The ESC/E(Z) complex, an effector of response to ovarian steroids, manifests an intrinsic difference in cells from women with premenstrual dysphoric disorder

N Dubey, J F Hoffman, K Schuebel, Q Yuan, P E Martinez, L K Nieman, D R Rubinow, P J Schmidt, D Goldman, N Dubey, J F Hoffman, K Schuebel, Q Yuan, P E Martinez, L K Nieman, D R Rubinow, P J Schmidt, D Goldman

Abstract

Clinical evidence suggests that mood and behavioral symptoms in premenstrual dysphoric disorder (PMDD), a common, recently recognized, psychiatric condition among women, reflect abnormal responsivity to ovarian steroids. This differential sensitivity could be due to an unrecognized aspect of hormonal signaling or a difference in cellular response. In this study, lymphoblastoid cell line cultures (LCLs) from women with PMDD and asymptomatic controls were compared via whole-transcriptome sequencing (RNA-seq) during untreated (ovarian steroid-free) conditions and following hormone treatment. The women with PMDD manifested ovarian steroid-triggered behavioral sensitivity during a hormone suppression and addback clinical trial, and controls did not, leading us to hypothesize that women with PMDD might differ in their cellular response to ovarian steroids. In untreated LCLs, our results overall suggest a divergence between mRNA (for example, gene transcription) and protein (for example, RNA translation in proteins) for the same genes. Pathway analysis of the LCL transcriptome revealed, among others, over-expression of ESC/E(Z) complex genes (an ovarian steroid-regulated gene silencing complex) in untreated LCLs from women with PMDD, with more than half of these genes over-expressed as compared with the controls, and with significant effects for MTF2, PHF19 and SIRT1 (P<0.05). RNA and protein expression of the 13 ESC/E(Z) complex genes were individually quantitated. This pattern of increased ESC/E(Z) mRNA expression was confirmed in a larger cohort by qRT-PCR. In contrast, protein expression of ESC/E(Z) genes was decreased in untreated PMDD LCLs with MTF2, PHF19 and SIRT1 all significantly decreased (P<0.05). Finally, mRNA expression of several ESC/E(Z) complex genes were increased by progesterone in controls only, and decreased by estradiol in PMDD LCLs. These findings demonstrate that LCLs from women with PMDD manifest a cellular difference in ESC/E(Z) complex function both in the untreated condition and in response to ovarian hormones. Dysregulation of ESC/E(Z) complex function could contribute to PMDD.

Trial registration: ClinicalTrials.gov NCT00001259 NCT00001322.

Conflict of interest statement

The Authors have no potential conflicts of interest or financial support regarding this manuscript.

Figures

Figure 1. LCL treatments and analyses
Figure 1. LCL treatments and analyses
Flow chart of LCL experiments. Blood is collected from women with PMDD or asymptomatic Controls and transformed into LCLs with Epstein Barr virus. Cells are then left untreated or treated with hormone (estradiol or progesterone, 100 nM each) and undergo RNA-sequencing. Pathway analyses were applied to results, and selected genes from the indicated LCL groups are followed up with qRT-PCR and protein analysis via ProteinSimple©.
Figure 2. mRNA and protein expression for…
Figure 2. mRNA and protein expression for hormone receptors and ESC/E(Z) Complex genes comparing untreated Control and PMDD lymphoblastoid cell lines
All graphs are presented as group mean ± SEM. (A–F) Student’s t-test was performed for each comparison. (A) qRT-PCR analysis of mRNA expression for hormone receptors ESR1, ESR2, PGRMC1, and PGRMC2 in lymphoblastoid cell lines (LCLs), n=15. Relative gene expression levels were calculated for ∆CT value with β-actin as the reference control for normalization; note a lower ∆CT value indicates higher expression. Both ESR1 and PGRMC1 were significantly higher in the PMDD group than the Control group (*p<0.05), and ESR2 was higher at a trend level of significance in the PMDD group than the Control group (#p<0.1). (B) Verification of β-actin as a normalizing protein for ProteinSimple© western analysis of protein expression in LCLs. The chemiluminescense ratio of β-actin over the ProteinSimple© Loading Control (PSCon) was calculated for each group and no significant difference was found (n=7, t12=0.02, p=0.9). (C–F) ProteinSimple© western analysis of protein expression for hormone receptors ESR1, ESR2, PGRMC1, and PGRMC2 (n=7). Target protein chemiluminescense was normalized to β-actin or PSCon if there was band interference with β-actin, as was the case for ESR2 and PGRMC2. There were no significant differences for any of the proteins. For (B–F), representative images of protein “bands” from the Compass software are shown below each group. (G–H) All graphs are calculated as percent change of values of the PMDD group from the Control group mean. Statistical analysis comes from Student’s t-test of the original group means (see Supplemental Figure 2). (G) Percent change from control of mRNA expression via RNAseq for the 13 genes of the ESC/E(Z) complex (n=9–10). Both MTF2 and SIRT1 were significantly higher in the PMDD group than the Control group (*p<0.05), and EED, HDAC2, and SUZ12 were higher at a trend level of significance in the PMDD group than the Control group (#p<0.1), while PHF19 was lower at a trend level of significance in the PMDD group than the control group (#p<0.1). (H) Percent change from control of mRNA expression via qRT-PCR for the 13 genes of the ESC/E(Z) complex (n=29–30). HDAC2 (t43=2.6), MTF2 (t37=3.7), SIRT1 (t46=3.2), and SUZ12 (t42=2.1) were significantly higher in the PMDD group than the Control group (*p<0.05), and RBBP7 (t33=1.8) was trending higher in the PMDD group than the Control group (#p<0.1). (I) Percent change from control of protein expression via ProteinSimple© western analysis for select genes of the ESC/E(Z) complex (n=7). NT indicates that protein was not tested. MTF2 (t11=3.4), PHF19 (t12=2.5), and SIRT1 (t11=3.2) were significantly lower in the PMDD group than the Control group (*p<0.05).
Figure 3. Selected ESC/E(Z) gene expressions from…
Figure 3. Selected ESC/E(Z) gene expressions from RNAseq: ANOVAs comparing untreated and estradiol- or progesterone-treated Control and PMDD lymphoblastoid cell lines
All graphs are presented as group mean ± SEM using RPKM values from RNA-sequencing. Two-way ANOVAs were performed for each gene of interest, followed by Sidak’s multiple comparisons post-hoc test in cases of a significant interaction. Untreated Control and PMDD lymphoblastoid cell lines compared with estradiol treatment (A) or progesterone treatment (B-F). (A) For JARID2 there was a significant interaction effect between diagnosis and estradiol treatment (F1,24=5.5, *p<0.05). Expression was significantly decreased after estradiol only in the PMDD group (t24=2.5, adjusted p-value<0.05). (B) For EED there was a significant interaction effect between diagnosis and progesterone treatment (F1,29=6.8, *p<0.05). Expression was significantly increased after progesterone only in the Control group (t29=2.8, adjusted p-value<0.05). (C) For EZH2 there was a significant interaction effect between diagnosis and progesterone treatment (F1,29=4.8, *p<0.05). Expression was significantly increased after progesterone only in the Control group (t29=2.4, adjusted p-value<0.05). (D) For MTF2 there was a significant interaction effect between diagnosis and progesterone treatment (F1,29=9.2, *p<0.05). Expression was significantly increased after progesterone only in the Control group (t29=2.7, adjusted p-value<0.05). Additionally, there was also a significant diagnosis effect (F1,29=5.5, *p<0.05) where MTF2 is more highly expressed in the PMDD group than the Control group (post-hoc untreated Control vs untreated PMDD t29=4.1, p<0.001). (E) For PHF19 there was a significant diagnosis effect (F1,29=16.1, *p<0.05) where expression is lower in the PMDD group compared with the Control group, but there was no interaction effect between diagnosis and progesterone treatment (F1,29=0.1, p=0.7). (E) For SIRT1 there was a significant diagnosis effect (F1,29=13.1, *p<0.05) where expression was higher in the PMDD group compared to the Control group, but there was no interaction effect between diagnosis and progesterone treatment (F1,29=0.5, p=0.5).
Figure 4. GeneMANIA interaction analysis of ESC/E(Z)…
Figure 4. GeneMANIA interaction analysis of ESC/E(Z) and steroid hormone receptor genes
GeneMANIA (genemania.org) was used to generate a gene network analysis inputting the 13 genes of the ESC/E(Z) complex (AEBP2, EED, EZH1, EZH2, HDAC2, JARID2, MTF2, PHF1, PHF19, RBBP4, RBBP7, SIRT1, SUZ12) and the 4 hormone receptor genes (ESR1, ESR2, PGRMC1, PGRMC2), whose expression was identified in LCLs. Genes and targets are included in the network for the input genes based on weights of interactions with the input genes, determined automatically by linear regression. Black circles indicate the input genes. Gray circles indicate additional genes identified in the network – a larger circle reflects a higher rank of connection with the input genes. Diamonds indicate miRNA or transcription factor targets. A thicker connecting bar between nodes reflects a stronger relationship between those two nodes. (A) The expanded network based on a search of these 17 input genes (September, 2015). The strongest relationships driving this network are Predicted Interactions (44.22% weight) and Physical Interactions (41.35% weight). (B) Selection of HDAC2 and highlighting its interactions within the predicted network. HDAC2 shows interactions with 9 of the 13 genes of the ESC/E(Z) complex (7 of them physical), as well as several other genes in the network. Importantly, it is the only component of the ESC/E(Z) complex that interacts with a receptor for both estradiol (ESR1) and progesterone (PGRMC2).

References

    1. American Psychiatric Association Diagnostic and Statistical Manual of Mental Disorders, Fifth Edition. Fifth. Arlington, VA: 2013.
    1. Yonkers K, O'Brien PMS, Eriksson E. Premenstrual syndrome. Lancet. 2008;371:1200–1210.
    1. Epperson CN, Steiner M, Hartlage SA, Eriksson E, Schmidt PJ, Jones I, et al. Premenstrual dysphoric disorder: evidence for a new category for DSM-5. Am J Psychiatry. 2012;169:465–475.
    1. Pearlstein T, Steiner M. Premenstrual dysphoric disorder: burden of illness and treatment update. J Psychiatry Neurosci. 2008;33:291–301.
    1. Halbreich U, Borenstein J, Pearlstein T, Kahn LS. The prevalence, impairment, impact, and burden of premenstrual dysphoric disorder (PMS/PMDD) Psychoneuroendocrinology. 2003;28:1–23.
    1. Schiller CE, Johnson AL, Abate AC, Rubinow DR, Schmidt PJ. Reproductive steroid regulation of mood and behavior. Comprehensive Physiology. 2016
    1. Hammarback S, Ekholm UB, Backstrom T. Spontaneous anovulation causing disappearance of cyclical symptoms in women with the premenstrual syndrome. Acta Endocrinol (Copenh) 1991;125:132–137.
    1. Schmidt PJ, Nieman LK, Danaceau MA, Adams LF, Rubinow DR. Differential behavioral effects of gonadal steroids in women with and in those without premenstrual syndrome. N Engl J Med. 1998;338:209–216.
    1. Helvacioglu A, Yeoman RR, Hazelton JM, Aksel S. Premenstrual syndrome and related hormonal changes: long-acting gonadotropin releasing hormone agonist treatment. J Reprod Med. 1993;38:864–870.
    1. Brown CS, Ling FW, Andersen RN, Farmer RG, Arheart KL. Efficacy of depot leuprolide in premenstrual syndrome: effect of symptom severity and type in a controlled trial. Obstet Gynecol. 1994;84:779–786.
    1. Freeman EW, Sondheimer SJ, Rickels K, Albert J. Gonadotropin-releasing hormone agonist in treatment of premenstrual symptoms: with and without comorbidity of depression: a pilot study. J Clin Psychiatry. 1993;54:192–195.
    1. West CP, Hillier H. Ovarian suppression with the gonadotrophin-releasing hormone agonist goserelin (Zoladex) in management of the premenstrual tension syndrome. Hum Reprod. 1994;9:1058–1063.
    1. Hussain SY, Massil JH, Matta WH, Shaw RW, O'Brien PMS. Buserelin in premenstrual syndrome. Gynecol Endocrinol. 1992;6:57–64.
    1. Protopopescu X, Pan H, Altemus M, Tuescher O, Polanecsky M, McEwen B, et al. Orbitofrontal cortex activity related to emotional processing changes across the menstrual cycle. Proc Natl Acad Sci U S A. 2005;102:16060–16065.
    1. Protopopescu X, Butler T, Pan H, Root J, Altemus M, Polanecsky M, et al. Hippocampal structural changes across the menstrual cycle. Hippocampus. 2008;18:985–988.
    1. Protopopescu X, Tuescher O, Pan H, Epstein J, Root J, Chang L, et al. Toward a functional neuroanatomy of premenstrual dysphoric disorder. J Affective Disord. 2007;108:87–94.
    1. Baller EB, Wei SM, Kohn P, Rubinow DR, Alarcon G, Schmidt PJ, et al. Abnormalities of dorsolateral prefrontal function in women with premenstrual dysphoric disorder: a multimodal neuroimaging study. Am J Psychiatry. 2013;170:305–314.
    1. Andreano JM, Cahill L. Menstrual cycle modulation of medial temporal activity evoked by negative emotion. Neuroimage. 2010;53:1286–1293.
    1. Goldstein JM, Jerram M, Abbs B, Whitfield-Gabrieli S, Makris N. Sex differences in stress response circuitry activation dependent on female hormonal cycle. J Neurosci. 2010;30:431–438.
    1. Gingnell M, Bannbers E, Wikstrom J, Fredrikson M, Sundstrom-Poromaa I. Premenstrual dysphoric disorder and prefrontal reactivity during anticipation of emotional stimuli. Eur Neuropsychopharmacol. 2013;23:1474–1483.
    1. Sundstrom PI, Gingnell M. Menstrual cycle influence on cognitive function and emotion processing-from a reproductive perspective. Front Neurosci. 2014;8:380.
    1. Dreher J, Schmidt PJ, Kohn P, Furman D, Rubinow D, Berman KF. Menstrual cycle phase modulates reward-related neural function in women. Proc Natl Acad Sci U S A. 2007;104:2465–2470.
    1. Berman KF, Schmidt PJ, Rubinow DR, Danaceau MA, Van Horn JD, Esposito G, et al. Modulation of cognition-specific cortical activity by gonadal steroids: a positron-emission tomography study in women. Proc Natl Acad Sci U S A. 1997;94:8836–8841.
    1. Toffoletto S, Lanzenberger R, Gingnell M, Sundstrom-Poromaa I, Comasco E. Emotional and cognitive functional imaging of estrogen and progesterone effects in the female human brain: a systematic review. Psychoneuroendocrinology. 2014;50:28–52.
    1. Schmidt PJ, Ben Dor R, Martinez PE, Guerrieri GM, Harsh VL, Thompson K, et al. Effects of Estradiol Withdrawal on Mood in Women With Past Perimenopausal Depression: A Randomized Clinical Trial. JAMA Psychiatry. 2015;72:714–726.
    1. Bloch M, Schmidt PJ, Danaceau M, Murphy J, Nieman L, Rubinow DR. Effects of gonadal steroids in women with a history of postpartum depression. Am J Psychiatry. 2000;157:924–930.
    1. Mehta D, Newport DJ, Frishman G, Kraus L, Rex-Haffner M, Ritchie JC, et al. Early predictive biomarkers for postpartum depression point to a role for estrogen receptor signaling. Psychol Med. 2014;44:2309–2322.
    1. Guintivano J, Arad M, Gould TD, Payne JL, Kaminsky ZA. Antenatal prediction of postpartum depression with blood DNA methylation biomarkers. Mol Psychiatry. 2014;19:560–567.
    1. Gladkevich A, Kauffman HF, Korf J. Lymphocytes as a neural probe: potential for studying psychiatric disorders. Prog Neuropsychopharmacol Biol Psychiatry. 2004;28:559–576.
    1. Mohr S, Liew CC. The peripheral-blood transcriptome: new insights into disease and risk assessment. Trends Mol Med. 2007;13:422–432.
    1. Nicholson AC, Unger ER, Mangalathu R, Ojaniemi H, Vernon SD. Exploration of neuroendocrine and immune gene expression in peripheral blood mononuclear cells. Brain Res Mol Brain Res. 2004;129:193–197.
    1. Sie L, Loong S, Tan EK. Utility of lymphoblastoid cell lines. J Neurosci Res. 2009;87:1953–1959.
    1. de Mello AS, Provin F, Michelin-Tireli K, Camelier MV, Coelho JC. Feasibility of using cryopreserved lymphoblastoid cells to diagnose some lysosomal storage diseases. Cell Prolif. 2010;43:164–169.
    1. Bahn S, Chan MK. What can we learn about depression from gene expression in peripheral tissues? Biol Psychiatry. 2015;77:207–209.
    1. Iga J, Ueno S, Ohmori T. Molecular assessment of depression from mRNAs in the peripheral leukocytes. Ann Med. 2008;40:336–342.
    1. Karsten SL, Kudo LC, Bragin AJ. Use of peripheral blood transcriptome biomarkers for epilepsy prediction. Neurosci Lett. 2011;497:213–217.
    1. Tylee DS, Kawaguchi DM, Glatt SJ. On the outside, looking in: a review and evaluation of the comparability of blood and brain "-omes". Am J Med Genet B Neuropsychiatr Genet. 2013;162B:595–603.
    1. American Psychiatric Association Diagnostic and Statistical Manual of Mental Disorders Fourth Edition. Washington, DC: American Psychiatric Association; 1994.
    1. Rubinow DR, Roy-Byrne PP, Hoban MC, Gold PW, Post RM. Prospective assessment of menstrually related mood disorders. Am J Psychiatry. 1984;141:684–686.
    1. Endicott J, Nee J, Cohen J, Halbreich U. Premenstrual changes: patterns and correlates of daily ratings. J Affective Disord. 1986;10:127–135.
    1. First MB, Spitzer RL, Gibbon M, Williams JBW. Biometrics Research. New York: New York State Psychiatric Institute; 2002. Structured Clinical Interview for DSM-IV-TR Axis I Disorders, Research Version, Patient Edition (SCID-IP)
    1. Halbreich U, Backstrom T, Eriksson E, O'Brien S, Calil H, Ceskova E, et al. Clinical diagnosis criteria for premenstrual syndrome and guidelines for their quantification for research studies. Gynecol Endocrinol. 2007;23:123–130.
    1. Halbreich U. The diagnosis of premenstrual syndromes and premenstrual dysphoric disorder - clinical procedures and research perspectives. Gynecol Endocrinol. 2004;19:320–334.
    1. Freeman EW, Sondheimer SJ, Rickels K. Gonadotropin-releasing hormone agonist in the treatment of premenstrual symptoms with and without ongoing dysphoria: a controlled study. Psychopharmacol Bull. 1997;33:303–309.
    1. Ben Dor R, Harsh VL, Fortinsky P, Koziol DE, Rubinow DR, Schmidt PJ. Effects of pharmacologically-induced hypogonadism on mood and behavior in healthy young women. Am J Psychiatry. 2013;170:426–433.
    1. Oh HM, Oh JM, Choi SC, Kim SW, Han WC, Kim TH, et al. An efficient method for the rapid establishment of Epstein-Barr virus immortalization of human B lymphocytes. Cell Prolif. 2003;36:191–197.
    1. Holmans P. Statistical methods for pathway analysis of genome-wide data for association with complex genetic traits. Adv Genet. 2010;72:141–179.
    1. O'Neill RA, Bhamidipati A, Bi X, Deb-Basu D, Cahill L, Ferrante J, et al. Isoelectric focusing technology quantifies protein signaling in 25 cells. Proc Natl Acad Sci U S A. 2006;103:16153–16158.
    1. Bhan A, Hussain I, Ansari KI, Bobzean SA, Perrotti LI, Mandal SS. Histone methyltransferase EZH2 is transcriptionally induced by estradiol as well as estrogenic endocrine disruptors bisphenol-A and diethylstilbestrol. J Mol Biol. 2014;426:3426–3441.
    1. Pal B, Bouras T, Shi W, Vaillant F, Sheridan JM, Fu N, et al. Global changes in the mammary epigenome are induced by hormonal cues and coordinated by Ezh2. Cell Rep. 2013;3:411–426.
    1. Bredfeldt TG, Greathouse KL, Safe SH, Hung MC, Bedford MT, Walker CL. Xenoestrogen-induced regulation of EZH2 and histone methylation via estrogen receptor signaling to PI3K/AKT. Mol Endocrinol. 2010;24:993–1006.
    1. Izzo F, Mercogliano F, Venturutti L, Tkach M, Inurrigarro G, Schillaci R, et al. Progesterone receptor activation downregulates GATA3 by transcriptional repression and increased protein turnover promoting breast tumor growth. Breast Cancer Res. 2014;16:491.
    1. Tamm-Rosenstein K, Simm J, Suhorutshenko M, Salumets A, Metsis M. Changes in the transcriptome of the human endometrial Ishikawa cancer cell line induced by estrogen, progesterone, tamoxifen, and mifepristone (RU486) as detected by RNA-sequencing. PLoS One. 2013;8:e68907.
    1. Han L, Wang P, Zhao G, Wang H, Wang M, Chen J, et al. Upregulation of SIRT1 by 17beta-estradiol depends on ubiquitin-proteasome degradation of PPAR-gamma mediated by NEDD4-1. Protein Cell. 2013;4:310–321.
    1. Moore RL, Dai Y, Faller DV. Sirtuin 1 (SIRT1) and steroid hormone receptor activity in cancer. J Endocrinol. 2012;213:37–48.
    1. Yang S, Jia Y, Liu X, Winters C, Wang X, Zhang Y, et al. Systematic dissection of the mechanisms underlying progesterone receptor downregulation in endometrial cancer. Oncotarget. 2014;5:9783–9797.
    1. Hwang C, Giri VN, Wilkinson JC, Wright CW, Wilkinson AS, Cooney KA, et al. EZH2 regulates the transcription of estrogen-responsive genes through association with REA, an estrogen receptor corepressor. Breast Caner Res Treat. 2008;107:235–242.
    1. Wu P, Zuo X, Deng H, Liu X, Liu L, Ji A. Roles of long noncoding RNAs in brain development, functional diversification and neurodegenerative diseases. Brain Res Bull. 2013;97:69–80.
    1. Margueron R, Reinberg D. The Polycomb complex PRC2 and its mark in life. Nature. 2011;469:343–349.
    1. Di CL, Helin K. Transcriptional regulation by Polycomb group proteins. Nat Struct Mol Biol. 2013;20:1147–1155.
    1. Simon JA, Kingston RE. Mechanisms of polycomb gene silencing: knowns and unknowns. Nat Rev Mol Cell Biol. 2009;10:697–708.
    1. Crowley SK, Girdler SS. Neurosteroid, GABAergic and hypothalamic pituitary adrenal (HPA) axis regulation: what is the current state of knowledge in humans? Psychopharmacology (Berl) 2014;231:3619–3634.
    1. Saha S, Hu Y, Martin SC, Bandyopadhyay S, Russek SJ, Farb DH. Polycomblike protein PHF1b: a transcriptional sensor for GABA receptor activity. BMC Pharmacol Toxicol. 2013;14:37.
    1. Epperson CN, Haga K, Mason GF, Sellers E, Gueorguieva R, Zhang W, et al. Cortical γ-aminobutyric acid levels across the menstrual cycle in healthy women and those with premenstrual dysphoric disorder: a proton magnetic resonance spectroscopy study. Arch Gen Psychiatry. 2002;59:851–858.
    1. Smith MJ, Adams LF, Schmidt PJ, Rubinow DR, Wassermann EM. Abnormal luteal phase excitability of the motor cortex in women with premenstrual syndrome. Biol Psychiatry. 2003;54:757–762.
    1. Schiller CE, Schmidt PJ, Rubinow DR. Allopregnanolone as a mediator of affective switching in reproductive mood disorders. Psychopharmacology (Berl) 2014;231:3557–3567.
    1. Hunter RG, McCarthy KJ, Milne TA, Pfaff DW, McEwen BS. Regulation of hippocampal H3 histone methylation by acute and chronic stress. Proc Natl Acad Sci U S A. 2009;106:20912–20917.
    1. Peter CJ, Akbarian S. Balancing histone methylation activities in psychiatric disorders. Trends Mol Med. 2011;17:372–379.
    1. Machado-Vieira R, Ibrahim L, Zarate CA., Jr Histone deacetylases and mood disorders: epigenetic programming in gene-environment interactions. CNS Neurosci Ther. 2011;17:699–704.
    1. Ernst C, Chen ES, Turecki G. Histone methylation and decreased expression of TrkB.T1 in orbital frontal cortex of suicide completers. Mol Psychiatry. 2009;14:830–832.
    1. Kishi T, Yoshimura R, Kitajima T, Okochi T, Okumura T, Tsunoka T, et al. SIRT1 gene is associated with major depressive disorder in the Japanese population. J Affect Disord. 2010;126:167–173.
    1. Kenworthy CA, Sengupta A, Luz SM, Ver Hoeve ES, Meda K, Bhatnagar S, et al. Social defeat induces changes in histone acetylation and expression of histone modifying enzymes in the ventral hippocampus, prefrontal cortex, and dorsal raphe nucleus. Neuroscience. 2014;264:88–98.
    1. Renthal W, Kumar A, Xiao G, Wilkinson M, Covington HE, III, Maze I, et al. Genome-wide analysis of chromatin regulation by cocaine reveals a role for sirtuins. Neuron. 2009;62:335–348.
    1. Tsankova NM, Berton O, Renthal W, Kumar A, Neve RL, Nestler EJ. Sustained hippocampal chromatin regulation in a mouse model of depression and antidepressant action. Nat Neurosci. 2006;9:519–525.
    1. Etchegaray JP, Yang X, DeBruyne JP, Peters AH, Weaver DR, Jenuwein T, et al. The polycomb group protein EZH2 is required for mammalian circadian clock function. J Biol Chem. 2006;281:21209–21215.
    1. Nasca C, Xenos D, Barone Y, Caruso A, Scaccianoce S, Matrisciano F, et al. L-acetylcarnitine causes rapid antidepressant effects through the epigenetic induction of mGlu2 receptors. Proc Natl Acad Sci U S A. 2013;110:4804–4809.
    1. Qi C, Liu S, Qin R, Zhang Y, Wang G, Shang Y, et al. Coordinated regulation of dendrite arborization by epigenetic factors CDYL and EZH2. J Neurosci. 2014;34:4494–4508.
    1. Pereira JD, Sansom SN, Smith J, Dobenecker MW, Tarakhovsky A, Livesey FJ. Ezh2, the histone methyltransferase of PRC2, regulates the balance between self-renewal and differentiation in the cerebral cortex. Proc Natl Acad Sci U S A. 2010;107:15957–15962.
    1. Yan W, Fang Z, Yang Q, Dong H, Lu Y, Lei C, et al. SirT1 mediates hyperbaric oxygen preconditioning-induced ischemic tolerance in rat brain. J Cereb Blood Flow Metab. 2013;33:396–406.
    1. Libert S, Pointer K, Bell EL, Das A, Cohen DE, Asara JM, et al. SIRT1 activates MAO-A in the brain to mediate anxiety and exploratory drive. Cell. 2011;147:1459–1472.
    1. Sparse whole-genome sequencing identifies two loci for major depressive disorder. Nature. 2015;523:588–591.
    1. Licinio J, Wong ML. The role of inflammatory mediators in the biology of major depression: central nervous system cytokines modulate the biological substrate of depressive symptoms, regulate stress-responsive systems, and contribute to neurotoxicity and neuroprotection. Mol Psychiatry. 1999;4:317–327.
    1. Gold PW. The organization of the stress system and its dysregulation in depressive illness. Mol Psychiatry. 2015;20:32–47.
    1. Miller AH, Raison CL. The role of inflammation in depression: from evolutionary imperative to modern treatment target. Nat Rev Immunol. 2016;16:22–34.
    1. Bertone-Johnson ER, Ronnenberg AG, Houghton SC, Nobles C, Zagarins SE, Takashima-Uebelhoer BB, et al. Association of inflammation markers with menstrual symptom severity and premenstrual syndrome in young women. Hum Reprod. 2014;29:1987–1994.
    1. Gold EB, Wells C, Rasor MO. The Association of Inflammation with Premenstrual Symptoms. J Womens Health (Larchmt) 2016
    1. Huo L, Straub RE, Schmidt PJ, Shi K, Vakkalanka R, Weinberger DR, et al. Risk for premenstrual dysphoric disorder is associated with genetic variation in ESR1, the estrogen receptor alpha gene. Biol Psychiatry. 2007;62:925–933.
    1. Tseng RC, Wang YC. SIRT1 (sirtuin (silent mating type information regulation 2 homolog) 1 (S. cerevisiae)) Atlas of Genetics and Cytogenetics in Oncology and Haematology. 2010;14:1152–1156.
    1. Roth M, Chen WY. Sorting out functions of sirtuins in cancer. Oncogene. 2014;33:1609–1620.
    1. Li X, Isono K, Yamada D, Endo TA, Endoh M, Shinga J, et al. Mammalian polycomb-like Pcl2/Mtf2 is a novel regulatory component of PRC2 that can differentially modulate polycomb activity both at the Hox gene cluster and at Cdkn2a genes. Mol Cell Biol. 2011;31:351–364.
    1. Kininis M, Kraus WL. A global view of transcriptional regulation by nuclear receptors: gene expression, factor localization, and DNA sequence analysis. Nucl Recept Signal. 2008;6:e005.
    1. Hah N, Kraus WL. Hormone-regulated transcriptomes: lessons learned from estrogen signaling pathways in breast cancer cells. Mol Cell Endocrinol. 2014;382:652–664.
    1. Villablanca AC, Lewis KA, Rutledge JC. Time- and dose-dependent differential upregulation of three genes by 17 beta-estradiol in endothelial cells. J Appl Physiol (1985) 2002;92:1064–1073.
    1. Brinton RD, Thompson RF, Foy MR, Baudry M, Wang J, Finch CE, et al. Progesterone receptors: form and function in brain. Front Neuroendocrinol. 2008;29:313–339.
    1. Bali N, Arimoto JM, Iwata N, Lin SW, Zhao L, Brinton RD, et al. Differential responses of progesterone receptor membrane component-1 (Pgrmc1) and the classical progesterone receptor (Pgr) to 17β-estradiol and progesterone in hippocampal subregions that support synaptic remodeling and neurogenesis. Endocrinology. 2012;153:759–769.
    1. Mani SK, Blaustein JD. Neural progestin receptors and female sexual behavior. Neuroendocrinology. 2012;96:152–161.
    1. Thomas P, Pang Y. Membrane progesterone receptors: evidence for neuroprotective, neurosteroid signaling and neuroendocrine functions in neuronal cells. Neuroendocrinology. 2012;96:162–171.
    1. Peluso JJ, DeCerbo J, Lodde V. Evidence for a genomic mechanism of action for progesterone receptor membrane component-1. Steroids. 2012;77:1007–1012.
    1. Thomas P. Characteristics of membrane progestin receptor alpha (mPRalpha) and progesterone membrane receptor component 1 (PGMRC1) and their roles in mediating rapid progestin actions. Front Neuroendocrinol. 2008;29:292–312.
    1. Schumacher M, Mattern C, Ghoumari A, Oudinet JP, Liere P, Labombarda F, et al. Revisiting the roles of progesterone and allopregnanolone in the nervous system: resurgence of the progesterone receptors. Prog Neurobiol. 2014;113:6–39.
    1. Tie F, Furuyama T, Prasad-Sinha J, Jane E, Harte PJ. The Drosophila Polycomb Group proteins ESC and E(Z) are present in a complex containing the histone-binding protein p55 and the histone deacetylase RPD3. Development. 2001;128:275–286.
    1. Van der Vlag J, Otte AP. Transcriptional repression mediated by the human polycomb-group protein EED involves histone deacetylation. Nat Genet. 1999;23:474–478.
    1. Kendler KS, Silberg JL, Neale MC, Kessler RC, Heath AC, Eaves LJ. Genetic and environmental factors in the aetiology of menstrual, premenstrual and neurotic symptoms: a population-based twin study. Psychol Med. 1992;22:85–100.
    1. Endicott J, Halbreich U. Retrospective report of premenstrual depressive changes: Factors affecting confirmation by daily ratings. Psychopharmacol Bull. 1982;18:109–112.
    1. Garcia-Gonzalo FR, Izpisua Belmonte JC. Albumin-associated lipids regulate human embryonic stem cell self-renewal. PLoS One. 2008;3:e1384.

Source: PubMed

3
Subscribe