not the physician is unable to corroborate the changes that the patient describes. On the other hand, it could be that changes in blood flow through the enlarged gland account for the symptoms.

Occasionally, a sudden increase in the size of the gland is associated with sharp pain and tenderness in one area. This event suggests hemorrhage into a nodular cyst of the goiter, which can be confirmed by ultrasound. Within 3-4 days the symptoms subside, and within 2-3 weeks the gland may revert to its previous dimensions. In such a situation, acute thyrotoxicosis may develop and subside spontaneously.

Rarely, if ever, do the patients become hypothyroid and if they do, the diagnosis is more probably Hashimoto´s thyroiditis than nodular goiter. In a study in clinically euthyroid subjects with multinodular goiter, 13 out of 22 had subnormal TSH release after TRH. (12) If the goiter is present for long time, thyrotoxicosis develops in a large number of patients. In a series collected many years ago at the Mayo Clinic, 60% of patients with MNG over 60 were thyrotoxic. The average duration of the goiter before the onset of thyrotoxicosis was 17 years; the longer the goiter had been present the greater was the tendency for thyrotoxicosis to develop. This condition appears to occur because with the passage of time, autonomous function of the nodules develops. In a study of patients with euthyroid multinodular goiter, thyroid function was autonomous in 64 and normal in 26. After a mean follow-up of 5.0 years (maximum 12 years) 18 patients with autonomous thyroid function became overtly hyperthyroid and in 6 patients with primarily normal thyroid function autonomy developed (25-26). Thyroid function tests is illustrated in a patient with multinodular goiter starting from complete euthyroidism on to overt thyrotoxicosis. Occasionally a single discrete nodule in the thyroid gland becomes sufficiently active to cause thyrotoxicosis and to suppress the activity of the rest of the gland. (see Chap13). If these patients are given thyroid hormone, continued function of nodules can be demonstrated by radioiodine scanning techniques. Thus, these nodules have become independent of pituitary control. When patients with euthyroid multinodular goiter are frequently tested, it appears that in some of them occasional transient increases of serum T3 and / or T4 are seen. The possibility that the abrupt development of hyperthyroidism may follow administration of large amounts of iodine to these patients was reviewed by Stanbury and collaboration (61). In several areas of the world previously iodine deficiency the introduction of iodine supplementation lead to an increase of hyperthyroidism (non-autoimmune) possibly by excessive thyroid hormone production by “hot” thyroid nodules.

MULTINODULAR GOITER AND CANCER

If surgical specimens of multinodular goiters are examined carefully, 4-17% are found to harbor a carcinoma (56-60, 62-64). The use of ultrasound-guided fine needle aspiration (FNA) for evaluating these patients is not clearly defined. The biopsy of all the numerous nodules is impractical. Recently, a retrospective study with 134 patients showed a significant incidence (46,3%) of thyroid cancer in patients with multinodular goiter and benign FNA (65).These carcinomas vary widely in size and are typically of the papillary variety. Similar tumors are occasionally found in thyroid glands affected by Hashimoto´s thyroiditis and in otherwise normal glands. Bisi et al (59) reported that 13% of the glands resected in thyroid operations for any reason contained papillary adenocarcinoma. In Japan, routine autopsies of patients who were not suspected of having thyroid disease and who had no known irradiation experience, 17% were found to have small carcinomas when careful serial sections of the thyroid glands were done (62). If the figures of Bisi et al (59) were confirmed (63, 64) truly represent the prevalence of invasive carcinoma, one would certainly be forced to conclude that all multinodular goiters should be resected in order to prevent dissemination of malignant disease. However, it seems quite unlikely that all lesions that appear to satisfy the histological criteria for malignant neoplasia are potentially lethal. This view is strongly supported by the final report of the study on the significance of nodular goiter carried out in Framingham (see ref. 24). They followed for 15 years all 218 nontoxic thyroid nodules previously detected in a total population of approximately 5,000 persons. None of these lesions showed any clinical evidence of malignancy at the end of that time. Despite of the low-quality, the evidence suggests a lower prevalence of thyroid caner in multinodular goiter compared to single nodules, particularly in iodine-deficient areas (66, 67).

A strong case can be made for the view that there is only minimal risk from carcinoma in multinodular goiter. The prevalence of clinical nodularity of the thyroid is at least 4%, or 40,000 per 1,000,000 populations. Use of a much higher figure can be justified by the autopsy studies described above. Despite the high frequency of nodular goiter, only 36-60 thyroid tumors appear per 1,000,000 persons each year or by analysis of reported statistics on thyroid surgical specimens (57-60). A recent national cancer survey in the United States found an incidence of 40 per 1,000,000. An overview of the incidence of thyroid cancer in 409 countries, both with and free of endemic goiter was reported previously (58). The range of incidence varied between 7.5 and 56 per 1,000,000 persons each year. The prevalence of significant thyroid carcinoma at routine autopsy is less than 0.1% and persons with this type of tumor are probably examined as frequently as are those with other forms of neoplasia. The United States mortality figures for thyroid carcinoma are constant at about 6 per 10-6 population each year. Riccabona also summarized death rates from thyroid cancer in non-endemic and in endemic countries. (64) For Austria this was 16 per 10-6 per year in 1952 and 10 per 10-6 per year in 1983. For Switzerland this was in 1952, 18 per 10-6 per year and in 1979, 9 per 10-6 per year. The death rate per year for the United States in 1979 was 3 per 10-6, for Israel in 1952 1 per 10-6 per year and for the UK 7 per 10-6 in 1963. Death rates from thyroid cancer in endemic goiter areas from regions in Austria, Yugoslavia, Finland and Israel were between 10 and 16 per 10-6 per year between 1980 and 1984.

Lastly, it should be recognized that meticulous examination of autopsy specimens from persons dying of nonthyroid disease may show small (less than 0.5 cm) papillary lesions in4-24% of human thyroid glands (63,64). A report of 1020 sequential autopsies revealed the presence of microscopic papillary carcinoma in 6%. (60) Although the prevalence of this type of lesion increases with age, there is no question that such lesions may be present even in younger persons. The proportion of these lesions that even become clinically apparent is unknown, but their presence in otherwise normal thyroid glands should be kept in mind when evaluating reports of similar prevalences of thyroid carcinoma in multinodular thyroid glands.

If 4% of patients with nodular goiter actually have thyroid carcinoma, the prevalence of tumor in the general population would be 1,600 per 1,000,000. It is remarkable that only about 25 of these 1,600 hypothetical tumors would become apparent each year, or that only about 10 would prove fatal. Thus, there appears to be a gross discrepancy between the mortality form thyroid carcinoma and its reported frequency in surgical specimens of multinodular goiters. Reasonable arguments can be mustered in an effort to reconcile the information. Perhaps the most important single factor is selection. Persons with nodular goiter who come to operation are not representative of the general population but are patients with clinically significant thyroid disease who have been selected by their physicians for thyroid surgery. One of the factors controlling the selection process is the suspicion of malignant tumor. In fact, the selection process is especially good, as reflected by the high recovery of malignant thyroid tumors in patients operated on with this presumptive diagnosis. A second factor is that the histologic diagnosis of thyroid carcinoma may not correlate well with true invasiveness. It is impossible to prove this thesis, but pathologists agree that the criteria for judging malignancy are variable and that it is exceedingly difficult to predict with any degree of certainty the growth potential of a particular thyroid lesion.

Other arguments may be used to defend a conservative therapeutic position. In the first place, the tumors that are usually found in multinodular goiters are papillary tumors, and their degree of invasiveness is low. Indeed, the survival rate for intrathyroid papillary carcinoma is only slightly less than that for normal persons of the same age and sex (69-74). Furthermore, prophylactic subtotal thyroidectomy is not a guarantee of protection from cancer arising in a nodular goiter, since the process is usually diffuse, and it may be assumed that abnormal tissue is left in the neck after operation. In fact, unless replacement therapy is given, partial thyroidectomy might be expected to induce a tremendous growth stimulus in the remaining gland (75-80). A further point is that thyroidectomy, even in the best of hands, carries its own risk and its own morbidity, with dimensions comparable to those of missing a small papillary carcinoma within a multinodular goiter (81-84). Obviously this last possibility does not apply when a focus of unusual induration or rapid growth rate is detected clinically.

Diagnosis

Many of the symptoms of multinodular goiter have already been described. They are chiefly due to the presence of an enlarging mass in the neck and its impingement upon the adjacent structures. There may be dysphagia, cough, and hoarseness. Paralysis of recurrent laryngeal nerve may occur when the nerve is stretched taut across the surface of an expanding goiter, but this event is very unusual. When unilateral vocal cord paralysis is demonstrated, the presumptive diagnosis is cancer. Pressure on the superior sympathetic ganglions and nerves may produce a Horner´s syndrome.

As the gland grows it characteristically enlarges the neck, but frequently the growth occurs in a downward direction, producing a substernal goiter. A history sometimes given by an older patient that a goiter once present in the neck has disappeared may mean that it has fallen down into the upper mediastinum, where its upper limits can be felt by careful deep palpation. Hemorrhage into this goiter can produce acute tracheal obstruction. Sometime substernal goiters are attached only by a fibrous band to the goiter in the neck and extend downward to the arch of the aorta. They have even been observed as deep in the mediastinum as the diaphragm. Occasionally the skilled physician can detect a substernal goiter by percussion, particularly if there is a hint from tracheal deviation, or the presence of a nodular mass in the neck above the manubrial notch.

Symptoms suggesting constriction of the trachea are frequent, and displacement of the trachea is commonly found on physical examination. Computer Tomography examination is useful in defining the extent of tracheal deviation and compression. Compression is frequently seen but rarely is functionally significant have expected to find softened tracheal cartilage after the removal of some large goiters, but tracheomalacia has been observed only on the rarest occasion. Patients may be remarkably tolerant of nodular goiter even when the enlargement is striking. This finding is especially true in the endemic goiter areas of the world.

It is generally agreed that, thyroid isotope or ultrasound scanning are of little or no use in the diagnosis of carcinoma in a multinodular goiter. Two aspects are important in the differentiation from malignancy. First, the clinical presentation, if the goiter is of longstanding, showing little or no growth, absence of a dominant node, familial, while there is no neck irradiation in the past, especially in childhood, no hoarse voice, and no suspicious lymphnodes in the neck, there is little fear for carcinoma.

Table 17-4      Clinical symptoms and investigations in the diagnosis of MNG

Simptoms and signs

Often family history of benign thyroid disease

Slowly growing anterior neck mass

Uni- or multinodularity on examination

Enlargment during pregnancy

Cosmetic complaints

Asymmetry, tracheal deviation, and/or compression

Rarelly upper airway obstruction, dyspnea, cough, and dysphagia

Sudden transient pain or enlargement secondary to hemorrhage

Gradually developing hyperthyroidism

Superior vena cava obstruction syndrome (rare)

Recurrent nerve palsy (rare)

Horner´s syndrome (rare)

Investigations

TSH normal or decreased, normal free T4, and free T3,

Serum Tg usually elevated

Thyroid autoantibodies (TPO and Tg) usually negative

Scintigraphy with solitary or multiple hot and/or cold areas

Ultrasound finding of solitary or multiple nodules with varying

echogenicity (nonhomogeneity)

Computed tomography and MR imaging demonstrating solitary or

multiple nodules with varying echogenicity

Lung function testing may demonstrate impaired inspiratory capacity

Fine-needle aspiration of solitary or dominant nodules – benign cytology

 

Modified and adapted from Hegedus et al (24)

Laboratory investigation

The choice of tests to investigate the functional status of a patient with a Simple diffuse goiter or Multinodular goiter may differ depending on the geographic areas of the world. Recent surveys conducted in the American, European and Latin American Thyroid Associations have indicated that the North American thyroidologists are quite restrictive in the choice of laboratory tests. Most of the experts, however, would perform a serum TSH and serum Free T4 test. In other settings Total T4 and Total T3 are also included because of the preferential secretion of T3 over T4 in mild iodine deficiency (53).

Antibodies against thyro-peroxidase (anti-TPO) and thyroglobulin (anti-TG) are measured, routinely, by most Europeans and Latin Americans thyroidologists. This seems to be relevant because thyroid auto antibodies are found approximately in 10% of the population and, consequently, autoimmunity may coexist with a goiter. Also diffuse or focal lymphocytic infiltration in an enlarged gland may represent chronic autoimmune thyroiditis.

Although serum TG correlates with the iodine status and the size of the enlarged thyroid gland it has little or no value in the diagnosis of goiter.

Diagnostic imaging

Neck palpation is notoriously imprecise with regard to thyroid morphology and size estimation (85). Several imaging methods are available in most settings: scintilography (with radioiodine, technetium), ultrasonography, computed tomography scans, magnetic resonance imaging and, less frequently used, positron emission tomography (PET). In Table 17-5 it is listed the characteristics, advantages and disadvantages of these imaging methods.

Ultrasonography of the thyroid

The main reasons for the widespread use of thyroid sonography are availability (several portable models are widely available at a relatively affordable price), the low cost of the procedure (if performed in the office or in the thyroid clinic), limited discomfort for the patient, and the non ionizing nature of the method. Ultrasonography may detect non palpable nodules cysts, will estimate nodule and goiter size (volume), will monitor the changes following therapy and will guide the Fine Needle Aspiration Biopsy (FNAB). After the introduction of ultrasonography it has become clear that nodules in the thyroid gland are very prevalent, ranging from 17% to 60% if older people are included in the study (85-95).

Hypoechogenicity, micro-calcifications, indistinct borders, increased nodular flow (visualized by DOPPLER) may have predictive value in distinguishing malignant from benign nodules (even in Multinodular Goiters).

The possibility of measuring thyroid volume is another highly useful feature of ultrasonographic studies particularly after therapy with L-T4 or radioiodine ablation. The volume of the goiter is usually based on the ellipsoid method (length, width depth X pi/6). This has an observer coefficient of variation of more than 10%. When compared to CT planimetry the ellipsoid method underestimate the goiter volume by 20%. Ultrasonography can not evaluate a multinodular goiter that has partially migrated to the upper mediastinum.

Ultrasound elastography can also provide information regarding malignant risk of thyroid nodules and multinodular goiter, however with questionable sensitivity (75%) and specificity (45,73%) (96).

Scintigraphy (isotope imaging)

It was used routinely in the past but at present has little place in the evaluation of a multinodular goiter (97-101). It is helpful in the determination of the functionality of the various nodules of a MNG. Thyroid scintigrams have been used through many years for measurement of the thyroid volume but compared to other methods is very inaccurate (24).

Computed tomography (CT) and Magnetic resonance (MR)

CT and MR provide high-resolution visualization of the goiter (Simple diffuse, multinodular). The major strength of CT and MR is their ability to diagnose and assess the extent of subesternal goiters (Fig. 17-7). Another advantage of the CT is the possibility for planimetric volume estimations, quite useful in irregularly enlarged multinodular goiter (102-105).

Recently the ionizing radiation delivered by a CT procedure has been source of concern for both clinicians and radiologists. Therefore the use of CT as an imaging method should be reserved for intra thoracic multinodular goiters, with tracheal compression.

 Table 17-5      Characteristics of imaging procedures in relation to nodular thyroid disease

 

  Advantages Disadvantages
Sonography

·  High Availability

·  High morphologic resolution

·  No ionizing irradiation

·  Dynamic picture

·  Blood flow visualization (Doppler)

·  Biopsy guidance, also of lymph nodes

·  Moderate precision in volume estimation

·  Operator dependency

·  No information of functionality

·  Not feasible in substernal goiter

·  Poor prediction of malignancy

Scintigraphy

·  Information of functionality

·  Differentiates between destructive and hyperthyroid conditions

·  Measurement of thyroid iodine uptake

·  Predictive of feasibility of ¹³¹I therapy

·  Detects ectopic thyroid tissue

·  Requires nuclear medicine

·  Ionizing irradiation

·  Poor resolution

·  Poor differentiation between solid and cystic cold nodules

·  Volume estimationinaccurate

CT Scan

·  High morphologic resolution

·  Visualization of adjacent structure

·  Ideal for substernal goiter

·  Planimetric volume estimation

·  Volume estimation probably accurate

·  Ionizing irradiation

·  No information of functionality

·  Poor prediction of malignancy

MR imaging

·  No ionizing irradiation

·  High morphologic resolution

·  Visualization of adjacent structure

·  Ideal for substernal goiter

·  Planimetric volume estimation

·  Volume estimation with high precision

·  Moderate availability

·  Long procedure time

·  Not usable with metallic objects inside patient

·  No information of functionality

·  Poor prediction of malignancy

PET

·  Information of functionality

·  Metabolic investigations

·  Good prediction of malignancy

·  Low availability and high cost

·  Requires specialized units

·  Ionizing irradiation

 

CT, Computed tomography. MR, magnetic resonance

 

Modified and adapted from Hegedus et al (24)

Treatment of multinodular goiter

In the past iodine supplementation seems to be an adequate approach because goiter development is associated with mild iodine deficiency in many countries worldwide. The effect of iodine once a multinodular goiter has developed a limited value in reducting the MNG. A major problem of iodine supplementation is the risk for inducing subclinical / clinical hyperthyroidism (Jod-Basedow). Therefore aside from a few European Countries iodine is no longer used alone or associated with L-T4 to treat thyroid enlargement (24).

This leaves in essence three modalities of therapy:

(1). L-T4 suppressive therapy

(2). Radioiodine (¹³¹I) alone or preceded by rhTSH

(3). Surgery

L-T4 suppressive therapy

L-T4 suppressive therapy is used extensively both in Europe, USA and Latin America, according to their respective surveys. A beneficial effect of L-T4 has been demonstrated in diffuse goiters in many controlled trials (106-112). A goiter reduction of 20-40% can be expected in 3-6 months of therapy, the goiter returning to the pre-treatment size after L-T4 withdrawal. The efficacy of L-T4 is shown to depend on the degree of TSH suppression. When it comes to the nontoxic MNG there are five controlled studies in which sonography was used for objective size monitoring. Berghout et al (113) in a randomized double-blind trial showed that the goiter volume was reduced by 15% (9 months of L-T4 therapy). In the placebo group the goiter continued to increase in size by more than 20% in the 9 months period. The goiter volume returned to baseline values after discontinuation of the therapy. Lima et al (109) studied 62 patients with nodular goiter. Thirty per cent of patients were regarded as responders (reduction > 50% of the initial volume). In the control group 87% showed no change or an increase in goiter size. Wesche et al (110) compared L-T4 with ¹³¹I therapy in a randomized trial. The median reduction of goiter volume in the radioiodine treated group was 38-44% whereas only 7% of the L-T4 treated patients had a significant goiter reduction.

Papini et al (111) treated 83 goitrous patients (nodular goiter) with suppressive doses of L-T4 comparing the results with a control group. The L-T4 therapy was extended for 5 years. There was a decrease in nodular size in the L-T4 treated group and a mean volume increase in the control group. After 5 years sonograms detected 28.5% new nodules in the control group but only 7.5% in the L-T4 treated group. In conclusion long term TSH suppression induced volume reduction in a subgroup of thyroid nodules but effectively prevented the appearance of new nodules.

Zelmanovitz et al (112) studied 42 women with a single colloid nodule. Twenty one patients were treated with 2.7µg/kg of L-T4 for one year. Six of the 21 treated patients had a >50% reduction of the nodule volume as evaluated by sonography as compared to only 2 (out of 24 patients) that received placebo. They concluded that L-T4 therapy is associated with 17% of reduction of a single colloid nodule and may inhibit growth in other patients. They also conducted a meta-analysis of 6 prospective controlled trials and concluded that four of seven studies favors treatment with L-T4. The treatment of single nodules or multinodular goiter with L-T4 is an open issue as the reduction of the nodule / MNG is only obtained in about one third of patients. The possible unwanted effects of L-T4 therapy have also to be considered (114, 115).

Table 17-6: Controlled studies of L-T4 therapy in multinodular goiter using a

precise thyroid size determination

Authors (Country) (n)

Duration of

L-T4 therapy

Dose of L-T4 Outcome of continuous L-T4 Therapy vs. Controls

Berghout et al

(The Netherlands)

55 9 months 2.5μg/kg 25% reduction among responders*

20% had

Increase of nodular volume

Lima et al (Brazil) 62 12 months 200μg/dia 30% reduction** No variation volume

Wesche et al

(The Netherlands)

57 24 months 2.5μg/kg 22% reduction 44% volume with Radioidine
Papini et al (Italy) 83 5 years

2,0μg/kg

7.5% new nodules

47.6% reduction

28.5% new

nodules

22% had reduction nodules

Zelmanovitz et al

(Brazil)

45 12 months 2.7μg/kg 28% reduction** 8.3% had reduction

(*) Effective response to L-T4 therapy: volume was reduced by 13% of basal

(**) Effective response to L-T4 therapy: volume reduction >50% of basal

Radioiodine ablation of goiter

General considerations: It has long been recognized that radioiodine administration results in shrinkage of the goitrous thyroid gland. Over 20 years ago ¹³¹I therapy reduced the MNG volume by approximately 40% in the first year, and 50-60% in the second year. In very large goiters with volume over 100 mL the reduction is less (around 35%). Patient with substernal MNG have also been treated with beneficial results. The individual response to radioiodine therapy, regarding goiter reduction and development of hypothyroidism is very difficult to predict. Goiter reduction is related to the absorbed thyroid dose. In most centers ¹³¹I doses of 3.7 MBq/g of thyroid tissue corrected for 100% 24h radioiodine uptake have been given. In other centers a fixed doses of radioiodine (100mCi, 150mCi) are administered according to the thyroid volume. The risk of permanent hypothyroidism after ¹³¹I therapy in MNG ranges from 11 to 58% after 1 to 8 years of follow-up (116-124).

 

The use of rhTSH for improving ¹³¹I therapy of nontoxic multinodular goiter

(1). Increased uptake and goiter volume reduction

In recent years, pretreatment with rhTSH has been used in patients with MNG (which typically have only a fraction of the normal RAIU) to increase ¹³¹I uptake in the goiter and allow treatment with lower doses of ¹³¹I to induce thyroid volume reduction (125-129). Accordingly, in a study of 15 patients with nontoxic MNG, pretreatment with a single low dose of rhTSH (0.01 or 0.03 mg 24 h before ¹³¹I administration) resulted in a doubling of RAIU (130). In addition, the single dose of rhTSH caused a more homogeneous distribution of ¹³¹I by stimulating more uptake in relatively cold areas than in hot areas, particularly in patients with low serum TSH levels (Figure 17- 7).

Various studies have demonstrated the effect of rhTSH on ¹³¹I therapy for MNG. Twenty-two patients with MNG were treated with ¹³¹I 24h after administration of 0.01 or 0.03 mg rhTSH (131). In this study, the dose of ¹³¹I was adjusted to the increase in uptake induced by rhTSH, aimed at 100 µCi/g thyroid tissue retained at 24h. Pretreatment with 0.01 and 0.03 mg rhTSH resulted in reductions in the ¹³¹I dose by a factor of 1.9 and 2.4, respectively. One year after treatment, there was a reduction in thyroid volume of 35% and 41% in the two groups, respectively. Despite delivering a good therapeutic response, the administration of ¹³¹I 100 µCi/g of thyroid tissue corrected for 24-h RAIU raises concerns of irradiation of the surrounding neck structures and potential risk for stomach, bladder, and breast cancer, which have been reported after ¹³¹I therapy for toxic nodular goiter (24). In another study (132), 16 patients with MNG were treated with a fixed dose of ¹³¹I (30 mCi) 72h after pretreatment with 0.3 mg rhTSH, or 24h after pretreatment with 0.9 mg rhTSH. The two regimens were equally effective, leading to a 30 to 40% reduction in thyroidal volume at 3 to 7 months. Giusti et al compared the 12-months outcome after RAI and rhTSH arbitrarily chosen (0.1mg for 24-h RAIU > 30 %; 0.2 mg for RAIU<30 %) between patients with basal non-toxic (TSH>0.3 mIU/l)) and non autimmune pre-toxic MNG (TSH<0.3 mIU/l). They confirmed the effectiveness of rhTSH adjuvant treatment in reducing thyroid volume after low RAI dose (<600 MBq) independently of the baseline TSH level. A more severe thyrotoxic phase after rhTSH was observed in patients with TSH<0.3 mIU/L, while L-T4 therapy was more frequently needed when initial TSH levels were > 0.3 mIU/l (133)

As mentioned, rhTSH was administered 24h before ¹³¹I therapy in most studies. However, results from a study published by Duick and Baskin (134, 135) suggested that the time interval may be even longer to achieve a maximum stimulation of the thyroid RAIU.

Recently in a phase II, single blinded, placebo-controlled study with 95 patients evaluating two low doses (0.01 and 0.03mg) of modified-release rhTSH, no statistical significant enhancement of thyroid volume reduction was achieved at three years follow-up (41% to 53%) . The modified-release rhTSH was developed to minimize the side effects related to thyroid hormone excess, (136).

(2). Tracheal compression and pulmonary function

Many elderly patients have significant intrathoracic extension of the MNG, which may cause tracheal compression with subsequent airflow reduction. Bonnema et al (137) evaluated upper airway obstruction by flow volume loops in 23 patients with very large goiter. In one third of the patients, there was impairment of the forced inspiratory flow at 50% of the vital capacity (FIF50%).The authors found a significant correlation between FIF50% and the smallest tracheal cross-sectional area. Reduction of the MNG volume after high dose of ¹³¹I had a remarkable effect in enlarging tracheal cross-sectional area and consequently improving inspiratory capacity in these patients.

(3). Transient hyperthyroidism after ¹³¹I ablation

Other studies using different doses of rhTSH and showing comparable RAIU increases with lower doses, demonstrated significant goiter reduction, but also transient hyperthyroidism after ¹³¹I therapy (131-144). A study in which 34 patients with large MNGs were randomized to receive ¹³¹I therapy (100 µCi/g of thyroid tissue) alone or following a single relatively high dose of rhTSH (0,45 mg) 24h before ¹³¹I administration, showed that patients who received rhTSH had transient elevations in thyroid hormone levels lasting a few weeks, a greater reduction in goiter size (60% vs. 40%), and a higher incidence of hypothyroidism (65% vs. 21%) (142). In another study, 18 patients received two 0.1 mg doses of rhTSH followed by 30 mCi of ¹³¹I. RAIU increased from 12 to 55%, free T 4 increased from 1.3 to 3.2 ng/dL, and goiter size reduced from 97 to 65 mL. However, about 30% of the patients experienced painful thyroiditis and 39% had mild hyperthyroidism (137). In a randomized trial of ¹³¹I treatment   calculated to deliver a thyroidal absorbed dose of 100 Gy (10 mrads) and administered 24h after rhTSH (0.3 mg) or placebo, patients with MNG (mean goiter volume of 55 cm³) who received rhTSH had more symptoms of hyperthyroidism and neck pain during the first week after treatment, a greater reduction in goiter size (52% vs. 46%), and a higher frequency of hypothyroidism (62% vs. 11%) (145). Using a similar study design, Bonnema et al (141) compared the effects of rhTSH (0.3 mg) or placebo, followed by a maximum dose of ¹³¹I 100 mCi on goiter volume reduction in 29 patients with very large goiters (median volume of 160 mL). After 12 months, the median goiter volume (monitored by magnetic resonance imaging) was reduced by 34% in the placebo group and by 53% in the rhTSH group. In the placebo group, the goiter reduction correlated positively with the retained thyroidal ¹³¹I dose, whereas this relationship was absent in the rhTSH group. Adverse effects, mainly related to thyroid pain and cervical compression, were more frequent in the rhTSH group. At 12 months, goiter-related complaints were significantly reduced in both groups without any between-group difference. One patient in the placebo group and three patients in the rhTSH group developed hypothyroidism.

Recently, an uncontrolled study (140) demonstrated the effect of rhTSH (0.1 mg, single dose) followed by ¹³¹I 30mCi 24h later in 17 patients with MNG (mean thyroid volume of 106 cm³). Pretreatment with rhTSH resulted in a mean RAIU increase from 18 to 50% and an increase in free T4 of 55% at 24h. Mean T3 levels increased by 86% and peaked at 48h, and median TG levels increased about 600% and peaked on the fifth day. Symptomatic tachycardia was promptly relieved with ß-blocker administration. After 12 months, mean thyroid volume measured by computed tomography had reduced by 46%. The adverse effects observed were transient hyperthyroidism (17.6%), painful thyroiditis (29.4%), and hypothyroidism (52.9%).

 

(4). Degree of goiter reduction, ¹³¹I dose, and rhTSH

Most investigators (Table 17-7) could not find any correlation of thyroid volume reduction with post-rhTSH RAIU, area under the curve of TSH, basal thyroid volume, or effective activity of ¹³¹I. Also, in the placebo-controlled study by Bonnema et al (141), no significant correlation was found, in either the placebo group of the rhTSH-treated group, between the degree of goiter reduction and the initial goiter size. However, in the placebo group, there was a correlation (r = 0.74) between the degree of goiter reduction and the retained ¹³¹I thyroid dose, an observation in agreement with previous reports (135). At variance, Albino et al (131) found a positive correlation (r = 0.68) between the degree of goiter volume reduction and the effective activity of administered post-rhTSH ¹³¹I dose. This issue, therefore, needs further clarification, but overall, these studies suggest that goiter reduction may be dependent on other factors caused by rhTSH pre-stimulation and not only on the applied ¹³¹I thyroid dose. For example, rhTSH could induce reactivation of inactive thyroid tissue or render the thyrocytes more vulnerable to ionizing radiation. Generally, the dose of ¹³¹I in these studies ranged from 75 to 400 µCi/g tissue, and most patients received doses between 100 and 200 µCi/g, similar to those used to treat hyperthyroidism.

Figure 17-7 – Goiter reduction volume (%) at last follow-up of patients treated only with radioiodine (left bars) as compared with patients that received recombinant human TSH plus radioiodine (right bars). Note the significant and early volume reduction with radioiodine preceded by rh TSH (146)

Figure 17-7 – Goiter reduction volume (%) at last follow-up of patients treated only with radioiodine (left bars) as compared with patients that received recombinant human TSH plus radioiodine (right bars). Note the significant and early volume reduction with radioiodine preceded by rh TSH (146)

Table 17-7. Studies on the effect of recombinant human TSH on goiter reduction  
in multinodular goiter patients.            

 

 

No. of subjects Dose of rhTSH (mg) Time interval between rhTSH and radioiodine (123I or 131I) Therapeutic dose of 131I (mCi) Peak increase in thyroid hormones (%) Goiter reduction(%) Time of follow-up Goiter size estimation (Methods)

Remarks

 

 
Nieuwlaat et al. (128) 12 0.01 24 h ~39 (mean) Free T4: 47       Free T3: 41 35 1 year MRI 0.01 mg: 131I activity reduced by a factor 1.9  
  10 0.03 24 h ~23 (mean) Free T4: 52       Free T3: 59 41 0.03 mg: 131I activity reduced by a factor 2.4;   Hypothyroidism: 36%  
Duick & Baskin (134) 6 0.3 72 h 30 NI NI 7 m Palpation 0.3 mg: increase in 4 h RAIU 72 h after rhTSH: from 3.9 to 17  
  10 0.9 24 h 30 NI 30-40     0.9 mg: remission of compressive symptoms in 69%                                           Hypothyroidism: 56%  
Silva et al.   (142) 17 none   ~96 (mean) Free T4: 34 T3: 33 40 1 year CT 131I:                                                   Hypothyroidism: 23%  
  17 0.45 24 h ~90 (mean) Free T4: 594         T3: 73 58     131I + rhTSH:                                     Hypothyroidism: 64%; hyperthyroidism: 100%  
Albino et al. (131) 18 2 x 0.1 24 h 30 Free T4: 146           T3: 191 39 6 m CT 24 h RAIU increased from 12 – 53%; Hypothyroidism: 65%; hyperthyroidism: 39%  
Giusti et al. (140) 8 none   NM NM 25 20 m US + CT    
  12 2x0.2 24 h NM Free T4: 290*        Free T3: 340* 44 22 m US + CT    
Cohen et al. (132) 17 0.03 24 h ~30 Free T4: 46 T3: 33 34 6 m CT 24 h RAIU increased from 26% to 43%; Hypothyroidism: 18%; hyperthyroidism: 18%  
Nielsen et al. (145) 29 none   14 (median) NM 46 1 year US 131I: 24 h RAIU decreased from 32 to 29; Hypothyroidism: 11%; hyperthyroidism: 21%  
  28 0.3 24 h ~16 (median) NM 62     131I + rhTSH: 24 h RAIU increased from 34 to 47; Hypothyroidism: 62%; hyperthyroidism: 36%  
Bonnema et al. (141) 15 none 24 h ~42 (median) NM 34 1 year MRI 131I: hypothyroidism: 7%  
  14 0.3   ~38 (median) NM 53     131I + rhTSH: hypothyroidism: 21%  
Paz-Filho et al. (137) 17 0.1 24 h 30 Free T4: 56 T3: 87 46 1 year CT 24 h RAIU increased from 18 to 50%; Hypothyroidism: 53%; hyperthyroidism: 18%  
Cubas et al. (147) 28

A: 0.1

B: 0.005

C: NONE

24 h

30

 

Free T4: 31

Free T4: 23

Free T4: 19

37.2

39.3

15.3

 

2 years

CT

43% had hypothyroid signs

25.9% had persistant hypothyroidism

 
Romão et al. (148)

Eu: 18

SCH: 18

CH: 6

0.1

24h

 

30

 

Free T4: 67

Free T4: 106

Free T4: 170

79.5

70.6

68.7

 

3 years

CT

Hypothyroidism: 50%

11%

16%

Side effects more commonly find in SCH and CH

 
Fast et al   Clontrol 24h   rh TSH>   1 year   see Figure 17-7  
(146)   Rh TSH     X control          
Fast et al. (136) 95

Placebo

0.01

0.03

24hs 100Gy NM

44

41

53

3 years CT Hypothyroidism: more frequently in 0.03 mg group  
Giusti et al (133) 26

TSH>0.3mlU/l

 

TSH<0.3mlU/l

24hs 600 MBq NM

67,1

 

 

61.7

55.3 ± 4.1m

 

57.2 ± 5.1m

Ultrasonography Several side effects in both groups  
                                 

M: months

(5). Increase in goiter size immediately after ablation

It is worth mentioning the possibility of increase in goiter size with rhTSH (142, 145). In a study of 10 patients with MNG who were given 0.3 mg of rhTSH, it was shown that 24h after rhTSH, the mean goiter volume increased by 9.8% and after 48h, by 24%, reverting to baseline at 1 week. This suggests that rhTSH may lead to significant cervical compression in patients with near obstructive goiters, when used for improving ¹³¹I therapy in patients with goiter (145). All side effects related to acute thyroid enlargement causing tenderness and dyspnea due to possible obstruction of tracheal airway were promptly resolved with corticosteroid therapy.

(6). Radioactive iodine and rhTSH in elderly with hyperthyroidism

Treatment with ¹³¹I following rhTSH stimulation is also an attractive alternative in elderly patients considered poor surgical candidates or who refuse surgery. The prevalence of MNG rises in the elderly, a population in whom comorbities prevail. Of even greater concern in iodine repleted areas is the development of subclinical or overt hyperthyroidism, since thyroid hyper-function may increase the mortality risk in these patients (148). An Italian study assessed 20 elderly patients with large goiters and compared treatment with ¹³¹I (10 to 15 mCi fixed dose) following two consecutive 0.2 mg doses of rhTSH (n = 12; 3 patients had subclinical hyperthyroidism with TSH <0.3 µU/ml) with treatment with ¹³¹I alone (n = 8; subclinical hyperthyroidism recorded in 5). Patients who received rhTSH had higher transient elevations in free T4 and Free T3 lasting 2 weeks, a greater reduction in goiter size (44% vs. 25%). Both groups had a 17% incidence of hypothyroidism ~ 2 years after ¹³¹I therapy. Symptomatic relief occurred in all but 1 patient following rhTSH with a 50% median reduction on thyroid volume after about 2 years (140). In study conducted by Silva et al (142), 17 elderly subjects with MNG treatment with ¹³¹I 24h after pretreatment with 0.45 mg rhTSH and were compared with 17 elderly controls treated with ¹³¹I alone. In patients pretreated with rhTSH, serum TSH and T3 levels rose to a peak level in 24h, returning to normal at 72h. Serum free T4 concentrations rose significantly at 48h returning to normal at 7 days. Serum TG increased and remained elevated during the following 12 months. Patients pretreated with rhTSH had a 58% reduction in goiter volume when compared with 40% in patients treated with ¹³¹I alone. Hypothyroidism was more frequent in pretreated patients (65% versus 21% in non-pretreated) after 1 year. No symptoms of hyperthyroidism were observed in these patients. Four years after ¹³¹I therapy, additional thyroid volume reduction was similar for patients treated with rhTSH prior to ¹³¹I or with ¹³¹I alone, but it was significantly more pronounced in the rhTSH group, mainly in the first year (149). Although no additional benefit of rhTSH was observed after a long follow-up, the initial difference in thyroid volume reduction was maintained, denoting the advantage of using rhTSH pretreatment to achieve higher thyroid volume reduction during the first treatment (Table 17-6).

In another report of a short-term observational study, the investigators assessed the efficacy of a low-dose (0.03 mg) rhTSH stimulation on a fixed therapeutic activity of ~ 30 mCi ¹³¹I in 17 patients with large nodular goiters (12 with overt or subclinical hyperthyroidism / TSH <0.5 µU/ml and five on treatment with thionamides) (147). RAIU increased from 26% to 43%, free T4 increased from 1.4 to 2.0 ng/dl, and goiter size decreased from 170 to 113 cm³ by 6 months. Symptomatic relief, improved well-being and / or reduction, or elimination of anti-hyperthyroid drug was seen in 76% of the patients. However, 3 (18%) patients presented transient neck pain or tenderness, 1 experienced asymptomatic thyroid enlargement, and 3 became hypothyroid by 3 months (Table 17-6). A recent paper (146) compared the results of RAI alone and RAI preceded by rhTSH (see Figure 17-8) clearly demonstrating the efficacy of pre-treatment with rh TSH.

(7). Cardiovascular events after RAI ablation

Cardiovascular parameters to detect transient elevation of serum thyroid hormones were evaluated in 27 of 42 patients (age range 42-80 years) with large MNGs who were treated with rhTSH before receiving ¹³¹I 30 mCi (150). All patients presented a transient surge in serum levels of free T4 and total T3 into the hyperthyroid range following therapy. However, post-treatment cardiovascular evaluation did not show significant changes when compared with baseline evaluation, suggesting that treatment of MNGs with RAI after rhTSH stimulation does not affect structural and functional parameters of the heart. These findings are reassuring, particularly when considering treatment for older adults with comorbidities that preclude surgery.

(8). Thyroid autoantibodies occurrence after ¹³¹I therapy

Some studies have reported the development of thyroid antibodies associated with ¹³¹I therapy (151), however a direct cause-effect linking to rhTSH has not been demonstrated. These observations have been interpreted as an immunological response caused by the release of thyroid antigens from destroyed follicular cells. In a study published by Rubio et al (152), it was found that rhTSH pretreatment had no significant effect in the development of antibodies (TSH receptor and TPO) when compared with treatment with ¹³¹I alone. As noted below, up to 5% of individuals develop auto-immune hyperthyroidism after 131-I therapy.

(9). Potential induction of malignancy

Although generally ignored, treatment with large doses of 131-I obviously raises the possibility of induction of malignancy. This has not so far been recorded in relation to therapy of MNG. Depending on functionality of the thyroid tissue, dose administered, size of the goiter, and size of the patient, whole body radiation could be up to 1 rad/mCi given, a dosage similar to that obtained during therapy of thyroid cancer. Perhaps the major use of this treatment will be in older individuals, with a shorter potential life span after treatment, which would presumably make this less of a concern.

Conclusions and comments

Given the limited experience published in the literature so far, before considering the routine use of rhTSH administration before ¹³¹I treatment of MNG, several issues must be taken into consideration (153-157).

  • ¹³³I treatment alone can lead to a 15-25% transient increase in thyroid volume during the first week after treatment;
  • rhTSH administration alone occasionally can lead to a significant increase, albeit transient, in thyroid volume, of up to 100% in normal subjects with 48h.
  • The combination of the two modalities may lead to a substantial acute increase in thyroid volume;
  • ¹³¹I treatment of MNG leads to transient hyperthyroidism during the first 2-3 weeks after therapy and the combination with rhTSH administration can enhance this effect, with potential consequences particularly for the elderly patients (148);
  • The optimal dose of rhTSH for pretreatment of MNG remains to be determined. Studies have used different doses and regimens or rhTSH administration, from as low as 0.01 or 0.03 mg to as high as 0.45 mg or   0.9 mg 24h before RAI treatment;
  • There is a significant occurrence of hypothyroidism after ¹³¹I treatment following rhTSH stimulation;
  • Although rare, autoimmune hyperthyroidism (approximate reported incidence of 4-5%) can develop after treatment of MNG with ¹³¹I;
  • Currently, rhTSH is not approved by the FDA as an adjuvant for ¹³¹I treatment of goiter.

 

Based on these results, pretreatment with rhTSH seems a promising alternative to thyroid surgery for the management of nontoxic MNG, particularly in elderly individuals. However, the optimal dose and timing of both, rhTSH and ¹³¹I as well as the criteria for patient eligibility remain to be determined.

Figure 17-8 – An elderly woman with a large and longstanding MNG that migrated to the upper mediastinal region with subsequent compression of the subclavian system. Note the subcutaneous enlarged venous circulation (a). In the next panel (b), it is presented the scintilographic studies after a tracer dose of 131I before and (c) after stimulation by 0.45 mg of recombinant human TSH. (ref. 142.).Figure 17-8 – An elderly woman with a large and longstanding MNG that migrated to the upper mediastinal region with subsequent compression of the subclavian system. Note the subcutaneous enlarged venous circulation (a). In the next panel (b), it is presented the scintilographic studies after a tracer dose of 131I before and (c) after stimulation by 0.45 mg of recombinant human TSH. (ref. 142.).

Surgery for MNG

As indicated by Fast et al (154) it is time to consider radioiodine treatment for MNG as an alternative to surgery. As indicated previously radioiodine (¹³¹I) is a simple, cost-effective and safe procedure with an impressive goiter reduction up to 65% of the original volume. Surgery of the MNG, however, is equally effective and the choice among the two procedures depends largely on their availability, clinical features, and last but not least the personal preference of the patient (and also the physician in charge). In many centers, specially in countries with large populations previously living in iodine deficiency, the number of patients with MNG, most of them, over 50 years old, are very common in the thyroid clinic daily routine. Therefore sending all those patients to surgery will inevitably, cause a logistic problem in terms of available surgical rooms, surgeons well trained in head and neck surgery, post surgical follow-up and all the costs involved. Moreover with the widespread use of ultrasonographic studies followed by Fine Needle Aspiration Biopsy (FNAB) the number of new cases of thyroid cancer has increased dramatically in the past few years. Obviously these patients will have precedence for a surgical therapy as compared with the patient with MNG. This situation is quite common in many countries where there is a long waiting list for a given patient to be selected for thyroid surgery. Frequently surgery of the thyroid due to a nodule harboring a papillary cancer in a relatively young subject has a definite preferential status over an elderly patient with a long standing MNG.

The preferred operation for MNG is subtotal thyroidectomy. The frequency of complications due to surgical depends on several factors and well-trained and experienced surgeons will reduce the rates of such complications. Recurrence after goiter surgery is rare and the frequency of hypothyroidism is low. It is advisable to introduce L-T4 therapy after surgery in order to avoid goiter recurrence although this option is considered highly controversial.

Laser Ablation Therapy

A retrospective study that assessed clinical records of 1534 benign nodules in 1531 patients treated with image-guided laser ablation therapy (LAT) showed that LAT induced a clinical relevant nodule volume reduction that ranges from 48 to 96% (72±11%) twelve months after treatment (158). This picture was more significantly in mixed nodules (range 70-92%, 79±7%). Most of the nodules (83%) received a single LAT dose while 13% received two doses and 3% three doses, with a total energy delivery based on the initial volume. The symptoms improved from 10% to 49% of cases (p<0.001), while major (voice changes) and minor complications (hematomas, skin burn) were rare. Thus, LAT is considered clinically effective and well-tolerated as an alternative to surgery for benign thyroid lesions.

To summarize: treatment of MNG with L-T4 suppressive doses is not accepted by many thyroidologists in spite of the fact that goiter reduction is achieved in one third of the patients and new nodules appearance is lower in the L-T4 treated patients.The results so far published using radioiodine preceded by rhTSH are quite encouraging . It is an excellent alternative when surgical teams are not available for all patients. LAT treatment should be performed only by experience operators. Finally, patients preference for a non- surgical alternative should always be taken into consideration.

SUMMARY

Perhaps the most common of all the disorders of the thyroid gland is multinodular goiter. Even in non-endemic regions it is clinically detected in about 4% of all adults beyond the age of 30. Pathologically it is much more frequent, the percentage incidence being roughly the same as the age of the group examined. The disease is much more common in women than in men.

Multinodular goiter is thought to be the result of primarily two factors. The first factor is genetic heterogeneity of follicular cells with regard to function (i.e. thyroid hormone synthesis) and growth. The second factor is the acquisition of new qualities that were not present in mother cells and become inheritable during further replication. Mutations may occur in follicular cells leading to constitutively activated adenomas and to hyperthyroidism. These factors may lead to loss of anatomical and functional integrity of the follicles and of the gland as a whole. These processes ultimately lead to goiter formation and are accelerated by stimulatory factors. These stimulatory factors are basically an elevated serum TSH, brought about by events such as iodine deficiency, inborn errors of thyroid hormone synthesis, goitrogens or local tissue growth-regulating factors. These basic and secondary factors may cause the thyroid to grow and gradually evolve into an organ containing hyperplastic islands of normal glandular elements, together with nodules and cysts of varied histologic pattern.

Nodular goiter is most often detected simply as a mass in the neck, but at times an enlarging gland produces pressure symptoms on the trachea and the esophagus. Occasionally tenderness and a sudden increase in size herald hemorrhage into a cyst. Hyperthyroidism develops in a large proportion of these goiters after a few decades frequently after iodine excess. Rare complications are paralysis of the recurrent laryngeal nerve, and pressure on the superior sympathetic ganglion causes a Horner´s syndrome.

The diagnosis is based on the physical examination. Thyroid function test results are normal or disclose subclinical or overt hyperthyroidism. Thyroid autoantibodies are usually absent or low, excluding Hashimoto´s thyroidits. Imaging procedures may reveal distortion of the trachea, calcified cysts, or impingement of the goiter on the esophagus. Sonographic studies, Scintilography (¹³¹I), CT and MRI are useful to detect details of the MNG and to provide an estimation of the volume before and after therapy.

From 4 to 17% of multinodular thyroids removed at operation contain foci that on microscopic examination fulfill the criteria of malignant change. The infrequency of thyroid cancer as a cause of death clearly proves that the vast majority of these lesions are not lethal or even clinically active. One of the reasons for the high incidence of cancer in surgical specimens is that patients with multinodular goiters were often selected for surgery because of a concern for carcinoma.

If a clinical and biochemically euthyroid multinodular goiter is small and produces no symptoms, treatment is controversial. T4 given in an effort to shrink the gland or to prevent further growth is effective in about one third of the patients. This therapy is more likely to be effective if begun at an early age while the goiter is still diffuse than in older patients in whom certain nodules may have already become autonomous. If the clinically euthyroid goiter is unsightly, shows subclinical hyperthyroidism or is causing, pressure symptoms, treatment with ¹³¹I preceded by recombinant human TSH is successful in virtually all cases but causes hypothyroidism at varying degree. Surgery is an acceptable alternative. The efficacy of T4 treatment after surgery, to prevent regrowth, is frequently used albeit debatable.

Overt toxic nodular goiter is usually treated with radioiodine. A gratifying reduction in the size of the goiter and control of the hyperthyroidism may be expected. Hypothyroidism often ensues.

During the past few years the use of recombinant human TSH has been used to enhance the uptake of radioiodine and to provide a more homogenous distribution of the radionuclide. Results have been rewarding with a 45-65% shrinkage of the MNG, even with an intrathoracic position. A surge of high levels of serum Free T4, total T3 and serum TG is observed in the first weeks after therapy. Clinically hyperthyroid patients seem to have more unwanted signs and symptoms as compared to euthyroid patients. Hypothyroidism (permanent) is commonly observed at 6-12 months after rhTSH plus RAI treatment. Taking all into account, this modality of treatment of MNG has a relatively low cost and it is considered a good alternative to surgery that might not be available for all patients with MNG in many centers around the world.

The term colloid is applied to glands composed of uniformly distended follicles appearing as a diffuse enlargement of the thyroid gland. The condition is found almost exclusively in young women. With time and due to a number of primary and secondary factors it may gradually develop into a multinodular goiter which becomes increasingly prominent as the decades pass. Appropriate therapy, if required, is the timely administration of thyroid hormone that may be continued for several years.

An intrathoracic goiter is usually an acquired rather than a development abnormality. It may come about in embryonic life by a carrying downward into the thorax of the developing thyroid anlage, or in adult life by protrusion of an enlarging thyroid through the superior thoracic inlet into the yielding mediastinal spaces. These lesions may produce pressure symptoms and may also be associated with hyperthyroidism. If too large for treatment with ¹³¹I, the appropriate therapy is resection of the goiter through the neck, if possible. Attachment of the intrathoracic goiter to the gland in the neck ordinarily proves the site of origin and provides a method for its easy surgical removal. In many of these patients a safe and easily performed therapy, in an outpatient mode, is the administration of a fixed dose of radioiodine (¹³¹I) of 30 mCi preceded by rhTSH.

 

REFERENCES

 

  1. Tunbridge WGM, Evered DC, Hall R, Appleton D, Brewis M, Clark F, Evans JG: The spectrum of thyroid disease in a community: The Whickham survey. Clin Endocrinol 7:481, 1977.
  2. Charib TGH, Thyroid incidentalomas: management approaches to non palpable nodules discovered incidentally on thyroid imaging. Ann Int Med 126:226-231, 1997.
  3. Sawin CT, Bigos ST, Land S, Bacharach P. The aging thyroid. Relationship between elevated thyrotropin levels and thyroid antibodies in elderly patients. Am J Med 79:591, 1985.
  4. Wang KW, Sum CF, Tan KT, Ng WY, Cheah JS, A study of non-toxic goiter. Ann Acad Med Singapore 19:439-442, 1990.
  5. Rallison ML, Dobyns BM, Meikle AW, et al. Natural history of thyroid abnormalities: prevalence, incidence, and regression of thyroid diseases in adolescents and young adults. Am J Med 91:363-370, 1991.
  6. Pinchera A, Aghini-Lombardi F, Antonangeli L, Vitti P. Multinodular goiter. Epidemiology and prevention. Ann Ital Chir 67:317-325, 1996.
  7. Jarlob AE, Nygaard B, Hegedus L, Hartling SC, Hansen JM: Oberver variation in the clinical and laboratory evaluation of patients with thyroid dysfunction and goiter. Thyroid 8:393-398, 1998.
  8. Marine D: Etiology and prevention of simple goiter. Medicine 3:453, 1924.
  9. Taylor S: The evolution of nodular goiter. J Clin Endocrinol Metab 13:1232, 1953.
  10. Beckers C, Cornette C: TSH production rate in nontoxic goiter. J Clin Endocrinol Metab 32:852, 1971.
  11. Dige-Petersen H, Hummer L: Serum thyrotropin concentrations under basal conditions and after stimulation with thyrotropin-releasing, hormone in idiopathic non-toxic goiter. J Clin Endocrinol Metab 44:1115, 1977.
  12. Smeulers J, Docter R, Visser TJ, Hennemann G: Response to thyrotrophin-releasing hormone and triiodothyronine suppressibility in euthyroid multinodular goiter. Clin Endocrinology 7:389, 1977.
  13. Berghout A, Wiersing WM, Smits NJ, Touber JL. Interrelationships between age, thyroid volume, thyroid nodularity, and thyroid function in patients with sporadic non-toxic goiter. Am J Med 89:602-608, 1995.
  14. Derwahl M, Studer H. Nodular goiter and goiter nodules: Where iodine deficiency falls short of explaining the facts. Exp Clin Endocrinol Diabetes 109:250-60, 2001.
  15. Studer H, Peter HJ, Gerber H: Natural heterogeneity of thyroid cells: The basis for understanding thyroid function and nodular growth. Endocr Rev 10:125, 1989.
  16. Peter JH, Gerber, Studer H, Smeds S: Pathogenesis of heterogeneity in human multinodular goiter. J Clin Invest 76:1992, 1985.
  17. Kopp P, Kimura ET, Aeschimann S, et al. Polyclonal and monoclonal nodules coexist with human multinodular goiters. J Clin Endocrinol Metab 79:134-139, 1994.
  18. Krohn K, Wohlgemuth S, Gerber H, Paschke R. Hot microscopic areas of iodine-deficient euthyroid goiters contain constitutively activating TSH receptor mutations. J Pathol 192:37-42, 2000.
  19. Holzapfel AP, Fuhrer D, Wonerow P, et al. Identification of constitutively activating somatic thyrotropin receptor mutations in a subset of toxic multinodular goiters. J Clin Endocrinol Metab 82:4292-4233, 1997.
  20. Tassi V, Di Cerbo A, Porcellini A, Papini E, Cisternino C, et al. Screening of thyrotropin receptor mutations by fine-needle aspiration biopsy in autonomous functioning thyroid nodules in multinodular goiters. Thyroid 9:353-357, 1999.
  21. Gabriel EM, Bergert ER, Grant CS, van Heerden JA, Thompson GB, et al. Germline polymorphism of codon 727 of human thyroid-stimulating receptor is associated with toxic multinodular goiter. J Clin Endocrinol Metab84:3328-3335, 1999.
  22. Krohn K, Führer D, Bayer Y, Eszlinger M, Brauer V, Neumann S, Paschke R. Molecular Pathogenesis of Euthyroid and Toxic Multinodular Goiter. Endocr Rev. 2005 June; 26(4):504-24.
  23. Masini-Repiso AM, Cobanillas AM, Bonaventura M, Coleoni AH. Dissociation of thyrotropin-dependent enzyme activities, reduced iodide transport, and preserved iodide organification in nonfunctioning adenoma and multinodular goiter. J Clin Endocrinol Metab 79:39, 1994.
  24. Hegedus L, Bonnema SJ, Bennedbek FN. Management of simple nodular goiter: current status and fature perspectives. Endocr Reviews 24:102-132, 2003.
  25. Krohn K, Paschke R. Progress in understanding the etiology of thyroid autonomy. J Clin Endocrinol Metab 86:3336-3345, 2001.
  26. Brix TH, Hededus L. Genetic and environmental factors in the aetiology of simple goiter. Ann Med 32:153-156, 2000.
  27. Brix TH, Kyvik KO, Hedegu¨s L. Major role of genes in the etiology of simple goiter in females: a population-based twin study. J Clin Endocrinol Metab 84:3071-3075, 1999.
  28. Brix TH, Kyvik KO, Hegedüs L. A population-based study of chronic autoimmune hypothyroidism in Danish twins. J Clin Endocrinol Metab 85:536-539, 2000.
  29. Medeiros-Neto G, Camargo RYC, Tomimori EK. Approach to and treatment of goiters. Med Clin N Am 92(2):351-358, 2012
  30. Knobel M & Medeiros-Neto G: An outline of inherited disorders of the thyroid hormone generating system. THYROID 13:771-801, 2003.
  31. Bignel GR, Canzian F, Shayeghi M, Stark M, Shugart YY, Biggs P, Mangion J, Hamoudi R, Rosenblatt J, Buu P, Sun S, Stoffer SS, Goldgar DE, Romeo G, Houlston RS, Narod SA, Stratton MR, Foulkes WD. Familial nontoxic multinodular thyroid goiter locus maps to chromosome 14q but does not account for familial nonmedullary thyroid cancer. Am J Hum Genet 61:1123-1130, 1997.
  32. Neumann S, Willgerodt H, Ackermann F, Reske A, Jung M, Reis A, Paschke R. Linkaged of familial euthyroid goiter to the multinodular goiter-1 locus and exclusion of the candidate genes thyroglobulin, thyroperoxidase and NA+/I- symporter. J Clin Endocrinol Metab 84:3750-3756, 1999.
  33. Mckay JD, Williamson J, Lesueur F, Stark M, Duffield A, Canzian F, Romeo G, Hoffman L. At least three genes account for familial papillary thyroid carcinoma: TCO and MNG1 excluded as susceptibility loci form a large Tasmanian family. Eur J Endocrinol 141:122-125, 1999.
  34. Capon F, Tacconelli A, Giardina E, Sciacchitano S, Bruno R, Tassi V, Trischitta V, Filetti S, Dallapiccola B, Novelli G. Mapping a dominant form of multinodular goiter to chromosome Xp22. Am J Hum Genet 67:1004-1007, 2000.
  35. Bayer Y, Neumann S, Meyer B. Genome-wide Linkage Analysis reveals evidence for four new susceptibility foci for familial euthyroid goiter. J Clin Endocr Metab 89:4044-53, 2004.
  36. Corval M, Perez R, Sanchez I, Mories MT, San Millan JL, Miralles JM, Gonzalez-Samiento R. Thyroglobulin gene point mutation associated with non-endemic simple goiter. Lancet 341:462-464, 1993.
  37. Takashi T, Nozaki JI, Komatsu M, Wada Y, Utsunomiya M, Inoue K, Takada G, Koisumi A. A new Locus for a dominant form of Multinodular Goiter on 3q26.1-q26.3 Bioch Biophy Res Comm 284:650-654, 2001.
  38. Samuels MH. Evaluation and treatment of sporadic nontoxic goiter-some answers and more questions. J Clin Endocrinol Metab 86:994-997, 2001.
  39. Bertelsen JB, Hegedüs L. Cigarette smoking and the thyroid. Thyroid 4:327-331, 1994.
  40. Brix TX, Hansen PS, Kyvik KO, Hegedüs L. Cigarette smoking and risk of clinically overt thyroid disease: a population-based twin case-control study. Arch Intern Med 160:661-666, 2000.
  41. Gaitan E. Environmental natural goitrogens. In: Peter F, Wiersinga WM, Hostalek U, eds. The thyroid and environment. New York: Schattauer; 69-78, 2000.
  42. Cheng YL. Birman KD, Schaudies RP, Ahmann AJ, d´Avis J, Geelhoed GW, Wartofsky L: Effects of epidermal growth factor on thyroglobulin and adenosine 3´,5´-monophosphate production by cultured human thyrocytes. J Clin Endocrinol Metab 69:771, 1989.
  43. Sugenoya A, Masuda H, Komatzu M, Jokojama S, Shimizu T, Fujimori M, Kobajashi S, Iida F. Adenomatous goitre: therapeutic strategy, postoperative outcome, and study of epidermal growth factor receptor. Brit J Surg 79:404, 1992.
  44. Maciel RM, Moses AC, Villone G, Tramontano D, Ingbar SH: Demonstration of the production and physiological role of insulin-like growth factor II in rat thyroid follicular cells in culture. J Clin Invest 82:1546, 1988.
  45. Phillips ID, Becks GP, Logan A, Wang JF, Smith C et al. Altered expression of insulin growth factor-1 (IGF-I) and IGF binding proteins during rat hyperplasia and involution. Growth Factors 10:207, 1994.
  46. Takahashi S-I, Conti M, Van Wyk JJ: Thyrotropin potentiation of insulin-like growth factor-I dependent deoxyribonucleic acid synthesis in FRTL-5 cells: Mediation by an autocrine amplification factor(s). Endocrinology 126:736, 1990.
  47. Vanelli GB, Barni T, Modigliani U, Paulin I, Serio M, Magi M, Fiorelli G, Balboni GC. Insulin-like growth factor-I receptors in non-functioning thyroid nodules. J Clin Endocrinol Metab 71:1175, 1990.
  48. deVito WJ, Chanoine J-P, Alex s, Fang S-L, Stone S. Effect of in vivo administration of recombinant acidic fibroblast growth factor on thyroid function in the rat: induction of colloid goiter. Endocrinology 131:729, 1992.
  49. Thompson SD, Franklyn JA, Watkinson JC, et al. Fibroblast growth factors 1 and 2 and fibroblast growth factor receptor 1 are elevated in thyroid hyperplasia. J Clin Endocrinol Metab 83:1336-1341, 1998.
  50. Francia FG, Azzolina L, Mantovani T, et al. Heterogeneity of nuclear DNA pattern and its relationship with cell cycle activity parameters in multinodular goiter. Clin Endocrinol 46:649-654, 1997.
  51. Bidey SP, Hill DJ, and Eggo MC. Growth factors and goitrogenesis. J Endocrinol 160:321-332, 1999.
  52. Gérard AC, Poncin S, Caetano B et al. Iodine deficiency induces a thyroid stimulating hormone-independent early phase of microvascular reshaping in the thyroid. Am J Pathol 172(3):748-60, 2008.
  53. Medeiros-Neto G and Knobel M. iodine deficiency disorders. In: deGroot LJ, Jameson JL, eds. Endocrinology 6th Ed. Chapter 88. New York, Elsevier, 2010.
  54. Hegedüs L. Thyroid size determined by ultrasound. Influence of physiological factors and non-thyroidal disease. Dan Med Bull 37:249-263.
  55. Rubio IGS, Medeiros-Neto G. Mutations of the thyroglobulin gene and its relevance to thyroid disorders. Curr Opin Endocrinol Diabetes Obes. 16(5):373-8, 2009.
  56. Pelizzo MR, Piotto A, Rubello D, Casara D, Fassina A, Busnardo B. High prevalence of occult papillary thyroid carcinoma in a surgical series for benign thyroid diseases. Tumori 76:255,1990.
  57. McCall A, Jarosz H, Lawrence AM, Paloyan E. The incidence of thyroid carcinoma in solitary cold nodules and in multinodular goiters. Surgery 100:1128, 1986.
  58. Koh KB, Chang KW. Carcinoma in multinodular goiter. Brit J Surg 79:266, 1992.
  59. Bisi H, Fernandes SO, Camargo RYA, Koch L, Abdo AH, Brito T. The prevalence of unsuspected thyroid pathology in 300 sequential autopsies with special reference to the incidental carcinoma. Cancer 64:1888-1893, 1989.
  60. Lang W, Borrusch H, Bauer L. Occult carcinomas of the thyroid. Evaluation of 1020 sequential autopsies. Am J Clin Pathol 90:72, 1988.
  61. Stanbury JB, Ermans AE, Bourdoux P, Todd C, Oken E, Tonglet R, Vidor G, Braverman LE, and Medeiros Neto G. Iodine-induced hyperthyroidism: Occurrence and epidemiology. Thyroid 8(1):83-100, 1998.
  62. Fukunaga FH, Lockett LJ: Thyroid carcinoma in the Japanese in Hawaii. Arch Pathol Lab Med 92:6, 1971.
  63. Sampson RJ, Key CR, Buncher CR, Iijima S: Smallest form of papillary carcinoma of the thyroid. Arch Pathol Lab Med 91:334, 1971.
  64. Sampson RJ, Woolner LB, Bahn RC, Kurland LT: Occult thyroid carcinoma in Olmsted County, Minnesota: Prevalence at autopsy compared with that in Hiroshima and Nagasaki. Jpn Cancer 34:2072, 1974.
  65. Campbell MJ, Seib CD, Candell L, Gosnell JE, Duh QY, Clark OH, Shen WT. The underestimated risk of cancer in patients with multinodulargoiters after a benign fine needle aspiration. World J Surg. 39(3):695-700, 2015.
  66. Brito JP, Yarur AJ, Prokop LJ, McIver B, Murad MH, Montori VM. Prevalence of thyroid cancer in multinodulargoiter versus single nodule: a systematic review and meta-analysis. 23(4):449-55, 2013. Review.

 

  1. Pasqualetti G, Caraccio N, Basolo F, Miccoli P, Monzani F.Prevalence of thyroid cancer in multinodulargoiter versus single nodule: iodine intake and cancer phenotypes. 24(3):604-5, 2014
  2. Riccabona OA. Thyroid cancer: its epidemiology, clinical features and treatment. Springer Verlag 1987.
  3. Tomusch O, Machens A, Sekulla C, Ukkat J, Lippert H, Gastinger I, Dralle H. Multivariate analysis of risk factors for postoperative complications in benign goiter surgery: prospective multicenter study in Germany. World J Surg 24:1335-1341, 2000.
  4. Abdel Rahim AA, Ahmed ME, Hassan MA. Respiratory complications after thyroidectomy and the need for tracheostomy in patients with a large goiter. Br J Surg 86:88-90, 1999.
  5. Mariotti RA, Zannini P, Viani MP, Voci C, Pezzuoli G. Surgical treatment of substernal goiters. Int Surg 76:12-17, 1991.
  6. Torre G, Borgonovo G, Amato A, Arezzo A, Ansaldo G, De Negri A, Ughe M, Mattioli F. Surgical management of substernal goiter: analysis of 237 patients. Am Surg 61:826-831, 1995.
  7. Vadasz P, Kotsis L. Surgical aspects of 175 mediatinal goiters. Eur J Cardiothorac surg 14:393-397, 1998.
  8. Allo MD, Thompson NW. Rationale for the operative management of substernal goiters. Surgery 94:969-977, 1983.
  9. Röjdmark J, Järhult J. High long term recurrence rate after subtotal thyroidectomy for nodular goiter. Eur J Surg 161:725-727, 1995.
  10. Berghout A, Wiersinga WM, Drexhage HA, van Trotsenburg P, Smits NJ, van der Gaag RD, Touber JL. The long-term outcome of thyroidectomy for sporadic non-toxic goitre. Clin Endocr (Oxf) 31:193-199, 1989.
  11. Geerdsen JP, Frolund L. Thyroid function after surgical treatment of nontoxic goitre. A randomized study of postoperative thyroxine administration. Acta Med Scand 220:341-345, 1986.
  12. Bistrup C, Nielsen JD, Gregersen G, Franch P. Preventive effect of levothyroxine in patients operated for non-toxic goitre: a randomized trial of one hundred patients with nine years follow-up. Clin Endocrinol (Oxf) 40:323-327, 1986.
  13. Miccoli P, Antonelli A, Iacconi P, Alberti B, Gambuzza C, Baschieri L. Prospective, randomized, double-blind study about effectiveness of levothyroxine suppressive therapy in prevention of recurrence after operation: result at the third year of follow-up. Surgery 114:1097-1101, 1993.
  14. Feldkamp J, Seppl T, Becker A, Klisch A, Schlagheck R, Goretzki PE, Roher HD. Iodide or L-thyroxine to prevent recurrent goiter in an iodine-deficient area: prospective sonographic study. World J Surg 21:10-14, 1997.
  15. Seiler CA, Glaser C, Wagner HE. Thyroid gland surgery in an endemic region. World J Surg 20:593-596, 1996.
  16. Mishra A, Agarwal A, Agarwal G, Mishra SK. Total thyroidectomy for benign thyroid disorders in an endemic region. World J Surg 25:307-310, 2001.
  17. Pappalardo G, Guadalaxara A, Frattaroli FM, Illomei G, Falaschi P. Total compared with subtotal thyroidectomy in benign nodular disease: personal series and review of published reports. Eur J Surg 164:501-506, 1998.
  18. Hisham AN, Azlina AF, Aina EN, Sarojah A. Total thyroidectomy: the procedure of choice for multinodular goitre. Eur J Surg 167:403-405, 2001.
  19. Wiest PW, Hartshorne MF, Inskip PD, Crooks LA, Vela BS, Telepak RI, Williamson MR, Blumhardt R, Bauman JM, Tekkel M. Thyroid palpation vs high-resolution thyroid ultrasonography in the detection of nodules. J Ultrasound Med 17:487-496, 1998.
  20. Borghout A, Wiersinga WM, Smits NJ, Touber JL. The value of thyroid volume measured by ultrasonography in the diagnosis of goitre. Clin Endocrinol (Oxf) 28:409-414, 1988.
  21. Knudsen N, Bols B, Bülow I, Jorgensen T, Perrild H, Ovesen L, Laurberg P. Validation of ultrasonography of the thyroid gland for epidemiological purposes. Thyroid 9:1069-1074, 1999.
  22. Ezzat S, Sarti DA, Cain DR, Braunstein GD. Thyroid incidentalomas. Prevalence by palpation and ultrasonography. Arch Intern Med 154:1838-1840, 1994.
  23. Tomimori E, Pedrinola F, Cavaliere H, Knobel M, Medeiros-Neto G. Prevalence of incidental thyroid disease in a relatively low iodine intake area. Thyroid 5:273-276, 1995.
  24. Brunn J, Block U, Ruf G, Bos I, Kunze WP, Scriba PC. [Volumetric analysis of thyroid lobes by real-time ultrasound.] Dtsch Med Wochenschr 106:1338-1340, 1981.

 

  1. Reinartz P, Sabri O, Zimmy M, Nowak B, Cremerius U, Setani K, Bull U. Thyroid volume measurement in patients prior to radioiodine therapy: comparison between three-dimensional magnetic resonance imaging and ultrasonography. Thyroid 12:713-717, 2002.
  2. Hegedus L, Perrild H, Poulsen LR, Andersen JR, Holm B, Schnohr P, Jensen G, Hansen JM. The determination of thyroid volume by ultrasound and its relationship to body weight, age, and sex in normal subjects. J Clin Endocrinol Metab 56:260-263, 1983.
  3. Nygaard B, Nygaard T, Court-Payen, Jensen LI, Soe-Jensen P, Gerhard NK, Fugl M, Hegedus L. Thyroid volume measured by ultrasonography and CT. Acta Radiol 43:269-274, 2002.
  4. Bonnema SJ, Andersen PB, Knudsen DU, Hegedus L. MR imaging of large multinodular goiters: observer agreement on volume vs observer disagreement on dimensions of the involved trachea. AJR Am J Roentgenol 179:259-266, 2002.
  5. Schlogl S, Werner E, Lassmann M, Terekhova J, Muffert S, Seybold S, Reiners C. The use of three-dimensional ultrasound for thyroid volumetry. Thyroid 11:569-574, 2001.
  6. Rivo-Vázquez Á, Rodríguez-Lorenzo Á, Rivo-Vázquez JE, Páramo-Fernández C, García-Lorenzo F, Pardellas-Rivera H, Casal-Núñez JE, Gil-Gil P. The use of ultrasound elastography in the assessment of malignancy risk in thyroid nodules and multinodular Clin Endocrinol (Oxf). 79(6):887-91, 2013.
  7. Hays MT, Wesselossky B. Simultaneous measurement of thyroidal trapping (99 mTcO4-) and binding (¹³¹I-): clinical and experimental studies in man. J Nucl Med 14:785-792, 1973.
  8. Arnold JE. Pinsky S. Comparison of 99 mTc and ¹²³I for thyroid imaging. J Nucl Med 17:261-267, 1976.
  9. Dige-Petersen H, Kroon S, Vadstrup S, Andersen ML, Roy-Poulsen NO. A comparison of 99Tc and ¹²³I scintigraphy in nodular thyroid disorders. Eur J Nucl Med 3:1-4, 1978.
  10. Ryo UY, Vaidya PV, Schneider AB, Bekerman C, Pinsky SM. Thyroid imaging agents: a comparison of I-123 and TC-99m pertechnetate. Radiology 148:819-822, 1983.
  11. Wesche MF, Tiel-Van Buul MM, Smits NJ, Wiersinga WM. Ultrasonographics vs. scintigraphic measurement of thyroid volume in patients referred for 131I therapy. Nucl Med Commun 19:341-346, 1998.
  12. Wanet PM, Sand A, Abramovici J. Physical and clinical evaluation of high-resolution thyroid pinhole tomography. J Nucl Med 37:2017-2020, 1996.
  13. Jennings A. Evaluation of substernal goiters using computed tomography and MR imaging. Endocrinol Metab Clin North Am 30:401-414, 2001.
  14. Belardinlli L, Gualdi G, Ceroni L, Guadalaxara A, Polettini E, Pappalardo G. Comparison between computed tomography and magnetic resonance data and pathologic findings in substernal goiters. Int Surg 80:65-69, 1995.
  15. Hermans r, Bouillon R, LagaK, Delaere PR, Foer BD, Marchal G, Baer AL. Estimation of thyroid gland volume by spiral computed tomography. Eur Radiol. 7:214-216, 1997.
  16. Gullu S, Gurses MA, Baskal N, Uysal AR, Kamel AN, Erdogan G. Suppressive therapy with levothyroxine for euthyroid diffuse and nodular goiter. Endocr J 46:221-226, 1999
  17. Peters H, Hackel D, Schleusener H. Treatment of euthyroid struma. Comparable volume reduction with 400 micrograms iodine, 100 micrograms levothyroxine combined with 100 micrograms iodine or individually dosed levothyroxine. Med Klin 92:63-67, 1997.
  18. Schumm-Draeger PM. Drug therapy of goiter. Iodine, thyroid hormones or combined therapy. Z Gesamte Inn Med 48:592-598, 1993.
  19. Lima N, Knobel M, Cavaliere H, Sztejnsznajd, Tomimori E, Medeiros-Neto G. Levothyroxine suppressive therapy is partially effective in treating patients with benign, solid thyroid nodules and multinodular goiters. Thyroid 7:691-697, 1997.
  20. Wesche MF, Tiel-Van Buul MM, Lips P, Smits NJ, Wiersinga WM. A randomized trial comparing levothyroxine with radioactived iodine in the treatment of sporadic nontoxic goiter. J Clin Endocrinol Metab 86:998-1005, 2001.
  21. Papini E, Petrucni R, Guglielmi R, Panunzi C, Rinaldi R, Bacci V, Crescenzi A, Nardi F, Fabbrini R, Pecella CM. Long Term changes in Nodular Goiter: a 5 yr prospective randomized trial of L-T4 suppressive therapy for benign thyroid nodules. J Clin Endocr Metab 83:780-783, 1998.
  22. Zelmanovitz F, Genro S, Gross JL. Suppressive therapy with L-T4 for solitary thyroid nodules: a double blind controlled clinical study and cumulative Meta-Analysis. J Clin Endocr Metab 83:3881-3885, 1998.
  23. Berghout A, Wiersinga WM, Smits NJ, Touber JL. Interrelationships between age, thyroid volume, thyroid nodularity, and thyroid function in patients with sporadic non-toxic goiter. Am J Med 89:602-608,1995.

 

  1. Leese GP, Jung RT, Guthrie C, Waugh N, Browning MC. Morbidity in patients on L-thyroxine: a comparison of those with a normal TSH to those with a suppressed TSH. Clin Endocrinol (Oxf) 37:500-503, 1992.
  2. Uzzan B, Campos J, Cucherat M, Nony P, Boissel JP, Perret GY. Effects on bone mass of long term treatment with thyroid hormones: a meta-analysis. J Clin Endocrinol Metab 81:4278-4289, 1996.

 

  1. Hegedüs L, Hansen BM, Knudsen N, Hansen JM: Reduction of size of thyroid with radioactive iodine in multinodualr nontoxic goitre. Brit Med J 297:661, 1988.
  2. Hamburger JI, Hamburger SW: Diagnosis and management of large toxic multinodular goiters. J Nucl Med 26:888, 1985.
  3. Kay TW, d´Emden MC, Andrews JT, Martin FI: Treatment of non-toxic multinodular goiter with radioactive iodine. Am J Med 84:19, 1988.
  4. Huysmans D, Hermus A, Edelbroek M, et al. Radioiodine for nontoxic multinodular goiter. Thyroid 7:235-239, 1997.
  5. Nygaard B, Hegedüs L, Gervil M, Hjalgrim H, Soe-Jensen G et al. Radioiodine treatment of multinodular non-toxic goitre. Br Med J 307:828, 1993.
  6. Danaci M, Veek CM, Notghi A, Merck MV, Padfield PL, Edwards CR. 131I radioiodine therapy for hyperthyroidism in patients with Graves´ disease, uninodular goiter and multinodular goiter. N Z Med J 101:784, 1988.
  7. Nygaard B, Hegedus L, Ulriksen P, Nielsen KG, Hansen JM. Radioiodine therapy for multinodular toxic goiter. Arch Intern Med 159:1364-1368, 1999.
  8. Aach R, Kissane J: Thyroid storm shortly after 131I therapy of a toxic multinodular goiter. Am J Med 52:786, 1972.
  9. Huysmans Ak, Hermus RM, Cortstens FAM, Barents JO, Kloppenborgh PWC. Large compressive goiters treated with radioiodine. Ann Int Med 121:757, 1994.
  10. Nygaard B, Hegedus L, Gervil M, Hjalgrim H, Soe-Jensen P, Hansen JM. Radioiodine treatment of multinodular non-toxic goitre. BMJ 307(6908):828-32, 1993.
  11. Nielsen VE, Bonnema SJ, Hegedus L. The effects of recombinant human thyrotropin, in normal subjects and patients with goitre. Clin Endocrinol (Oxf) 61(6)655-63, 2004.
  12. Nielsen VE, Bonnema SJ, Boel-Jorgensen H, Grupe P, Hegedus L. Stimulation with 0.3 mg recombinant human thyrotropin prior to iodine 131I therapy to improve the size reduction of benign nontoxic nodular goiter: a prospective randomized double-blind trial. Arch Intern Med 166(14):1476-82, 2006.
  13. Nieuwlaat Wa, Huysmans DA, van den Bosch HC, Sweep CG, Ross HA, Corstens FH, et al. Pretreatment with a single, low dose of recombinant human thyrotropin allows dose reduction of radioiodine therapy in patients with nodular goiter. J Clin Endocrinol Metab 3121-9, 2003.
  14. Weetman AP. Radioiodine treatment for benign thyroid diseases. Clin Endocrinol (Oxf) 66(6):757-64, 2007.
  15. Huysmans DA, Nieuwlaat WA, Erdtsieck RJ, Schellekens AP, Bus JW, Bravenboer B. Administration of a single low dose of recombinant human thyrotropin significantly enhances thyroid radioiodide uptake in nontoxic nodular goiter. J Clin Endocrinol Metab 85(10):3592-6, 2000.
  16. Albino CC, Mesa CO, Jr, Olandoski M, Ueda CE, Woellner LC, Goedert CA, et al. Recombinant human thyrotropin as adjuvant in the treatment of multinodular goiters with radioiodine. J Clin Endocrinol Metab 90(5):2775-80, 2005.
  17. Cohen O,m Ilany J, Hoffman C, Olchovsky D, Dabhi S, Karasik A. Low-dose recombinant human thyrotropin-aided radioiodine treatment of large, multinodular goiters in elderly patients. Eur J Endocrinol 154(2):243-52, 2006.
  18. Giusti M, Caorsi V, Mortara L, Caputo M, Monti E, Schiavo M, Bagnara MC, Minuto F, Bagnasco M. Long-term outcome after radioiodine therapy with adjuvant rhTSH treatment: comparison between patients with non-toxic and pre-toxic large multinodular goitre. Endocrine (2):221-9, 2014.
  19. Duick DS, Baskin HJ. Utility of recombinant human thyrotropin for augmentation of radioiodine uptake and treatment of nontoxic and toxic multinodular goiters. Endocr Pract 9(3):204-9, 2003.
  20. Duick DS, Baskin HJ. Significance of radioiodine uptake at 72 hours versus 24 hours after pretreatment with recombinant human thyrotropin for enhancement of radioiodine therapy in patients with symptomatic nontoxic multinodular goiter. Endocr Pract 10(3):253-60, 2004.
  21. Fast S, Hegedüs L, Pacini F, Pinchera A, Leung AM, Vaisman M, Reiners C, Wemeau JL, Huysmans DA, Harper W, Rachinsky I, de Souza HN, Castagna MG, Antonangeli L, Braverman LE, Corbo R, Düren C, Proust-Lemoine E, Marriott C, Driedger A, Grupe P, Watt T, Magner J, Purvis A, Graf H. Long-term efficacy of modified-release recombinant human thyrotropin augmented radioiodine therapy for benign multinodulargoiter: results from a multicenter, international, randomized, placebo-controlled, dose-selection study. 2014.
  22. Paz-Filho GJ, Mesa-Junior CO, Olandoski M, Woellner LC, Goedert CA, Boguszewski CL, et al. Effect of 30 mCi radioiodine on multinodular goiter previously treated with recombinant human thyroid stimulating hormone. Braz J Med Biol Res 40(12):1661-70, 2007.
  23. Nielsen VE, Bonnema SJ, Boel-Jorgensen H, Veje A, Hegedus L. Recombinant human thyrotropin markedly changes the 131I kinetics during 131I therapy of patients with nodular goiter: an evaluation by a randomized double-blinded trial. J Clin Endocrinol Metab 90(1):79-83, 2005.
  24. Niewlaat WA, Hermus AR, Sivro-Pmdelj F, Corstens FH, Huysmans DA. Pretreatment with recombinant human TSH changes the regional distribution of radioiodine on thyroid scintigrams of nodular goiters. J Clin Endocrinol Metab 86(11):5330-6, 2001.
  25. Giusti M, Cappi C, Santaniello B, Ceresola E, Augeri C, Lagasio C, et al. Safety and efficacy of administering 0.2 mg of recombinant human TSH for two consecutive days as an adjuvant to therapy with low radioiodine doses in elderly out-patients with large nontoxic multinodular goiter. Minerva Endocrinol 31(3):191-209, 2006.
  26. Bonnema SJ, Nielsen VE, Boel-Jorgensen H, Grupe P, Andersen PB, Bastholt L, et al. Improvement of goiter volume reduction after 0.3 mg recombinant human thyrotropin-stimulated radioiodine therapy in patients with a very large goiter: a double-blinded, randomized trial. J Clin Endocrinol Metab 92(9):3424-8, 2007.
  27. Silva MN, Rubio IG, Romao R, Gebrin EM, Buchpiguel C, Tomimori E, Camargo RYA, Cardia MS, Medeiros-Neto G. Administration of a single dose of recombinant human thyrotropin enhances the efficacy of radioiodine treatment of large compressive multinodular goitres. Clin Endocrinol (Oxf) 60(3):300-8, 2004.
  28. Paz-Filho G, Mesa C, Carvalho G, Goedert C, Graf H. Recombinant human TSH associated with radioiodine does not have further effects on thyroid volume and function after two years. Clin Endocrinol (Oxf) 2007 (Letter|).
  29. Nieuwlaat WA, Hermus AR, Ross HA, Buijs WC, Edelbroek MA, Bus JW, et al. Dosimetry of radioiodine therapy in patients with nodular goiter after pretreatment with a single, low dose or recombinant human thyroid-stimulating hormone. J Nucl Med 45(4):626-33, 2004.
  30. Nielsen VE, Bonnema SJ, Hegedus L. Transient goiter enlargement after administration of 0.3 mg of recombinant human thyrotrophin in patients with benign nontoxic nodular goiter: a randomized, double-blind, crossover trial. J Clin Endocrinol Metab 91(4):1317-22, 2006.
  31. Fast S., Nielsen VE, Grupe P, Boel-Jaegersen H, Bastholt L, Andersen PB, Bonnema SJ, Hegedus L. Prestimulation with recombinant human thyrotropin (rhTSH) improves the long-term outcome of radioiodine therapy for multinodular nontoxic goiter. J Clin Endocrinol Metab 98::2653-2660, 2012
  32. Cubas ER, Paz-Filho GJ, Olandoski M, Goedert CA, Woellner LC, Carvalho GA, Graf H. Recombinant human TSH increases the efficacy of a fixed activity of radioiodine for treatment of multinodular goitre. Int J Clin Pract 63:585-590, 2009.
  33. Romão R, Rubio IGS, Tomimori EK, Camargo RYA, Knobel M, Medeiros-Neto G. High prevalence of side effects after rhTSH-stimulated radioiodine treatment with 30 mCi in patients with multinodular goiter and subclinical / clinical hyperthyroidism. Thyroid 19:945-951, 2009.
  34. Cardia MS, Rubio IG, Medeiros-Neto G. Prolonged follow-up of multinodular goitre patients treated with radioiodine preceded or not by human recombinant TSH. Clin Endocrinol (Oxf) 64(4):474, 2006.
  35. Barca MF, Gruppi C, Oliveira MT, Romao R, Cardia MS, Rubio IGS, Knobel M, Medeiros-Neto G. Cardiovascular assessment of hyperthyroid patients with multinodular goiters before and after radioiodine treatment preceded by stimulation with recombinant human TSH. Endocrine 70:810-813, 2009.
  36. Nygaard B, Knudsen JH, Hedegus L, Scient AV, Hansen JE. Thyrotropin receptor antibodies and Graves´ disease, a side-effect of 131I treatment in patients with nontoxic goiter. J Clin Endocrinol Metab. 82(9):2926-30, 1997.
  37. Rubio IG, Perone BH, Silva MN, Knobel M, Medeiros-Neto G. Human recombinant TSH preceding a therapeutic dose of radioiodine for multinodular goiters has no significant effect in the surge of TSH-receptor and TPO antibodies. Thyroid 15(2):134-9, 2005.
  38. Medeiros-Neto G, Marui S, Knobel M. An outline concerning the potential use of recombinant human thyrotropin for improving radioiodine therapy of multinodular goiter. Endocrine 33:109-117, 2009.
  39. Fast S, Nielsen UE, Bonnema SJ, Hegedus L. Time to reconsider non- surgical therapy of benign non-toxic multinodular goiter: focus on recombinant human TSH augmented radioiodine therapy. Eur J Endocrinol 160:517-528, 2009.
  40. Magner J. Problems associated with the use of thyrogen in patients with a thyroid gland. N Engl J Med 359:1738-9, 2008.
  41. Fast S, Nielsen VE, Bonnema SJ, Hegedus L. Optimizing 131I uptake after rhTSH stimulation in Patients with multinodular goiter: evidence from a prospective double blind study. J Nucl Med 50:732-737, 2009.
  42. Bonnema SJ, Hegedus L, A 30 year perspective on radioiodine therapy of benign nontoxic multinodular goiter. Cur Op Endocr Metab Diab Obes. 16:379-384, 2009.
  43. Pacella CM,Mauri GAchille GBarbaro DBizzarri GDe Feo PDi Stasio EEsposito RGambelunghe GMisischi I,Raggiunti BRago TPatelli GLD'Este SVitti PPapini E. Outcomes and Risk Factors for Complications of Laser Ablation for Thyroid Nodules: A Multicenter Study on 1531 Patients. J Clin Endocrinol Metab. 100(10):3903-10, 2015.