REVIEWS

Hormonal control of T‑cell development in health and disease

Wilson Savino1, Daniella Arêas Mendes-da-Cruz1, Ailin Lepletier1 and Mireille Dardenne2 Abstract | The physiology of the , the primary lymphoid organ in which T cells are generated, is controlled by . Data from animal models indicate that several and nonpeptide hormones act pleiotropically within the thymus to modulate the proliferation, differentiation, migration and death by apoptosis of developing thymocytes. For example, growth and can enhance thymocyte proliferation and migration, whereas glucocorticoids lead to the apoptosis of these developing cells. The thymus undergoes progressive age-dependent atrophy with a loss of cells being generated and exported, therefore, hormone-based therapies are being developed as an alternative strategy to rejuvenate the organ, as well as to augment thymocyte proliferation and the export of mature T cells to peripheral lymphoid organs. Some hormones (such as and progonadoliberin‑1) are also being used as therapeutic agents to treat immunodeficiency disorders associated with thymic atrophy, such as HIV infection. In this Review, we discuss the accumulating data that shows the thymus gland is under complex and multifaceted hormonal control that affects the process of T‑cell development in health and disease.

Once formed in the thymus, mature T cells migrate and hypothalamic–growth-hormone/prolactin axes. to peripheral lymphoid organs to initiate the cell- Additionally, the nervous system can influence the thymus mediated immune response. Importantly, hormones via direct innervation of blood vessels and parenchymal can control thymus physiology, including T‑cell devel- cells, which has been reviewed elsewhere7. opment, and the thymus itself can also affect endocrine Given the variety of interactions between neurons, axes1. Investigations in the 1970s and 1980s showed endocrine and immune cells, that dysfunctions in one that neo­natal thymectomy generated abnormal devel- of these systems can affect the other is unsurprising, opment of secondary sexual organs2, and that a given and can lead to diseases such as those observed during cell-mediated antigen stimulation triggered a feedback ageing and several infectious diseases, such as Chagas response invol­ving the production of glucocorticoids disease and malaria. In this Review, we describe the and IL‑13. During the 1980s, the existence of a com- effects of hormones of the functioning of the thymus, mon syntax in neuro­endocrine and lymphoid tissues as well as their role in T‑cell development in both was also proposed4. Accordingly, immune cells can healthy and diseased states.­ 1Laboratory of Thymus produce and release mol­ecules classically defined as Research, Oswaldo Cruz hormones, and cells of the nervous and endocrine T‑cell development Institute, Oswaldo Cruz tissues also produce cytokines5. The entire process of T‑cell differentiation takes place Foundation, Avenue Brasil 4365, 21045-900, In pioneering work, stress-induced activation within a 3D network called the thymic microenviron­ Manguinhos, Rio de Janeiro, of the –pituitary–adrenal axis was ment (FIG. 1). This network is composed of non­lymphoid Brazil. shown to trigger severe atrophy of the thymus6. Since cells (such as thymic epithelial cells (TECs)), extra­cellular 2Hôpital Necker, CNRS UMR this research was published, the cross-talk between matrix and soluble moieties such as cyto­kines or chemo­ 8147, Université Paris the immune and neuroendocrine systems has been kines, as well as classic thymic hormones such as thymu­ Descartes, 75015 Paris, 8,9 France. ext­ended to other endo­crine axes, including those medi- lin, thymo­poietin and . Other hormones, Correspondence to W.S. ated by the interactions­ between lymphocytes and micro­ including growth hormone (GH), prolactin, [email protected] environmental cells in the thymus, spleen and lymph and glucocorticoids, which are pro­duced by the endo­ doi:10.1038/nrendo.2015.168 nodes and hormones secreted by the hypothalamic– crine glands such as the pituitary gland and the adrenals, Published online 6 Oct 2015 pituitary–gonadal, hypothalamic–pituitary– are also involved in this thymic microenvironment10–12.

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 77

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

Key points In the medulla of the thymus, peripheral tissue antigens (PTAs) are expressed by TECs21. These PTAs • The thymus is the primary lymphoid organ responsible for the generation of T cells are proteins or specifically expressed in given • Thymus physiology and T‑cell development can be controlled by hormones, tissues of the body that represent self-antigens of most via a variety of endocrine and paracrine pathways paren­chymal organs, such as , which is expressed • Microenvironmental cells in the thymus constitutively produce hormones that are by the endo­crine cells of pancreatic islets of Langerhans. typically secreted by the pituitary gland, such as growth hormone, prolactin, oxytocin PTA expression is controlled by AIRE expression, which and enables the presentation of self-peptides by medullary • Glucocorticoids induce thymocyte depletion through caspase-dependent apoptosis, TECs (mTECs) to developing thymocytes22,23. This whereas growth hormone enhances thymocyte proliferation and migration process avoids the development of self-antigen reactive • Considering the variety of the interactions between the endocrine, the nervous and cells and consequently prevents autoimmunity. Mature the immune systems, dysfunctions in one of these systems can affect the other SP thymocytes that have survived this negative selec- • Acute infection by Trypanosoma cruzi (the causative agent of Chagas disease) induces tion process can ultimately emigrate from the thymus thymic atrophy through glucocorticoid-mediated thymocyte depletion, which can be towards secondary lymphoid organs (that is, the spleen, counteracted by exogenous prolactin lymph nodes, tonsils and appendix) and constitute the peripheral pool of T lymphocytes (FIG. 1a)24,25. During differentiation, thymocytes migrate Lymphocyte precursors that originate from the within the thymic lobules. DN cells are found in the bone marrow enter the thymus through postcapillary sub-capsular outer cortex of the thymic lobules, whereas venules located close to the cortico-medullary region DP cells are found in the inner cortex and mature SP of thymic lobules (FIG. 1a)13. Developing thymocytes thymocytes are located in the medulla (FIG. 1a)24,25. differen­tially express molecules that characterize Developing thymocytes interact with different compo- well-defined stages of normal T‑cell development. nents of the thymic microenvironment in various ways Among these molecules,­ the expression of the T‑cell including by direct cell–cell interaction, as well as with receptor (TCR) at the membrane, together with the extra­cellular matrix and soluble moieties, such as accessory receptors CD4 and CD8, can be used to trace , chemo­kines, sphingolipids and hormones intrathymic differenti­ation, particularly by using cyto- (FIGS 1b,2)26–28. In early intrathymic T‑cell development, fluorometry. The TCR is a hetero­dimer (mostly formed a key cell–cell interaction is mediated by neurogenic by one α and one β chain) that results from somatic locus notch homo­logue protein 1 (commonly known recombination of distinct gene segments (known as V, as Notch), which is expressed by T-cell precursors, and D and J)14. Such a mechanism enables the generation the protein Delta-like protein 4 (Dll4), expressed by the of a vast number of distinct TCR, which collectively thymus epithelium29. Disruption of Dll4 expression in form the T‑cell repertoire. Up to 1020 events of recom- TEC completely impedes thymopoiesis30. bination can occur during thymus selection, which is The thymus undergoes physiological age-dependent associated with the generation of 108 T cells in mice atrophy, owing to the depletion of developing thymo- and 1012 T cells in humans15. cytes, together with changes in the microenvironmen- Most immature thymocytes express neither the tal com­partments, and infiltration with , TCR, nor CD4 and CD8, and are called double-nega- particularly secondary to fibroadipogenetic transfor- tive (DN) thymocytes, as they lack both CD4 and CD8 mation31–33. The cortico-medullary tissue architecture is on their membranes. Cells that start to express both progressively lost, which occurs in parallel with a decline CD4 and CD8 co-receptors become double-positive in the export of thymocytes34. This process affects the (DP). At this stage, developing thymocytes express peripheral T‑cell compartment, with a consequent low levels of the membrane TCR, and can be pheno­ decrease in immuno­responsiveness as an indivi­dual typed as TCRlowCD4+CD8+16. These thymocytes ages35. In addition, the epithelial compartment of ageing occupy most of the cortical region of thymic lobules. mice is altered in both cortical and medullary regions, During this phase, DP cells undergo positive selec- with reduced numbers of TECs (including mTECs tion, which enables cells that express TCRs to interact expressing AIRE) and conse­quently reduced expression with the microenvironmental­ cells and continue dif- of PTAs36,37, which in theory favours the generation of ferentiation, whereas those that do not express TCRs autoimmune events in ageing individuals. The molec- undergo apoptosis17. ular mechanisms underlying age-­dependent thymic As thymocytes differentiate further, these cells atrophy are not fully understood. Even so, a number of enhance the expression of TCR and cease to express pre-clinical and clinical approaches to rejuvenate the either CD4 or CD8, thereby becoming single- organ have already been developed that result in sub- positive (SP) cells with the phenotypes of either stantial improvement in thymus function, such as sex TCRhighCD4+CD8– or TCRhighCD4–CD8+18. These SP steroid ablation and administration of GH and IL‑734,38. cells undergo apoptosis if they have a high or very low avidity when interacting with a TCR and the endogen­ Thymic hormone-mediated circuits ous peptide coupled to major histocompatibility com- Hormonal control of the thymus includes both endo­crine plex (MHC) class I or class II molecule expressed by and paracrine/autocrine pathways that act on thymic microenvironmental cells in the thymus (particularly microenvironment cells and thymocytes via specific TECs and thymic dendritic cells (TDCs))17,19,20. hormone receptors. Accordingly, hormones regulate

78 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

the proliferation and survival of both lymphoid and enables their action in the (also known as thymic microenvironment cells, as well as selection of the adenohypophysis) gland40. Hormones secreted by the the T-cell repertoire, which is a mechanism partly related adenopituitary gland can directly stimulate other endo- to the modulation of TCR activation upon MHC-peptide crine targets to release hormones and affect tissue growth, binding39. Many different hormones also regulate the as observed for many aspects of thymic physiology (FIG. 2). migration and export of developing T cells (TABLES 1,2)8. The hypothalamic hormones that affect the thymus GH–IGF‑1 are peptides derived from specialized neurons in specific Isolated ex vivo human thymocytes and primary TEC hypothalamic nuclei, where axonal processes release the cultures can produce and secrete GH41. The expression of hormonal content in a specialized vascular network that GH receptor was initially reported in cultured human

a

Cortex High proliferation Positive selection

Cortico- medullary junction

Negative Very low selection proliferation Medulla

b Cortical TEC TCR–MHC complex 2 3 Medullary TEC Notch 1 Macrophage Delta-like protein 4

Dendritic cell Integrin-type membrane receptor

DN Laminin, fibronectin

4 DP Chemokines

SP CD4+ G protein-coupled receptor

SP CD8+ , peptidic hormone 5 Blood vessel Cytokine/hormone receptor

Figure 1 | Intrathymic T‑cell differentiation. a | T‑cell progenitors enter the thymus via post-capillaryNature Reviews venules | Endocrinology in the cortico-medullary region of the thymus lobules and pass through distinct microenvironment regions: the outer cortex, followed by the inner cortex and finally the medulla. The thymocytes interact with cells in the thymus microenvironment such as cortical TECs, medullary TECs, dendritic cells and macrophages. Developing T cells also regulate the expression levels of different proteins, such as the TCR heterodimer, the CD3 complex, CD4, CD8, CD24, CD25, CD44, CD69 and CD62L, which can be used as markers to define a given stage of development within the organ. b | Cells of the thymus microenvironment interact via multiple mechanisms including: via the TCR and MHC/endogenous peptides (1), direct cell–cell interaction such as Notch or Delta-like protein 4 (2), the extracellular matrix, such as laminins and fibronectin, which bridge thymocytes to TEC via integrin-type membrane receptors (3), chemokines and other soluble moieties that activate G protein-coupled receptors (4) and via cytokines and peptidic hormones, such as IL‑1, IL‑7, growth hormone and prolactin (5). Abbreviations: DN, double negative; DP, double positive; MHC, major histocompatibility complex; TEC, thymus epithelial cell; TCR, T‑cell receptor; SP, single positive.

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 79

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

structure of the thymus in these patients has not been evaluated by noninvasive methods. However, in one case reported in the literature, a patient with acro- megaly who had high levels of GH and insulin-like growth factor 1 (IGF‑1) presented with thymic hyper- Higher brain centres plasia, as ascertained by radiological examination and CT scanning50. Hypothalamus GH also enhances the deposition of proteins involved in cell migration, such as laminin species and stromal Progonadoliberin-1 cell-derived factor 151,52. The migration of thymocytes CRH GHRH derived from GH transgenic mice or from mice injected TRH Anterior pituitary with GH directly into the thymus, is enhanced towards sources of stromal cell-derived factor 1 and laminins51,52. Interestingly, the number of cells derived from the thy- mus are also increased in peripheral lymphoid organs in both of these mouse models52. This finding suggests GH TSH FSH PRL LH ACTH APV that GH induces changes in the repertoire of both thymic OT and peripheral T cells, although no experimental data Spinal DRG cord yet supports this assertion. A large proportion of GH effects in the body are mediated by the production and release of IGF‑1. Thymic Nerve cells can produce and release IGF‑1, and also express the Thyroid Adrenals endings corre­sponding receptor, IGF1R53. Moreover, the enhan­ T E GC 3 2 cing effects of GH upon thymulin production, extra­ T Prog Epi ACh NE 4 cellular matrix expression and adhesion of developing Thymus thymo­cytes to TECs can be abrogated by treating cells 41,53 Cortex with GH, IGF‑1 and IGF1R antibodies . Although the is considered to be the major source of peripheral , this hormone is Cortico- also expressed in immune cells and regulates T‑cell medullary 54 junction activation and inflammation . Ghrelin promotes the release of GH by acting on a specific G protein- Medulla coupled receptor: the GH secretagogue receptor type 1 (GHS‑R)55. GHS‑R mRNA and the protein have also been shown to be present on human and rodent T cells; Figure 2 | Immuno–neuro–endocrine interactions in theNature thymus. Reviews The | Endocrinology thymus can acylated and des-acyl ghrelin are also produced and be influenced by both peptide and nonpeptide hormones. The neuroendocrine secreted upon T‑cell activation54. GHS‑R is expressed control of the thymus includes sympathetic and parasympathetic innervation of the in the thymus and the various thymocyte subsets, with organ, with local release of neurotransmitters such as NE and ACh. Abbreviations: ACh, acetylcholine; ACTH, adrenocorticopropic hormone; AVP, arginine vasopressin; the highest expression in DP cells, and expression is also decreased during ageing56. Functionally, infusion of ghre- CRH, corticotropin-releasing hormone; DRG, dorsal root ganglion; E2, ; Epi, epinephrine; GC, glucocorticoid; GH, growth hormone; GHRH, GH-releasing lin into aged mice led to recovery of the age-associated hormone; LH, ; NE, ; OT, oxytocin; PRL, prolactin; changes in thymic architecture and increased numbers of Prog, ; TRH, thyrotropin-releasing hormone. early thymocyte progenitors, which resulted in augmen- tation of thymo­cyte numbers as well as recent thymic emigrants in the periphery56. Receptors for GH, prolactin and belong to TECs42 and confirmed with the co-localization with the class I cytokine receptor superfamily, and are com- cyto­keratin labelling in avian thymus43. Expression posed of an extracellular ligand binding domain, a of GH receptor is particularly evident in immature trans­membrane domain and an intracellular domain57. thymo­cytes, as seen in both mice and humans41,44. These receptors lack intrinsic kinase activity and trans­ Several functions in the thymus are controlled duce signals through kinases that interact with its cyto­ by GH45. Transgenic mice that overexpress GH have plasmic tail. The main signalling pathway activated by an enlarged thymus, and similar effects have been these receptors is the Janus kinase/signal transducer observed in mice and humans treated with recombinant and activator of transcription (STAT) pathway (FIG. 3). forms of the hormone46–48. GH can also modulate the Binding of the ligand induces the dimerization of the thymic micro­environment by increasing the secretion receptor and activation of Janus kinase 2 (JAK2), which of cytokines, chemokines and thymulin41. Interestingly, in turn phosphorylates multiple tyrosine residues, patients with acromegaly have high serum levels of thy- enabling the binding of STAT1, STAT3 and STAT558. mulin compared with age-matched healthy individu- Tyrosine-phosphorylated STAT5 dissociates from the als or patients with acromegaly who have been treated receptor, dimerizes and translocates to the nucleus, by pituitary surgery or octreotide49. Nevertheless, the where it binds to the promoters of target genes.

80 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

Table 1 | Intrathymic expression of hormones and correspondent receptors the immunosupressive effects of gluco­corticoids on the thymus under stress conditions, as demonstrated by Hormone Site of production Receptor expression Study different in vivo models in which an increase in circu- Thymocytes TECs Thymocytes TECs lating levels of prolactin protected thymocytes­ from Hypothalamus/neurohypophysis glucocorticoid-induced apoptosis60,70. TRH + ND + ND 88,90 Progonadoliberin‑1 + ND + ND 99,100 Thymus physiology is also influenced by neuropeptides CRH + ND ND ND 121 such as oxytocin and vasopressin, which are released AVP ND + + ND 12,72,73,164 by the posterior pituitary lobe (known as the neuro­ 71,72 OT ND + + ND 12,72,73,164 hypophysis) in the brain . TECs produce both vaso­ pressin and oxytocin, but no vasopressin production has Adenohypophysis been detected in thymocytes12. The G protein-coupled GH + + + + 41–44 oxytocin receptor has been detected in all thymocyte TSH ND + + ND 86,87 subsets, whereas the vasopressin V1b receptor was only found in DP and CD8+ SP cells12,72,73. Functionally, inhib­ Prolactin + + + + 44,59–61 ition of the oxytocin receptor in fetal thymus organ cul- ACTH ND + + + 118–120,165 tures increases the amount of early apoptosis of CD8+ Other organs mature T cells, while vasopressin V1b receptor antag­ Ghrelin + + + + 54,56,148 onists, such as nelivaptan, inhibited T‑cell differenti­ ation and favoured the development of CD8+ T cells72,74. Leptin ND + + + 75,77,166 However, much work still has to be done to elucidate the

T3 ND ND + + 84,167 intrathymic role of these neuropeptides. Glucocorticoids + + + + 106,107,111, 112,116 Leptin Androgens ND ND + + 93,94 Leptin is expressed in the thymus and functions via the leptin receptor75. The db/db mouse line, which ND ND + + 92 lacks the , develops type 2 diabetes melli­ + indicates that the hormone or its receptor has been detected at these sites. Abbreviations: ACTH, adrenocorticotropic hormone; AVP, arginine vasopressin; CRH, corticotropin- tus and obesity. In our own research, we have shown that releasing hormone; GH, growth hormone; ND, not determined; OT, oxytocin; db/db mice also undergo a precocious thymic involu-

T3, ; TEC, thymic epithelial cell; TRH, thyrotropin-releasing hormone. tion, with severe loss of thymocytes and accumulation of ­adipocytes within the organ76. Expression of the leptin receptor in the thymus seems Prolactin to be restricted to the microenvironmental cells and its Developing thymocytes constitutively produce prol- activation by leptin protects against loss of lymphoid and actin59 and mTECs seem to be the main source of this TEC populations during stress-induced acute atrophy protein60. We have also shown that of the thymus77. However, no effect on thymopoiesis is is expressed on mTECs and their activation leads to the observed with the administration of leptin to healthy proliferation of this cellular subset61. Moreover, most mice78. Accordingly, intrathymic leptin mRNA is selec- thymic dendritic cells express the PRLR gene (which tively increased in microenvironmental cells in the encodes the pro­lactin receptor), and when these cells young rats from protein-deprived dams, which in turn are treated with prolactin they increase responsiveness protects thymocytes from apoptosis75. The function of in allogeneic mixed leukocyte reactions62. Such an effect both leptin and ghrelin support thymopoiesis. In addi- is probably related to the upregulation of MHC surface tion to changing the TCR repertoire, both hormones expression and the co-stimulatory molecule CD8063. partially reverse age-associated­ thymic involution, albeit PRLR is also expressed by thymocytes, and this expres- by distinct mechanisms56,79. sion is independent of the thymocyte maturation stage and prolactin receptor expression increases in response Thyroid hormone to stimulation with mitogenic factors such as con­ Several lines of evidence suggest that circuitry associ- cavalin A64,65, as well as high levels of prolactin66. Despite ated with thyroid hormone can modulate the function normal development of the immune system in mice defi- of the thymus. For instance, patients with hyperthyroid- cient in the production of prolactin67,68 or expression of ism have increased numbers of thymocytes, which leads PRLR69, prolactin has been shown to be an important fac- to thymic hyperplasia in these individuals80. Consistent

tor for both survival and proliferation of early T‑cell pre- with this finding, mice exogenously injected with T3 cursors such as CD25+CD4–CD8– DN cells63. Accordingly, exhibit an increase in the volume of the thymus, its cel- monoclonal antibodies against prolactin and the prolactin lularity and the cycling of thymocytes81. Moreover, sys-

receptor block T‑cell development, which leads to accu- temic treatment or intrathymic injection of T3 enhances mulation of DN cells in the thymus63. These data indicate thymocyte adhesion and migration towards extracellular that prolactin contributes to the physiological modula- matrix mol­ecules82,83. Such biological effects occur via

tion of thymus function, but is not essential for thymus activation of T3 nuclear receptors, which are expressed development. Furthermore, prolactin seems to counteract in both developing thymocytes and TECs84.

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 81

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

Although the intrathymic production of thyroid hor- of peripheral T cells is not associated with enhanced mone has not been reported, immunocytochemical data thymopoiesis, as observed after orchidectomy98. In this indicate that TSH can be produced by TECs, as detected study, the effect seems to be due to an expansion of the by immunohistochemistry with β-TSH anti­bodies in numbers of pre-existing­ cells in the periphery98. the subcapsular and cortical thymic zones85. Moreover, Although production of sex steroids has not been TSHR (which encodes , the recep- seen in the thymus, progonadoliberin‑1 (also known tor for TSH) is differentially expressed during human as luteinizing hormone-releasing hormone), which T‑cell development in the thymus, and mice lacking stimulates the hypothalamus–pituitary–gonad axis, is functional thyrotropin receptor expression have lower known to be expressed by thymocytes in the thymus numbers of DP and SP thymocytes than wild-type mice, of rats99. In rats, progonadoliberin‑1 can also stimulate which suggests that TSH functions as a growth factor for the proliferation of these cells through direct action on developing T cells86. Accordingly, deregulation of TSHR- ­progonadoliberin‑1‑specific receptors100. mediated regulatory activity of gene expression, through the inhibition of its transcriptional repressor promyelo- Glucocorticoids cytic leukemia zinc finger protein in the thymus, triggers Immature DP thymocytes are major targets of gluco­ thyroid auto­immunity due to the escape of TSHR- corticoid-associated immunosuppression in the thy- reactive T cells from negative selection87. Additionally, mus, and this population of cells undergoes high levels TRH, which encodes prothyrotropin-releasing hormone, of apoptosis and decreased proliferation upon gluco­ is also transcribed in the thymus of the rat88. Given the corticoid stimulation101,102. However, the distinct effects local production of prothyrotropin-releasing hormone, of glucocorticoids on thymocytes seem to be dose the intrathymic expression of the thyrotropin receptor dependent. Although high glucocorticoid levels are and the fact that this is able to enhance required for the induction of apoptosis in these cells, thymocyte prolif­eration89,90, an autocrine/paracrine the exposure of thymo­cytes to low glucocorticoid levels circuit mediated by thyrotropin receptor might control rescues these cells from TCR-mediated apoptosis103,104. the thymus physiol­ogy. However, whether a complete Moreover, a local circuitry mediated by TEC-produced

thyrotropin-releasing hormone–TSH–T3/T4 cascade glucocorticoid has been suggested as the main medi­ator occurs in the thymus remains to be defined. of a glucocorticoid-positive role in thymocyte selec- tion39, decreasing the affinity with which TCR binds Sex hormones to MHC-presented self-ligands and rescuing thymo­ Receptors for androgens and have been found cyte cells from a negative selection route. Although in both thymocytes and microenvironmental cells8,91,92. the absence of gluco­corticoid receptor signalling has Developing androgen receptors are mainly found in no impact on thymo­cyte development105, the deletion the DN and CD8+ SP subsets93,94. Estrogen receptor of glucocorti­coid receptors before selection leads to a signalling is required for the normal development of decreased thymus size. This effect is associated with the thymus in mice95,96. This hormone also prevents the increased antigen-specific negative selection of DP development of human thymoma through the inhib­ thymo­cytes and alterations in the TCR ­repertoire of ition of TEC proliferation, as ascertained by the addition polyclonal T cells103. of estrogen and anti-estrogens to primary cultures of Glucocorticoid receptors are expressed in both the human thymoma epithelial cells97. Although both estro- cytoplasm and nucleus of thymocytes106,107. On bind- gen and androgens have inhibitory effects on thymus ing, glucocorticoid–glucocorticoid receptors com- growth, in ovariectomized­ mice the increase in numbers plexes translocate to the cell nucleus to modulate the

Table 2 | Pleiotropic effects of hormones on the thymus Parameters Hormones Study

AVP OT GH Ghrelin Prolactin Leptin T3 GC Androgens

Size of the organ ND ND ↑ ↑ ↑ ↑­ ↑ ↓ ↓ 81,127,168–171

Cellularity ND ND ↑ ↑­ ↑ ↑­ ↑­ ↑ or ↓ ↓ 45,60,78,172,81,140 Thymocyte proliferation ND ND ↑ ↑ ↑ ↑ ↑ ↓ ↓ 77,79,140,173,174

Thymocyte differentiation ↑­ ↔ ↑­ ↑ ↑ ↑ ↑­ ↑ or ↓ ↓ 45,51,52,82,171,175 Thymocyte death ↔ ↓ ↔ ND ↓ ↓ ↑ ↑ or ↓ ↓ 70,72,74,77,176–178 Intrathymic T‑cell repertoire ND ND ↑ ↑ ↓ ↑ ND ↑ ↑ 39,56,179,180 TEC proliferation ND ND ↑ ↑ ↑ ↑ ND ↓ ↓ 36,61,181,182 TEC death ND ND ND ND ND ↓ ND ↑ ND 77,183 Thymic hormone production ND ND ↑ ND ↑ ND ND ↑ ↑ 170,184

↑, increase; ↓, decrease; ↔, no effect. Abbreviations: AVP, arginine vasopressin; GC, glucocorticoid; GH, growth hormone; ND, not determined; OT, oxytocin; TEC, thymic epithelial cell.

82 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

a b Control GH c GH

JAK2 5 Cytoplasm JAK2 P 4 * P STAT5 STAT5 STAT5 ) 5 3 STAT5 P P STAT5 2 CDK2 TEC numbers (×10

Gene expression STAT5 c-MYC, CDK 1 P P growth factors, STAT5 cytokines CRE CDK4 0 Control GH Nucleus Treatment

Figure 3 | GH signalling in TECs stimulates proliferation of T cells. a | GH binding to itsNature receptor Reviews recruits | Endocrinology and activates the receptor-associated JAK2 that in turn phosphorylates tyrosine residues within itself and the GH receptor. These tyrosines form binding sites for a number of signalling proteins, including members of the STAT family. Among the known signalling molecules for GH, STAT5 proteins have a particularly prominent role in the regulation of gene transcription. In the nucleus, phosphorylated STAT5 binds to the promoters of target genes, such as theCRE . b | A mouse TEC line treated with GH has increased expression of JAK2, STAT5, CDK2 and CDK4 (as determined by immunohistochem- istry). c | The overall numbers of TECs are also increased when treated with GH. Abbreviations: CDK, cyclin-dependent kinase; CRE, cAMP response element; GH, growth hormone; JAK2, Janus kinase 2; P, phosphorylated; STAT5, signal transducers and activators of transcription 5; TEC, thymus epithelial cell. Panel c modified with permission from Wiley © Savino, W. et al. Scand. J. Immunol. 55, 442–452 (2002)138.

expression of target genes (FIG. 4)108–110. Correlating POMC (the gene that encodes for proopiomelano­ with the increased sensitivity to the effects of gluco- cortin, which is cleaved into adrencorticotropic hor- corticoids, the density of glucocorticoid receptors mone) also seems to be constitutively expressed in is increased in immature thymocytes in the human TECs and TDCs118,119. The adrencorticotropic hor- thymus compared with mature thymocytes111. mone receptor, receptor subtype 2, is In our own research, we have demonstrated that expressed on TECs and its activation regulates thymo­ increased exposure to circulating glucocorticoid cyte expansion when the systemic concentration of decreases glucocorticoid receptor‑α expression in the glucocorticoid is low120. The intrathymic production DP cell subset in mice60. Glucocorticoid receptors have of corticotropin-releasing factor by thymocytes has also been detected in TEC cultures112, and the β isoform also been reported121. Taken together, these findings of this receptor, which has been reported to have a dom- suggest that a circuitry, similar to that seen in the inant negative effect on glucocorticoid α‑induced trans- hypothalamus–­pituitary–adrenal axis, might also activation of glucocorticoid response element-driven occur within the thymus, but this system needs to be promoters113, was detected in situ in medullary TECs114. investigated further. In addition to expressing glucocorticoid receptors, both TECs and developing thymocytes express all the Infection and T‑cell development enzymes and cofactors required for the production Atrophy of the thymus is a common feature in acute of glucocorticoids115. Corticosterone is produced by infections of viruses, bacteria, parasites or fungi122. Such TECs, whereas the synthesis of glucocorticoid by thy- atrophy is frequently associated with hypothalamus– mocytes was also elegantly demonstrated in a cell line pituitary–adrenal dysfunction, and can be exem- transfected with glucocorticoid receptors and coincu- plified by the changes seen in patients with Chagas bated with thymocytes116. Specifically, DP cells seem disease, which is caused by the protozoan parasite to be the main thymocyte subset that produces gluco- Trypanosoma cruzi and is a major public health issue corticoids117. Thymocyte-derived glucocorticoids have in Latin America. We have previously shown that the an anti-­proliferative effect on the cell line and induced atrophy seen in mice who have acute T. cruzi infection apoptosis in thymocytes117. is characterized by a loss of DP thymocytes expressing

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 83

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

a b GC c GC PRL PRL

GC GC

GC GC JAK2 Trypanosoma cruzi GC GC P infection Pituitary P ACTH STAT5 STAT5

STAT5 Adrenal ↓ PRL GC GC P STAT5 P

↑ GC

Gene Gene GC GC STAT5 expression expression P P ↑ PRL Pro-apoptotic Pro-apoptotic ↓ GC STAT5 Thymus genes GRE genes GRE

Increased apoptosis Partial blockade of apoptosis

Figure 4 | Trypanosoma cruzi infection and thymus homeostasis. a | T. cruzi infection resultsNature in Reviews an increased | Endocrinology systemic production of glucocorticoids by the without altering levels of adrenocorticotropic hormone. At the same time, the infection decreases PRL synthesis and/or release by the pituitary. T. cruzi infection also decreases the intrathymic contents of glucocorticoids and increases levels of locally produced PRL. b | Immature DP thymocytes sensitive to T. cruzi infection-induced apoptosis have increased levels of GR (blue rectangle). Upon GC binding, the GR complex translocates to the nucleus where it drives the transcription of pro-apoptotic proteins from the GRE. c | Increased PRL-mediated STAT5 signalling (which is secondary to exogenous prolactin administration) has a protective effect duringT. cruzi infection by binding to the GR complex and inhibiting the pro-apoptotic effects of GC. Abbreviations: ACTH, adrenocorticotropic hormone; DP, double positive; GC, glucocorticoid; GR, glucocorticoid receptor; GRE, glucocorticoid response element; JAK2, Janus kinase 2; P, phosphorus; PRL, prolactin; STAT5, signal transducers and activators of transcription 5.

low levels of TCR, together with an abnormal release of Increased serum levels of glucocorticoid was accom- immature thymocytes60,123,124. Some of these thymocytes panied by a decrease in intrathymic levels of cortico­ have TCR Vβ families, which under normal conditions sterone during acute T. cruzi infection, which indicates should have undergone apoptosis, but persist and that the control of intrathymic glucocorticoid produc- might, therefore, lead to an autoimmune reaction­ 125. tion is independent of systemic glucocorticoid levels60. The T. cruzi-induced progressive atrophy of the DP thymocytes from mice infected with T. cruzi also thymus was paralleled by increased circulating levels of had reduced levels of glucocorticoid receptor mRNA, in glucocorticoids126 and could be prevented in mice pre- addition to elevated expression of PRLR60. This finding viously adrenalectomized and treated with the gluco­ suggests that prolactin counteracts the effect of gluco- corticoid receptor antagonist, RU486127. Interestingly, corticoid by directly affecting gluco­corticoid receptor we also found decreased levels of corticoliberin in the signalling in DP cells (FIG. 4b). Accordingly, signalling hypothalamus of T. cruzi-infected mice, with no sig- generated by prolactin receptor, which is mediated by nificant changes in the circulating amounts of adren- the STAT5 pathway, abro­gates glucocorticoid-induced ocorticotropic hormone, but with high levels of IL‑6, a apoptosis in T cells131. Consequently, re-establishing cytokine known to function directly on adrenal cells to systemic prolactin levels by treatment with metyrapone stimulate corticosterone release128,129. Moreover, levels (which stimulates synthesis of adrenocorticotropic hor- of GH and prolactin were diminished in the pituitary mone) can prevent thymic atrophy by decreasing both gland of T. cruzi-infected mice and in vitro infection DP cell apoptosis and the numbers of DP cells in the with the parasite of the rat pituitary tumour cell line, periphery of the immune system60. This finding indi- GH3, as determined using immunohistochemistry130. cates that prolactin-mediated protection of the thymus Prolactin has emerged as another stress-adaptation also influences the abnormal export of these immature molecule with altered production during T. cruzi potentially autoreactive T cells60. infection. Reversed kinetic changes have been found in the serum levels of glucocorticoids and prolactin in Hormone therapy mice undergoing experimental acute Chagas disease60. The thymus undergoes a physiological age-dependent This result suggests that the immunosuppression that involution with changes in both microenvironment and appears after T. cruzi infection is partially related to a lymphoid compartments that usually result in the loss of stress hormone imbalance (FIG. 4a). T‑cell export to the periphery of the immune system34,132.

84 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

Although the overall numbers of T cells are not dimin- However, treating ageing rodents with either GH or ished in the periphery of individuals as they age, naive IGF‑1 did not restore thymus cellularity to the levels T cells are replaced by memory T cells in both humans seen in healthy young mice146. As the combined admin- and mice133. Furthermore, the ageing process results in istration of IGF‑1 and young bone marrow cells resulted suppression of T-cell proliferation, cytokine production in a higher cellularity of thymus in these mice than in and CD8‑mediated cytotoxicity, which contributes to a animals treated with GH or IGF‑1, or bone marrow reduced response of T cells134. transplantation alone, ageing-related additional defects An accelerated involution of the organ can be seen in the bone marrow might limit the magnitude of the in pathological situations, such as malnutrition, and rejuvenating effects of these hormones during ageing. acute infections such as those caused by T. cruzi and The expression of ghrelin and its receptor is also Paracoccidioides brasiliensis, as well as some immuno­ decreased in ageing mice56 and deletion of Ghrl accel- deficiency disorders such as HIV infection135,136. erates thymic involution, fibroadipogenetic trans- Accordingly, replenishing the thymus has been pursued formation and adipogenesis in the mouse thymus148. as a potential t­herapeutic strategy to prevent involution Importantly, re-establishing the presence of ghrelin in of the organ. the serum partially reverses the thymic involution pro- cess, increases the numbers of thymic emigrants and Age-dependent thymus involution improves TCR diversity of peripheral T cells in ageing The serum concentrations of both GH and IGF‑1 mice in a mechanism independent of IGF‑1 induc- decline during ageing137. GH increases TEC pro- tion56. However, further studies are needed to elucidate liferation in vitro (FIG. 3b) and synergizes with if the disruption of a ghrelin–GH–IGF‑1 axis underlies anti‑CD3 in its stimulatory effect on thymocyte pro- the effects of ageing on the thymus. liferation138. Accordingly, GH is a potential adjuvant Treating 14‑month-old mice with leptin can also therapeutic agent to revert age-dependent thymic lead to the recovery of thymopoiesis, which is charac- involution. Positive effects on the growth of the thy- terized by an increase in thymus size and weight, the mus in 24‑month-old rats implanted with cells from number of thymocytes, TECs and thymic emigrants56. a pituitary tumour producing both GH and prolac- In addition, during states of malnutrition, the balance tin have been reported139. Conversely, GH receptor between the levels of leptin and glucocorticoid is dis- knockout mice have a severe precocious decrease of rupted. This situation leads to a decrease in levels of thymulin production, indicating a functional defect leptin and consequent enhancement of glucocorticoid- of the thymic epithelial network140. Although studies mediated thymo­cyte apoptosis with atrophy of the thy- on thymus function have not been reported in chil- mus; exogenous administration of leptin might reverse dren with typical Laron syndrome (who have muta- these effects149. tions in the gene that encodes GH receptor), they do Atrophy of the thymus accelerates at puberty95, exhibit major immune dysfunctions141,142. Interestingly, which suggests that increasing levels of sex steroids leads in mutations in STAT5b (which encodes a component to this effect. Accordingly, exogenous­ administration of of a GH signalling pathway), growth defects correlates sex steroids, such as dihydro­ and estradiol, with immunodeficiency141,142. in adult mice causes a collapse of thymo­poiesis that is GH has the potential to improve thymic function, associated with increased apoptosis of cortical thymo- enhance bone marrow engraftment and stimulate cytes150. In addition, ablation of the production of sex haematopoiesis in immunosuppressed and aged ani- steroids by surgical removal of the gonads or chemi- mals143,144. Moreover, in ageing mice, GH treatment cal blockage (by interfering with the hypothalamus- enhanced the diversity of the peripheral T‑cell reper- derived progonadoliberin‑1) can ­rejuvenate the thymus toire (as determined by PCR analysis of CDR3 length in rodents94,151,152. for several TCR Vβ genes)45. In addition to these find- Progonadoliberin‑1 signalling depends on the ings in mice45,52,145, GH administration in adult humans activation of its receptor, which is a member of the who have GH deficiency enables the recovery of the G protein-coupled receptor family. The continuous numbers of thymic emigrant cells in the periphery46. exposure to the cognate ligand leads to downregula- These results clearly indicate that GH is important for tion of this receptor and consequently a decrease in the maintenance of T-cell output and thymocyte pro- the release of sex hormones by the gonads153. Studies liferation and that GH might therefore be useful for conducted in rat thymocytes revealed a decrease in hormone replacement therapy46. binding sites for progonadoliberin­ ‑­1 with age; the abil- Many of the effects of GH administration are medi- ity of ­progonadoliberin‑1 to modulate thymus function ated by IGF‑1. Indeed, administration of IGF‑1 in vivo indicates that progonadoliberin‑1 action has a direct can partially reverse the age-associated mouse thymic effect on this organ100. Similar to the effect of testoster- involution and enhance thymopoiesis146. Furthermore, one removal by castration, the administration of pro- in mice lacking the IGF‑1 receptor on thymocytes gonadoliberin‑1 agonists restores thymus weight and and T cells, development proceeds normally upon increases progonadoliberin‑1 binding in the thymus IGF‑1 treatment, which seems to be dependent on an of aged rats, resulting in a partial restoration of thymic expansion in the numbers of TECs and their capac- structure and an increased thymocyte proliferation100,152. ity to regulating the entrance of T‑cell precursors into In addition to a decrease in thymocyte apoptosis (which the thymus147. is also high in ageing), a substantial increase in DN, DP

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 85

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

cells and CD4+ SP subsets is observed in mice after cas- circles (TRECs; which are small DNA circles present tration36. Such an effect seems to be partly related to the in T cells as a residue after TCR gene rearrangement) re-establishment of the epithelial compartment36. in circulating T cells, as well as the numbers of naive Sex steroid ablation also increases bone marrow and total CD4+ T cells in the blood of these patients162. recovery in ageing animals152, although whether the thy- These findings provide compelling evidence that GH mus recruits precursor T cells in this scenario is unclear. enhances T-cell production and export and, therefore, Nevertheless, the interaction between Notch on develop- facilitates CD4+ T‑cell recovery in individuals with HIV ing thymocytes and Dll4 on TECs can be modulated by infection. A similar effect was found in another study, in sex hormones, as progonadoliberin‑1 antagonists, which which a daily treatment with low dose of recombinant mimic sex steroid ablation, enhance thymopoiesis in aged human GH over 40 weeks stimulated thymopoiesis and and female mice by upregulation of Dll4 expression154. the frequency of T-cell receptor excision circles in CD4+ Although ovarian estrogen ablation in ageing females T cells from patients with HIV infection undergoing also increases thymus cellularity, regeneration is more highly aggressive anti-retroviral therapy­ 163. pronounced in men than in women, which proba- The improvement of function after removal of hor- bly reflects the increased extragonadal production of mone therapy suggests that this treatment enables the estrogen by adipose tissue and seen after oophor­ entrance of thymocyte progenitors and developing ectomy155,156. Despite the difference in the thymus physi­ thymocytes into the atrophied thymus, as well as TEC ology, progonadoliberin‑1 analogues have been used in growth and T‑cell output. However, despite these prom- the treatment of both prostate and breast cancer that are ising results, studies with larger cohorts of patients are hormonally sensitive157. still required before GH therapy in patients with HIV infection enter into clinical practice. Hormone therapy in HIV HIV infection leads to severe atrophy of the thymus, Conclusions with a substantial decrease in the total number of devel- Hormonal control of thymus physiology is complex oping thymocytes, with thymic microenvironmental and involves both endocrine and paracrine signalling cells also being affected135,158. Accordingly, the restora- pathways. The molecular mechanisms underlying this tion of thymus function would be greatly beneficial for control have been only partly determined, including the patients with HIV infection. GH and IGF‑1 can promote intrathymic expression of respective ligands and recep- the re-establishment of T-cell development in patients tors. The intracellular signalling has also been partially with HIV infection. Exogenously administered GH has defined, but much work remains. Genetic manipulation a thymopoietic role in humans with HIV infection159. of animal models, including conditional knockouts for In this study, the investigators examined the thymus of signalling pathways that target thymic-specific cell pro- patients with HIV‑1 infection who received GH therapy, moters, will also contribute to our deeper understand- and found a marked increase in overall thymic mass and ing of the molecular mechanisms that underpin the in the numbers of CD4+ T cells in the blood159. hormonal control of the thymus. Specifically, the func- In addition to the increased numbers of naive T cells tional rele­vance of the various intrathymic hormonal in GH‑treated patients who have HIV infection, many circuitries will hopefully be defined. of the T cells are derived from expansion of the periph- In addition to physiological mechanisms, dysfunc- eral naive CD4+ T-cell pool160. Whether these increased tion in the hormonal control of the thymus can be numbers of naive T cells were due to the increase in cells induced by endogenous or exogenous stimuli, trigger- that have left the thymus was unclear. These cells might ing or enhan­cing disease. The consequences on the derive from the infiltration of mature peripheral T cells progression or regression of the disease in response to and from adipose tissue within the perivascular spaces hormone therapy are largely unknown, except for some of the thymic lobules161. However, with the use of highly specific cases, such as HIV infection. However, investi- aggressive anti-retroviral therapy and improved control gators are beginning to apply hormonal manipulation of the inflammation in these patients, GH was found to of the thymus as an alternative therapeutic approach for enhance the number of mature thymocytes162, as meas- specific acute infectious diseases as well as noninfectious ured by both the frequency of T-cell receptor excision immunodeficiency conditions,­ including ageing.

1. Savino, W. Intrathymic migration is a 6. Selye, H. Implications of stress concept. 11. Savino, W., Smaniotto, S., Mendes‑da‑Cruz, D. A. & multivectorial process under a complex N. Y. State J. Med. 75, 2139–2145 (1975). Dardenne, M. Growth hormone modulates migration of neuroendocrine control. Neuroimmunomodulation 17, 7. Ader, R., Felton, D. L. & Cohen, H. (Eds) thymocytes and peripheral T cells. Ann. N. Y. Acad. Sci. 142–145 (2010). Psychoneuroimmunology 4th edn (Academic Press, 1261, 49–54 (2012). 2. Besedovsky, H. O. & Sorkin, E. Thymus involvement in 2006). 12. Moll, U. M., Lane, B. L., Robert, F., Geenen, V. & Legros, female sexual maturation. Nature 249, 356–358 8. Savino, W. & Dardenne, M. Neuroendocrine control J. J. The neuroendocrine thymus. Abundant occurrence (1974). of thymus physiology. Endocr. Rev. 21, 412–443 of oxytocin‑, vasopressin‑, and neurophysin-like peptides 3. Besedovsky, H., del Rey, A., Sorkin, E. & Dinarello, (2000). in epithelial cells. Histochemistry 89, 385–390 (1988). C. A. Immunoregulatory feedback between 9. Reggiani, P., Martines, E., Ferese, C., Goya, R. 13. Lind, E. F., Prockop, S. E., Porritt, H. E. & Petrie, H. T. interleukin‑1 and glucocorticoid hormones. Science & Console, G. Morphological restoration of Mapping precursor movement through the postnatal 233, 652–654 (1986). gonadotrope population by thymulin gene therapy in thymus reveals specific microenvironments supporting 4. Blalock, J. E., Smith, E. M. & Meyer, W. J. 3rd. The nude mice. Histol. Histopathol. 24, 729–735 (2009). defined stages of early lymphoid development. J. Exp. pituitary-adrenocortical axis and the immune system. 10. de Mello-Coelho, V., Villa-Verde, D. M., Dardenne, M. Med. 194, 127–134 (2001). Clin. Endocrinol. Metab. 14, 1021–1038 (1985). & Savino, W. Pituitary hormones modulate cell-cell 14. Attaf, M., Huseby, E. & Sewell, A. K. αβ T cell receptors 5. Savino, W. & Dardenne, M. Immune-neuroendocrine interactions between thymocytes and thymic as predictors of health and disease. Cell. Mol. Immunol. interactions. Immunol. Today 16, 318–322 (1995). epithelial cells. J. Neuroimmunol. 76, 39–49 (1997). 12, 391–399 (2015).

86 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

15. Miles, J. J., Douek, D. C. & Price, D. A. Bias in the αβ 40. Charlton, H. Hypothalamic control of anterior pituitary 64. Gagnerault, M. C., Touraine, P., Savino, W., Kelly, P. A. T‑cell repertoire: implications for disease pathogenesis function: a history. J. Neuroendocrinol. 20, 641–646 & Dardenne, M. Expression of prolactin receptors in and vaccination. Immunol. Cell Biol. 89, 375–387 (2008). murine lymphoid cells in normal and autoimmune (2011). 41. de Mello-Coelho, V. et al. Growth hormone and its situations. J. Immunol. 150, 5673–5681 (1993). 16. Godfrey, D. I. & Zlotnik, A. Control points in early T‑cell receptor are expressed in human thymic cells. 65. Dardenne, M., de Moraes Mdo, C., Kelly, P. A. development. Immunol. Today 14, 547–553 (1993). Endocrinology 139, 3837–3842 (1998). & Gagnerault, M. C. Prolactin receptor expression in 17. Fu, G. et al. Fine-tuning T cell receptor signaling to 42. Ban, E. et al. Specific binding sites for growth hormone human hematopoietic tissues analyzed by flow control T cell development. Trends Immunol. 35, in cultured mouse thymic epithelial cells. Life Sci. 48, cytofluorometry. Endocrinology 134, 2108–2114 311–318 (2014). 2141–2148 (1991). (1994). 18. Egawa, T. Regulation of CD4 and CD8 43. Hull, K. L., Thiagarajah, A. & Harvey, S. Cellular 66. Feng, J. C., Loh, T. T. & Sheng, H. P. Lactation coreceptor expression and CD4 versus CD8 lineage localization of growth hormone receptors/binding increases prolactin receptor expression in spleen and decisions. Adv. Immunol. 125, 1–40 (2015). proteins in immune tissues. Cell Tissue Res. 286, thymus of rats. Life Sci. 63, 111–119 (1998). 19. Spidale, N. A., Wang, B. & Tisch, R. Cutting edge: 69–80 (1996). 67. Foster, M., Montecino-Rodriguez, E., Clark, R. Antigen-specific thymocyte feedback regulates 44. Gagnerault, M. C., Postel-Vinay, M. C. & Dardenne, M. & Dorshkind, K. Regulation of B and T cell homeostatic thymic conventional dendritic cell Expression of growth hormone receptors in murine development by anterior pituitary hormones. Cell. maturation. J. Immunol. 193, 21–25 (2014). lymphoid cells analyzed by flow cytofluorometry. Mol. Life Sci. 54, 1076–1082 (1998). 20. Klein, L., Kyewski, B., Allen, P. M. & Hogquist, K. A. Endocrinology 137, 1719–1726 (1996). 68. Horseman, N. D. et al. Defective mammopoiesis, but Positive and negative selection of the T cell repertoire: 45. Taub, D. D., Murphy, W. J. & Longo, D. L. Rejuvenation normal hematopoiesis, in mice with a targeted what thymocytes see (and don’t see). Nat. Rev. of the aging thymus: growth hormone-mediated and disruption of the prolactin gene. EMBO J. 16, Immunol. 14, 377–391 (2014). ghrelin-mediated signaling pathways. Curr. Opin. 6926–6935 (1997). 21. Derbinski, J., Schulte, A., Kyewski, B. & Klein, L. Pharmacol. 10, 408–424 (2010). 69. Bouchard, B., Ormandy, C. J., Di Santo, J. P. Promiscuous gene expression in medullary thymic 46. Morrhaye, G. et al. Impact of growth hormone (GH) & Kelly, P. A. Immune system development and epithelial cells mirrors the peripheral self. Nat. deficiency and GH replacement upon thymus function function in prolactin receptor-deficient mice. Immunol. 2, 1032–1039 (2001). in adult patients. PLoS ONE 4, e5668 (2009). J. Immunol. 163, 576–582 (1999). 22. Laan, M. & Peterson, P. The many faces of Aire 47. Polgreen, L., Steiner, M., Dietz, C. A., Manivel, J. C. & 70. Krishnan, N., Thellin, O., Buckley, D. J., in central tolerance. Front. Immunol. 4, 326 (2013). Petryk, A. Thymic hyperplasia in a child treated with Horseman, N. D. & Buckley, A. R. Prolactin 23. Passos, G. A., Mendes‑da‑Cruz, D. A. & Oliveira, E. H. growth hormone. Growth Horm. IGF Res. 17, 41–46 suppresses glucocorticoid-induced thymocyte The thymic orchestration involving aire, mi­RNAs, and (2007). apoptosis in vivo. Endocrinology 144, 2102–2110 cell-cell interactions during the induction of central 48. Kermani, H. et al. Expression of the growth hormone/ (2003). tolerance. Front. Immunol. 6, 352 (2015). insulin-like growth factor axis during Balb/c thymus 71. Melis, M. R., Mauri, A. & Argiolas, A. Opposite 24. Ciofani, M. & Zuniga-Pflucker, J. C. The thymus as an ontogeny and effects of growth hormone upon ex vivo changes in the content of oxytocin- and vasopressin- inductive site for T lymphopoiesis. Annu. Rev. Cell Dev. T cell differentiation. Neuroimmunomodulation 19, like immunoreactive peptides in the rat thymus during Biol. 23, 463–493 (2007). 137–147 (2012). aging. Regul. Pept. 59, 335–340 (1995). 25. Petrie, H. T. & Zuniga-Pflucker, J. C. Zoned out: 49. Timsit, J. et al. Growth hormone and insulin-like 72. Hansenne, I. et al. Ontogenesis and functional aspects functional mapping of stromal signaling growth factor‑I stimulate hormonal function and of oxytocin and vasopressin gene expression in the microenvironments in the thymus. Annu. Rev. Immunol. proliferation of thymic epithelial cells. J. Clin. thymus network. J. Neuroimmunol. 158, 67–75 25, 649–679 (2007). Endocrinol. Metab. 75, 183–188 (1992). (2005). 26. Savino, W., Mendes‑da‑Cruz, D. A., Silva, J. S., 50. Bazzoni, N. et al. Acromegaly and thymic hyperplasia: 73. Elands, J., Resink, A. & De Kloet, E. R. Dardenne, M. & Cotta‑de‑Almeida, V. Intrathymic T‑cell a case report. J. Endocrinol. Invest. 13, 931–935 Neurohypophyseal hormone receptors in the rat migration: a combinatorial interplay of extracellular (1990). thymus, spleen, and lymphocytes. Endocrinology matrix and chemokines? Trends Immunol. 23, 51. Smaniotto, S. et al. Combined role of extracellular 126, 2703–2710 (1990). 305–313 (2002). matrix and chemokines on peripheral lymphocyte 74. Hansenne, I. et al. Neurohypophysial receptor gene 27. Cotta‑de‑Almeida, V. et al. Trypanosoma cruzi infection migration in growth hormone transgenic mice. Brain expression by thymic T cell subsets and thymic T cell modulates intrathymic contents of extracellular matrix Behav. Immun. 24, 451–461 (2010). lymphoma cell lines. Clin. Dev. Immunol. 11, 45–51 ligands and receptors and alters thymocyte migration. 52. Smaniotto, S. et al. Growth hormone modulates (2004). Eur. J. Immunol. 33, 2439–2448 (2003). thymocyte development in vivo through a combined 75. da Silva, S. V. et al. Increased leptin response and 28. Savino, W., Mendes‑Da‑Cruz, D. A., Smaniotto, S., action of laminin and CXC chemokine ligand 12. inhibition of apoptosis in thymocytes of young rats Silva-Monteiro, E. & Villa-Verde, D. M. Molecular Endocrinology 146, 3005–3017 (2005). offspring from protein deprived dams during mechanisms governing thymocyte migration: combined 53. de Mello Coelho, V. et al. Functional insulin-like lactation. PLoS ONE 8, e64220 (2013). role of chemokines and extracellular matrix. J. Leukoc. growth factor‑1/insulin-like growth factor‑1 receptor- 76. Dardenne, M., Savino, W., Gastinel, L. N., Nabarra, B. Biol. 75, 951–961 (2004). mediated circuit in human and murine thymic & Bach, J. F. Thymic dysfunction in the mutant 29. Michie, A. M. et al. Constitutive Notch signalling epithelial cells. Neuroendocrinology 75, 139–150 diabetic (db/db) mouse. J. Immunol. 130, promotes CD4 CD8 thymocyte differentiation in the (2002). 1195–1199 (1983). absence of the pre-TCR complex, by mimicking pre-TCR 54. Dixit, V. D. et al. Ghrelin inhibits leptin- and activation- 77. Gruver, A. L., Ventevogel, M. S. & Sempowski, G. D. signals. Int. Immunol. 19, 1421–1430 (2007). induced proinflammatory cytokine expression by Leptin receptor is expressed in thymus medulla and 30. Sambandam, A. et al. Notch signaling controls the human monocytes and T cells. J. Clin. Invest. 114, leptin protects against thymic remodeling during generation and differentiation of early T lineage 57–66 (2004). endotoxemia-induced thymus involution. progenitors. Nat. Immunol. 6, 663–670 (2005). 55. Howard, A. D. et al. A receptor in pituitary and J. Endocrinol. 203, 75–85 (2009). 31. Varas, A. et al. Age-dependent changes in thymic hypothalamus that functions in growth hormone 78. Hick, R. W., Gruver, A. L., Ventevogel, M. S., macrophages and dendritic cells. Microsc. Res. Tech. release. Science 273, 974–977 (1996). Haynes, B. F. & Sempowski, G. D. Leptin selectively 62, 501–507 (2003). 56. Dixit, V. D. et al. Ghrelin promotes thymopoiesis augments thymopoiesis in leptin deficiency and 32. Aw, D., Silva, A. B., Maddick, M., von Zglinicki, T. & during aging. J. Clin. Invest. 117, 2778–2790 (2007). lipopolysaccharide-induced thymic atrophy. Palmer, D. B. Architectural changes in the thymus of 57. Bole-Feysot, C., Goffin, V., Edery, M., Binart, N. J. Immunol. 177, 169–176 (2006). aging mice. Aging Cell 7, 158–167 (2008). & Kelly, P. A. Prolactin (PRL) and its receptor: actions, 79. Lee, J. H. et al. Ghrelin augments murine 33. Aw, D., Taylor-Brown, F., Cooper, K. & Palmer, D. B. signal transduction pathways and phenotypes T‑cell proliferation by activation of the Phenotypical and morphological changes in the thymic observed in PRL receptor knockout mice. Endocr. Rev. phosphatidylinositol‑3‑kinase, extracellular signal- microenvironment from ageing mice. Biogerontology 19, 225–268 (1998). regulated kinase and protein kinase C signaling 10, 311–322 (2009). 58. DaSilva, L. et al. Prolactin recruits STAT1, STAT3 and pathways. FEBS Lett. 588, 4708–4719 (2014). 34. Lepletier, A., Chidgey, A. P. & Savino, W. Perspectives STAT5 independent of conserved receptor tyrosines 80. Chen, Y. K. et al. The frequency and spectrum of for improvement of the thymic microenvironment TYR402, TYR479, TYR515 and TYR580. Mol. Cell thymus 2‑[fluorine‑18] fluoro‑2‑deoxy‑D‑glucose through manipulation of thymic epithelial cells: a mini- Endocrinol. 117, 131–140 (1996). uptake patterns in hyperthyroidism patients. Acad. review. Gerontology http://dx.doi.org/10.1159/ 59. Montgomery, D. W. et al. Human thymocytes express Radiol. 18, 1292–1297 (2011). 000375160. a prolactin-like messenger ribonucleic acid and 81. Villa-Verde, D. M., de Mello-Coelho, V., 35. Qi, Q. et al. Diversity and clonal selection in the human synthesize bioactive prolactin-like proteins. Farias‑de‑Oliveira, D. A., Dardenne, M. & Savino, W. T‑cell repertoire. Proc. Natl Acad. Sci. USA 111, Endocrinology 131, 3019–3026 (1992). Pleiotropic influence of triiodothyronine on thymus 13139–13144 (2014). 60. Lepletier, A. et al. Trypanosoma cruzi disrupts thymic physiology. Endocrinology 133, 867–875 (1993). 36. Gray, D. H. et al. Developmental kinetics, turnover, and homeostasis by altering intrathymic and systemic 82. Ribeiro-Carvalho, M. M., Farias‑de‑Oliveira, D. A., stimulatory capacity of thymic epithelial cells. Blood stress-related endocrine circuitries. PLoS Negl. Trop. Villa-Verde, D. M. & Savino, W. Triiodothyronine 108, 3777–3785 (2006). Dis. 7, e2470 (2013). modulates extracellular matrix-mediated interactions 37. Griffith, A. V., Fallahi, M., Venables, T. & Petrie, H. T. 61. Dardenne, M., Kelly, P. A., Bach, J. F. & Savino, W. between thymocytes and thymic microenvironmental Persistent degenerative changes in thymic organ Identification and functional activity of prolactin cells. Neuroimmunomodulation 10, 142–152 (2002). function revealed by an inducible model of organ receptors in thymic epithelial cells. Proc. Natl Acad. 83. Ribeiro-Carvalho, M. M. et al. Triiodothyronine regrowth. Aging Cell 11, 169–177 (2012). Sci. USA 88, 9700–9704 (1991). modulates thymocyte migration. Scand. J. Immunol. 38. Ventevogel, M. S. & Sempowski, G. D. Thymic 62. Carreno, P. C., Jimenez, E., Sacedon, R., Vicente, A. & 66, 17–25 (2007). rejuvenation and aging. Curr. Opin. Immunol. 25, Zapata, A. G. Prolactin stimulates maturation and 84. Villa-Verde, D. M. et al. Identification of nuclear 516–522 (2013). function of rat thymic dendritic cells. J. Neuroimmunol. triiodothyronine receptors in the thymic epithelium. 39. Vacchio, M. S., Lee, J. Y. & Ashwell, J. D. Thymus- 153, 83–90 (2004). Endocrinology 131, 1313–1320 (1992). derived glucocorticoids set the thresholds for 63. Carreno, P. C., Sacedon, R., Jimenez, E., Vicente, A. & 85. Batanero, E. et al. The neural and neuro-endocrine thymocyte selection by inhibiting TCR-mediated Zapata, A. G. Prolactin affects both survival and component of the human thymus. II. Hormone thymocyte activation. J. Immunol. 163, 1327–1333 differentiation of T‑cell progenitors. J. Neuroimmunol. immunoreactivity. Brain Behav. Immun. 6, 249–264 (1999). 160, 135–145 (2005). (1992).

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 87

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

86. van der Weerd, K. et al. Thyrotropin acts as a T‑cell 109. Carlberg, C. & Seuter, S. Dynamics of nuclear 131. Lechner, J., Welte, T. & Doppler, W. Mechanism of developmental factor in mice and humans. Thyroid receptor target gene regulation. Chromosoma 119, interaction between the glucocorticoid receptor and 24, 1051–1061 (2014). 479–484 (2010). Stat5: role of DNA-binding. Immunobiology 198, 87. Stefan, M. et al. Genetic-epigenetic dysregulation of 110. Stavreva, D. A. et al. Ultradian hormone stimulation 112–123 (1997). thymic TSH receptor gene expression triggers induces glucocorticoid receptor-mediated pulses of 132. Gruver, A. L., Hudson, L. L. & Sempowski, G. D. thyroid autoimmunity. Proc. Natl Acad. Sci. USA gene transcription. Nat. Cell Biol. 11, 1093–1102 Immunosenescence of ageing. J. Pathol. 211, 111, 12562–12567 (2014). (2009). 144–156 (2007). 88. Montagne, J. J., Ladram, A., Nicolas, P. & Bulant, M. 111. Ranelletti, F. O. et al. Glucocorticoid receptors and 133. Kovaiou, R. D. et al. Age-related differences in Cloning of thyrotropin-releasing hormone precursor corticosensitivity of human thymocytes at discrete phenotype and function of CD4+ T cells are due to a and receptor in rat thymus, adrenal gland, and stages of intrathymic differentiation. J. Immunol. phenotypic shift from naive to memory effector CD4+ testis. Endocrinology 140, 1054–1059 (1999). 138, 440–445 (1987). T cells. Int. Immunol. 17, 1359–1366 (2005). 89. Matre, V. et al. The human neuroendocrine 112. Dardenne, M., Itoh, T. & Homo-Delarche, F. Presence 134. Chiu, B. C., Martin, B. E., Stolberg, V. R. thyrotropin-releasing hormone receptor promoter is of glucocorticoid receptors in cultured thymic & Chensue, S. W. Cutting edge: Central memory CD8 activated by the haematopoietic transcription factor epithelial cells. Cell. Immunol. 100, 112–118 (1986). T cells in aged mice are virtual memory cells. c‑Myb. Biochem. J. 372, 851–859 (2003). 113. Kino, T., Su, Y. A. & Chrousos, G. P. Human J. Immunol. 191, 5793–5796 (2013). 90. Pawlikowski, M., Zerek-Melen, G. & Winczyk, K. glucocorticoid receptor isoform β: recent 135. Grody, W. W., Fligiel, S. & Naeim, F. Thymus involution Thyroliberin (TRH) increases thymus cell proliferation understanding of its potential implications in in the acquired immunodeficiency syndrome. in rats. Neuropeptides 23, 199–202 (1992). physiology and pathophysiology. Cell. Mol. Life Sci. Am. J. Clin. Pathol. 84, 85–95 (1985). 91. Kawashima, I. et al. Localization of estrogen 66, 3435–3448 (2009). 136. Nezelof, C. Thymic pathology in primary and receptors and estrogen receptor-mRNA in female 114. Oakley, R. H., Webster, J. C., Sar, M., Parker, C. R. Jr secondary immunodeficiencies. Histopathology 21, mouse thymus. Thymus 20, 115–121 (1992). & Cidlowski, J. A. Expression and subcellular 499–511 (1992). 92. Nancy, P. & Berrih-Aknin, S. Differential estrogen distribution of the β-isoform of the human 137. Hermann, M. & Berger, P. Hormonal changes in aging receptor expression in autoimmune myasthenia glucocorticoid receptor. Endocrinology 138, men: a therapeutic indication? Exp. Gerontol. 36, gravis. Endocrinology 146, 2345–2353 (2005). 5028–5038 (1997). 1075–1082 (2001). 93. Viselli, S. M., Olsen, N. J., Shults, K., Steizer, G. & 115. Lechner, O. et al. Glucocorticoid production in the 138. Savino, W., Postel-Vinay, M. C., Smaniotto, S. Kovacs, W. J. Immunochemical and flow cytometric murine thymus. Eur. J. Immunol. 30, 337–346 & Dardenne, M. The thymus gland: a target organ for analysis of androgen receptor expression in (2000). growth hormone. Scand. J. Immunol. 55, 442–452 thymocytes. Mol. Cell Endocrinol. 109, 19–26 116. Vacchio, M. S., Papadopoulos, V. & Ashwell, J. D. (2002). (1995). Steroid production in the thymus: implications for 139. Kelley, K. W. et al. GH3 pituitary adenoma cells can 94. Olsen, N. J., Watson, M. B., Henderson, G. S. thymocyte selection. J. Exp. Med. 179, 1835–1846 reverse thymic aging in rats. Proc. Natl Acad. Sci. USA & Kovacs, W. J. Androgen deprivation induces (1994). 83, 5663–5667 (1986). phenotypic and functional changes in the thymus of 117. Qiao, S., Chen, L., Okret, S. & Jondal, M. 140. Savino, W., Smaniotto, S., Binart, N., adult male mice. Endocrinology 129, 2471–2476 Age-related synthesis of glucocorticoids in Postel-Vinay, M. C. & Dardenne, M. In vivo effects of (1991). thymocytes. Exp. Cell Res. 314, 3027–3035 (2008). growth hormone on thymic cells. Ann. N. Y. Acad. Sci. 95. Yellayi, S. et al. Normal development of thymus in 118. Lacaze-Masmonteil, T., de Keyzer, Y., Luton, J. P., 992, 179–185 (2003). male and female mice requires estrogen/estrogen Kahn, A. & Bertagna, X. Characterization of 141. Rosenfeld, R. G. et al. Growth hormone insensitivity receptor-α signaling pathway. Endocrine 12, transcripts in human resulting from post-GH receptor defects. Growth 207–213 (2000). nonpituitary tissues. Proc. Natl Acad. Sci. USA 84, Horm. IGF Res. 14 (Suppl. A), S35–S38 (2004). 96. Staples, J. E. et al. Estrogen receptor α is necessary 7261–7265 (1987). 142. Pugliese-Pires, P. N. et al. A novel STAT5B mutation in thymic development and estradiol-induced thymic 119. Jessop, D. S., Renshaw, D., Lightman, S. L. causing GH insensitivity syndrome associated with alterations. J. Immunol. 163, 4168–4174 (1999). & Harbuz, M. S. Changes in ACTH and β-endorphin hyperprolactinemia and immune dysfunction in two 97. Ishibashi, H. et al. Estrogen inhibits cell proliferation immunoreactivity in immune tissues during a chronic male siblings. Eur. J. Endocrinol. 163, 349–355 through in situ production in human thymoma. Clin. inflammatory stress are not correlated with changes (2010). Cancer Res. 11, 6495–6504 (2005). in corticotropin-releasing hormone and arginine 143. Dardenne, M., Smaniotto, S., de Mello-Coelho, V., 98. Utsuyama, M. & Hirokawa, K. Hypertrophy of the vasopressin. J. Neuroimmunol. 60, 29–35 (1995). Villa-Verde, D. M. & Savino, W. Growth hormone thymus and restoration of immune functions in mice 120. Talaber, G., Tuckermann, J. P. & Okret, S. ACTH modulates migration of developing T cells. Ann. and rats by gonadectomy. Mech. Ageing Dev. 47, controls thymocyte homeostasis independent N. Y. Acad. Sci. 1153, 1–5 (2009). 175–185 (1989). of glucocorticoids. FASEB J. 29, 2526–2534 (2015). 144. French, R. A. et al. Age-associated loss of bone 99. Azad, N., Emanuele, N. V., Halloran, M. M., 121. Aird, F., Clevenger, C. V., Prystowsky, M. B. & Redei, marrow hematopoietic cells is reversed by GH and Tentler, J. & Kelley, M. R. Presence of luteinizing E. Corticotropin-releasing factor mRNA in rat thymus accompanies thymic reconstitution. Endocrinology hormone-releasing hormone (LHRH) mRNA in rat and spleen. Proc. Natl Acad. Sci. USA 90, 143, 690–699 (2002). spleen lymphocytes. Endocrinology 128, 7104–7108 (1993). 145. Smaniotto, S., Ribeiro-Carvalho, M. M., Dardenne, M., 1679–1681 (1991). 122. Savino, W. The thymus is a common target organ in Savino, W. & de Mello-Coelho, V. Growth hormone 100. Marchetti, B. et al. Luteinizing hormone-releasing infectious diseases. PLoS Pathog. 2, e62 (2006). stimulates the selective trafficking of thymic hormone (LHRH) agonist restoration of age- 123. Mendes‑da‑Cruz, D. A., de Meis, J., CD4+CD8- emigrants to peripheral lymphoid organs. associated decline of thymus weight, thymic LHRH Cotta‑de‑Almeida, V. & Savino, W. Experimental Neuroimmunomodulation 11, 299–306 (2004). receptors, and thymocyte proliferative capacity. Trypanosoma cruzi infection alters the shaping of the 146. Montecino-Rodriguez, E., Clark, R. & Dorshkind, K. Endocrinology 125, 1037–1045 (1989). central and peripheral T‑cell repertoire. Microbes Effects of insulin-like growth factor administration and 101. Cohen, J. J. Glucocorticoid-induced apoptosis in the Infect. 5, 825–832 (2003). bone marrow transplantation on thymopoiesis in aged thymus. Semin. Immunol. 4, 363–369 (1992). 124. Morrot, A. et al. Chagasic thymic atrophy does not mice. Endocrinology 139, 4120–4126 (1998). 102. Berki, T., Palinkas, L., Boldizsar, F. & Nemeth, P. affect negative selection but results in the export of 147. Chu, Y. W. et al. Exogenous insulin-like growth factor 1 Glucocorticoid (GC) sensitivity and GC receptor activated CD4+CD8+ T cells in severe forms of enhances thymopoiesis predominantly through thymic expression differ in thymocyte subpopulations. Int. human disease. PLoS Negl. Trop. Dis. 5, e1268 epithelial cell expansion. Blood 112, 2836–2846 Immunol. 14, 463–469 (2002). (2011). (2008). 103. Mittelstadt, P. R., Monteiro, J. P. & Ashwell, J. D. 125. Mendes‑da‑Cruz, D. A., Silva, J. S., Cotta‑de‑Almeida, 148. Youm, Y. H. et al. Deficient ghrelin receptor-mediated Thymocyte responsiveness to endogenous V. & Savino, W. Altered thymocyte migration during signaling compromises thymic stromal cell glucocorticoids is required for immunological fitness. experimental acute Trypanosoma cruzi infection: microenvironment by accelerating thymic adiposity. J. Clin. Invest. 122, 2384–2394 combined role of fibronectin and the chemokines J. Biol. Chem. 284, 7068–7077 (2009). 104. Pálinkás, L. et al. Developmental shift in CXCL12 and CCL4. Eur. J. Immunol. 36, 1486–1493 149. Greenwood, P. L. & Bell, A. W. Consequences of intra- TcR-mediated rescue of thymocytes from (2006). uterine growth retardation for postnatal growth, glucocorticoid-induced apoptosis. Immunobiology 126. Perez, A. R. et al. Thymus atrophy during metabolism and pathophysiology. Reprod. Suppl. 61, 213, 39–50 (2008). Trypanosoma cruzi infection is caused by an immuno- 195–206 (2003). 105. Purton, J. F., Boyd, R. L., Cole, T. J. & Godfrey, D. I. endocrine imbalance. Brain Behav. Immun. 21, 150. Barr, I. G. et al. Dihydrotestosterone and estradiol Intrathymic T cell development and selection 890–900 (2007). deplete corticosensitive thymocytes lacking in proceeds normally in the absence of glucocorticoid 127. Roggero, E. et al. Endogenous glucocorticoids cause receptors for these hormones. J. Immunol. 128, receptor signaling. Immunity 13, 179–186 (2000). thymus atrophy but are protective during acute 2825–2828 (1982). 106. Tsawdaroglou, N. G., Govindan, M. V., Schmid, W. & Trypanosoma cruzi infection. J. Endocrinol. 190, 151. Kendall, M. D. et al. Reversal of ageing changes in the Sekeris, C. E. Dexamethasone-binding proteins in 495–503 (2006). thymus of rats by chemical or surgical castration. Cell cytosol and nucleus of rat thymocytes. Purification of 128. Correa‑de‑Santana, E. et al. Hypothalamus‑ Tissue Res. 261, 555–564 (1990). three receptor proteins. Eur. J. Biochem. 114, pituitary‑adrenal axis during Trypanosoma cruzi 152. Heng, T. S. et al. Effects of castration on thymocyte 305–313 (1981). acute infection in mice. J. Neuroimmunol. 173, development in two different models of thymic 107. McGimsey, W. C., Cidlowski, J. A., Stumpf, W. E. 12–22 (2006). involution. J. Immunol. 175, 2982–2993 (2005). & Sar, M. Immunocytochemical localization of the 129. Salas, M. A., Evans, S. W., Levell, M. J. & Whicher, 153. Greenstein, B. D., Fitzpatrick, F. T., Kendall, M. D. & glucocorticoid receptor in rat brain, pituitary, liver, J. T. Interleukin‑6 and ACTH act synergistically to Wheeler, M. J. Regeneration of the thymus in old male and thymus with two new polyclonal antipeptide stimulate the release of corticosterone from adrenal rats treated with a stable analogue of LHRH. antibodies. Endocrinology 129, 3064–3072 gland cells. Clin. Exp. Immunol. 79, 470–473 J. Endocrinol. 112, 345–350 (1987). (1991). (1990). 154. Velardi, E. et al. Sex steroid blockade enhances 108. Ratman, D. et al. How glucocorticoid receptors 130. Correa‑de‑Santana, E. et al. Modulation of growth thymopoiesis by modulating notch signaling. J. Exp. modulate the activity of other transcription factors: a hormone and prolactin secretion in Trypanosoma Med. 211, 2341–2349 (2014). scope beyond tethering. Mol. Cell. Endocrinol. 380, cruzi-infected mammosomatotrophic cells. 155. Mattsson, C. & Olsson, T. Estrogens and 41–54 (2013). Neuroimmunomodulation 16, 208–212 (2009). glucocorticoid hormones in adipose tissue

88 | FEBRUARY 2016 | VOLUME 12 www.nature.com/nrendo

© 2016 Macmillan Publishers Limited. All rights reserved REVIEWS

metabolism. Curr. Med. Chem. 14, 2918–2924 Hemidactylus flaviviridis: an in vitro study. Gen. Comp. 182. Sakabe, K., Kawashima, I., Urano, R., Seiki, K. (2007). Endocrinol. 144, 10–19 (2005). & Itoh, T. Effects of sex steroids on the proliferation 156. Zhao, H., Tian, Z., Hao, J. & Chen, B. Extragonadal 170. De Mello-Coelho, V., Savino, W., Postel-Vinay, M. C. & of thymic epithelial cells in a culture model: a role of aromatization increases with time after ovariectomy in Dardenne, M. Role of prolactin and growth hormone protein kinase C. Immunol. Cell Biol. 72, 193–199 rats. Reprod. Biol. Endocrinol. 3, 6 (2005). on thymus physiology. Dev. Immunol. 6, 317–323 (1994). 157. Limonta, P. et al. GnRH receptors in cancer: from cell (1998). 183. Talaber, G. et al. Wnt‑4 protects thymic epithelial biology to novel targeted therapeutic strategies. 171. Olsen, N. J., Olson, G., Viselli, S. M., Gu, X. & Kovacs, cells against dexamethasone-induced senescence. Endocr. Rev. 33, 784–811(2012). W. J. Androgen receptors in thymic epithelium Rejuvenation Res. 14, 241–248 (2011). 158. Savino, W. et al. Thymic epithelium in AIDS. modulate thymus size and thymocyte development. 184. Dardenne, M. et al. Thymic hormone-containing An immunohistologic study. Am. J. Pathol. 122, Endocrinology 142, 1278–1283 (2001). cells. VII. Adrenals and gonads control the 302–307 (1986). 172. Sacedon, R. et al. Partial blockade of T‑cell in vivo secretion of thymulin and its plasmatic 159. Napolitano, L. A. et al. Increased thymic mass and differentiation during ontogeny and marked inhibitor. J. Immunol. 136, 1303–1308 (1986). circulating naive CD4 T cells in HIV‑1‑infected adults alterations of the thymic microenvironment in 185. Coura, J. R. & Vinas, P. A. Chagas disease: treated with growth hormone. AIDS 16, 1103–1111 transgenic mice with impaired glucocorticoid receptor a new worldwide challenge. Nature 465, S6–S7 (2002). function. J. Neuroimmunol. 98, 157–167 (1999). (2010). 160. Tesselaar, K. & Miedema, F. Growth hormone 173. Jondal, M., Pazirandeh, A. & Okret, S. Different roles 186. Perez, A. R. et al. Immunoneuroendocrine resurrects adult human thymus during HIV‑1 infection. for glucocorticoids in thymocyte homeostasis? Trends alterations in patients with progressive forms of J. Clin. Invest. 118, 844–847 (2008). Immunol. 25, 595–600 (2004). chronic Chagas disease. J. Neuroimmunol. 235, 161. Haynes, B. F. HIV infection and the dynamic interplay 174. Lin, B., Kinoshita, Y., Hato, F. & Tsuji, Y. Enhancement 84–90 (2011). between the thymus and the peripheral T cell pool. of thymic lymphocyte proliferation by the culture 187. Savino, W., Leite‑de‑Moraes, M. C., Hontebeyrie- Clin. Immunol. 92, 3–5 (1999). supernatant of thymus epithelial cells stimulated by Joskowicz, M. & Dardenne, M. Studies on the 162. Napolitano, L. A. et al. Growth hormone enhances prolactin. Cell. Mol. Biol. (Noisy‑le‑grand) 43, thymus in Chagas’ disease. I. Changes in the thymic thymic function in HIV‑1‑infected adults. J. Clin. Invest. 361–367 (1997). microenvironment in mice acutely infected with 118, 1085–1098 (2008). 175. Pazirandeh, A., Xue, Y., Prestegaard, T., Jondal, M. & Trypanosoma cruzi. Eur. J. Immunol. 19, 163. Herasimtschuk, A. A. et al. Low-dose growth hormone Okret, S. Effects of altered glucocorticoid sensitivity in 1727–1733 (1989). for 40 weeks induces HIV‑1‑specific T cell responses in the T cell lineage on thymocyte and T cell homeostasis. 188. Vilar-Pereira, G. et al. Trypanosoma cruzi-induced patients on effective combination anti-retroviral FASEB J. 16, 727–729 (2002). depressive-like behavior is independent of therapy. Clin. Exp. Immunol. 173, 444–453 (2013). 176. Mihara, S. et al. Effects of on meningoencephalitis but responsive to parasiticide 164. Melis, M. R., Stancampiano, R. & Argiolas, A. apoptotic cell death of human lymphocytes. J. Clin. and TNF-targeted therapeutic interventions. Brain Oxytocin- and vasopressin-like immunoreactivity in Endocrinol. Metab. 84, 1378–1385 (1999). Behav. Immun. 26, 1136–1149 (2012). the rat thymus: characterization and possible 177. Olsen, N. J., Viselli, S. M., Fan, J. & Kovacs, W. J. 189. Perez, A. R., Bottasso, O. & Savino, W. The impact of involvement in the immune response. Regul. Pept. 45, Androgens accelerate thymocyte apoptosis. infectious diseases upon neuroendocrine circuits. 269–272 (1993). Endocrinology 139, 748–752 (1998). Neuroimmunomodulation 16, 96–105 (2009). 165. Johnson, E. W., Hughes, T. K. Jr & Smith, E. M. ACTH 178. Belloni, A. S. et al. Effect of ghrelin on the apoptotic receptor distribution and modulation among murine deletion rate of different types of cells cultured in vitro. Acknowledgements mononuclear leukocyte populations. J. Biol. Regul. Int. J. Mol. Med. 14, 165–167 (2004). Work in the authors’ laboratories was supported with grants Homeost. Agents 15, 156–162 (2001). 179. Alpdogan, O. et al. Insulin-like growth factor‑I from CNPq, Capes, Faperj and Fiocruz (Brazil), FOCEM 166. Kim, S. Y. et al. Preferential effects of leptin on CD4 enhances lymphoid and myeloid reconstitution after (Mercosur countries) and CNRS (France). The conjoint work T cells in central and peripheral immune system are allogeneic bone marrow transplantation. was developed in the framework of the Fiocruz–CNRS critically linked to the expression of leptin receptor. Transplantation 75, 1977–1983 (2003). International Associated Laboratory of Immunology Biochem. Biophys. Res. Commun. 394, 562–568 180. Montgomery, D. W., Krumenacker, J. S. & Buckley, and Immunopathology. (2010). A. R. Prolactin stimulates phosphorylation of the 167. Brtko, J. & Knopp, J. Rat thymus: demonstration of human T‑cell antigen receptor complex and ZAP‑70 specific thyroxine receptors in nuclear extract. tyrosine kinase: a potential mechanism for its Author contributions Endocrinol. Exp. 17, 3–9 (1983). immunomodulation. Endocrinology 139, 811–814 All authors researched data for the article, wrote, reviewed 168. Howard, J. K. et al. Leptin protects mice from (1998). and edited the manuscript before submission. D.A.M.‑d.-C., starvation-induced lymphoid atrophy and increases 181. Tsuji, Y., Kinoshita, Y., Hato, F., Tominaga, K. A.L. and W.S. made substantial contribution to discussion of thymic cellularity in ob/ob mice. J. Clin. Invest. 104, & Yoshida, K. The in vitro proliferation of thymus the content. 1051–1059 (1999). epithelial cells stimulated with growth hormone and 169. Hareramadas, B. & Rai, U. Mechanism of androgen- insulin-like growth factor‑I. Cell. Mol. Biol. Competing interests statement induced thymic atrophy in the wall lizard, (Noisy‑le‑grand) 40, 1135–1142 (1994). The authors declare no competing interests.

NATURE REVIEWS | ENDOCRINOLOGY VOLUME 12 | FEBRUARY 2016 | 89

© 2016 Macmillan Publishers Limited. All rights reserved