Fluoxetine effects on molecular, cellular and behavioral endophenotypes of depression are driven by the living environment

S Alboni, R M van Dijk, S Poggini, G Milior, M Perrotta, T Drenth, N Brunello, D P Wolfer, C Limatola, I Amrein, F Cirulli, L Maggi, I Branchi, S Alboni, R M van Dijk, S Poggini, G Milior, M Perrotta, T Drenth, N Brunello, D P Wolfer, C Limatola, I Amrein, F Cirulli, L Maggi, I Branchi

Abstract

Selective serotonin reuptake inhibitors (SSRIs) represent the most common treatment for major depression. However, their efficacy is variable and incomplete. In order to elucidate the cause of such incomplete efficacy, we explored the hypothesis positing that SSRIs may not affect mood per se but, by enhancing neural plasticity, render the individual more susceptible to the influence of the environment. Consequently, SSRI administration in a favorable environment promotes a reduction of symptoms, whereas in a stressful environment leads to a worse prognosis. To test such hypothesis, we exposed C57BL/6 mice to chronic stress in order to induce a depression-like phenotype and, subsequently, to fluoxetine treatment (21 days), while being exposed to either an enriched or a stressful condition. We measured the most commonly investigated molecular, cellular and behavioral endophenotypes of depression and SSRI outcome, including depression-like behavior, neurogenesis, brain-derived neurotrophic factor levels, hypothalamic-pituitary-adrenal axis activity and long-term potentiation. Results showed that, in line with our hypothesis, the endophenotypes investigated were affected by the treatment according to the quality of the living environment. In particular, mice treated with fluoxetine in an enriched condition overall improved their depression-like phenotype compared with controls, whereas those treated in a stressful condition showed a distinct worsening. Our findings suggest that the effects of SSRI on the depression- like phenotype is not determined by the drug per se but is induced by the drug and driven by the environment. These findings may be helpful to explain variable effects of SSRI found in clinical practice and to device strategies aimed at enhancing their efficacy by means of controlling environmental conditions.

Conflict of interest statement

The authors declare no conflict of interest.

Figures

Figure 1
Figure 1
Experimental design and effects of fluoxetine (FLX) administered either in enriched or stressful condition. (a) FLX treatment in the enriched condition. (b) FLX treatment in the stressful condition. In both cases, before treatment, mice are exposed to a 14-day period of stress to induce a depression-like phenotype. (c) Liking-type anhedonia (saccharin preference). Although no difference was found between FLX and vehicle (VEH) mice exposed to an enriched condition, a significantly increase of anhedonia (that is, reduction of saccharin preference) was found in FLX mice at the end of treatment in the stressful condition. VEH, n=23; FLX, n=24. (d) Wanting-type anhedonia (progressive ratio). When treatment was administered in the enriched condition, FLX led to a reduction of anhedonia (that is, a higher break point) at week 1, whereas in the stressful condition mice showed an increase in anhedonia (that is, lower break point) both at week 1 and 3, compared with VEH mice. Enrich: n=8 in all groups; stress: VEH, n=10, FLX, n=11. (e) Cognitive bias. In an enriched condition, FLX mice displayed a significantly ‘more optimistic' attitude than VEH by responding significantly more often to the ambiguous stimulus. No difference was found in the stressful condition. *P<0.05 vs relative VEH group. VEH, n=6, FLX, n=5. Data are means±s.e.m.
Figure 2
Figure 2
Fluoxetine (FLX) treatment administered in stressful condition leads to a reduction of proliferation and hippocampal volume. (a) Representative sections of histological (Giemsa) and immunohistochemical (Ki67 and doublecortin (DCX) counterstained with hematoxylin) stainings in the mid region of the matrix-embedded, straightened hippocampal dentate gyrus. Scale bar: 250 μm, insert scale bar: 10 μm. (b) Ki67 cell number was not significantly affected by FLX administered in enriched condition, whereas it was significantly decreased in FLX compared with vehicle (VEH) mice when treatment was administered in the stressful condition, *P<0.05 vs VEH group. (c) DCX cell number was not significantly affected by FLX administered in both conditions. (d) Schematic drawing of the straightened hippocampus. Gray part represents CA1. The analysis has been performed independently in the septal, mid and temporal part of the hippocampus, because it has been reported that the effects of SSRIs and the environment are region specific. Boundaries of hippocampal fields are illustrated in the Giemsa-stained section of the mid region of the straightened hippocampus cut perpendicular to the septotemporal axis. S, septal; T, temporal. (e) Analysis of anatomically aligned data of volumetric measurements showed no significant differences between groups in enriched condition. However, septal CA1 volume was significantly reduced in FLX compared with VEH when treatment was administered in stressful condition. *P<0.05. n=8 in all groups. Data shown as mean±s.e.m.
Figure 3
Figure 3
Fluoxetine (FLX) treatment affects hippocampal signaling pathways and HPA axis activity according to the quality of the environment. (a) No significant treatment effect was found for pERK1/ERK1 and in pERK2/ERK2 ratios in the enriched condition. Whereas both ratios were reduced in the cytoplasmic, but not in the nuclear, fraction in FLX mice in the stressful condition. (b) No treatment effect on CREB phosphorylation was observed in the nuclear enriched fraction in both environmental conditions. No difference in the total hippocampal CREB protein levels was found in the enriched condition, but it was reduced by treatment in the stressful condition. (c) In enriched conditions, reduced proBDNF and increased mBDNF levels were found in FLX mice compared with vehicle (VEH) mice. No difference in the stressful condition. (d and e) Real-time PCR analysis revealed that BDNF and p11 mRNA levels were significantly increased by FLX in the enriched condition but were not affected in the stressful condition. (f) Representative western blottings are shown. (g) Corticosterone levels were significantly reduced by treatment in FLX compared with VEH mice in the enriched but not in the stressful condition. (h) Glucocorticoid receptor (GR) mRNA expression was reduced by treatment in the stressful condition. No difference for GR expression in the enriched condition and mineralocorticoid receptor (MR) expression in both condition was found. MR, mineralocorticoid receptor. Corticosterone analysis: VEH, n= 6; FLX, n=7. For all the other analyses, n=8 in all groups. Data shown as mean±s.e.m. *P<0.05 vs respective VEH group.
Figure 4
Figure 4
Fluoxetine (FLX) modifies molecular and cellular correlates of synaptic plasticity in an environment-dependent manner. (a) Hippocampal levels of AMPA receptor subunits GluR1 and GluR2, and (b) their phosphorylation measured by western blotting. FLX induced opposite effects on GluR1 phosphorylation at Ser 845 according to the quality of the environment. (c) Hippocampal levels of NMDA receptor subunits GluN1, GluN2A and GluN2B measured by western blotting. (d) Representative western blots. n=8 for all groups in western blot analyses. Data shown as mean±s.e.m. of the control. *P<0.05 vs respective vehicle (VEH) group. (e) FLX increased paired-pulse ratio in the stressful condition but had no effect in the enriched condition. VEH: n=18/5; FLX: n=15/4. (f) FLX affect CA1 plasticity in mice exposed to stress condition. In the enriched condition, both FLX and VEH mice developed a robust LTP 45 min after stimulation (single 100 Hz burst; VEH: n=12/5; FLX: n=11/5). In the stressful condition, FLX mice showed a remarkable increase in the LTP amplitude compared with VEH (VEH: n=14/6; FLX: n=10/5). Arrows indicate time of application of HFS. fEPSP, field excitatorypostsynaptic potential; PPR, paired-pulse ratio. Data shown as mean±s.e.m. *P<0.05 and **P<0.01 vs respective VEH group.

References

    1. WHOThe Global Burden of Disease: 2004 Update. World Health Organization: Geneva, Switzerland, 2008.
    1. Balak N, Elmaci I. Costs of disorders of the brain in Europe. Eur J Neurol 2007; 14: e9.
    1. Trivedi MH, Rush AJ, Wisniewski SR, Nierenberg AA, Warden D, Ritz L et al. Evaluation of outcomes with citalopram for depression using measurement-based care in STAR*D: implications for clinical practice. Am J Psychiatry 2006; 163: 28–40.
    1. Kirsch I, Deacon BJ, Huedo-Medina TB, Scoboria A, Moore TJ, Johnson BT. Initial severity and antidepressant benefits: a meta-analysis of data submitted to the Food and Drug Administration. PLoS Med 2008; 5: e45.
    1. Bessa JM, Ferreira D, Melo I, Marques F, Cerqueira JJ, Palha JA et al. The mood-improving actions of antidepressants do not depend on neurogenesis but are associated with neuronal remodeling. Mol Psychiatry 2009; 14: 764–773, 739.
    1. Rygula R, Abumaria N, Flugge G, Hiemke C, Fuchs E, Ruther E et al. Citalopram counteracts depressive-like symptoms evoked by chronic social stress in rats. Behav Pharmacol 2006; 17: 19–29.
    1. Malberg JE, Eisch AJ, Nestler EJ, Duman RS. Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 2000; 20: 9104–9110.
    1. Nibuya M, Nestler EJ, Duman RS. Chronic antidepressant administration increases the expression of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci 1996; 16: 2365–2372.
    1. Pariante CM. Glucocorticoid receptor function in vitro in patients with major depression. Stress 2004; 7: 209–219.
    1. Uys JD, Muller CJ, Marais L, Harvey BH, Stein DJ, Daniels WM. Early life trauma decreases glucocorticoid receptors in rat dentate gyrus upon adult re-stress: reversal by escitalopram. Neuroscience 2006; 137: 619–625.
    1. Bath KG, Jing DQ, Dincheva I, Neeb CC, Pattwell SS, Chao MV et al. BDNF Val66Met impairs fluoxetine-induced enhancement of adult hippocampus plasticity. Neuropsychopharmacology 2012; 37: 1297–1304.
    1. Brenes JC, Fornaguera J. The effect of chronic fluoxetine on social isolation-induced changes on sucrose consumption, immobility behavior, and on serotonin and dopamine function in hippocampus and ventral striatum. Behav Brain Res 2009; 198: 199–205.
    1. Prendergast MA, Yells DP, Balogh SE, Paige SR, Hendricks SE. Fluoxetine differentially suppresses sucrose solution consumption in free-fed and food-deprived rats—reversal by amantadine. Med Sci Monit 2002; 8: BR385–BR390.
    1. Sammut S, Bethus I, Goodall G, Muscat R. Antidepressant reversal of interferon-alpha-induced anhedonia. Physiol Behav 2002; 75: 765–772.
    1. Tonissaar M, Mallo T, Eller M, Haidkind R, Koiv K, Harro J. Rat behavior after chronic variable stress and partial lesioning of 5-HT-ergic neurotransmission: effects of citalopram. Prog Neuropsychopharmacol Biol Psychiatry 2008; 32: 164–177.
    1. David DJ, Samuels BA, Rainer Q, Wang JW, Marsteller D, Mendez I et al. Neurogenesis- dependent and -independent effects of fluoxetine in an animal model of anxiety/depression. Neuron 2009; 62: 479–493.
    1. Marlatt MW, Potter MC, Bayer TA, van Praag H, Lucassen PJ. Prolonged running, not fluoxetine treatment, increases neurogenesis, but does not alter neuropathology, in the 3xTg mouse model of Alzheimer's disease. Curr Topics Behavi Neurosci 2013; 15: 313–340.
    1. Wu MV, Shamy JL, Bedi G, Choi CW, Wall MM, Arango V et al. Impact of social status and antidepressant treatment on neurogenesis in the baboon hippocampus. Neuropsychopharmacology 2014; 39: 1861–1871.
    1. Klomp A, Vaclavu L, Meerhoff GF, Reneman L, Lucassen PJ. Effects of chronic fluoxetine treatment on neurogenesis and tryptophan hydroxylase expression in adolescent and adult rats. PLoS One 2014; 9: e97603.
    1. Possamai F, dos Santos J, Walber T, Marcon JC, dos Santos TS, Lino de Oliveira C. Influence of enrichment on behavioral and neurogenic effects of antidepressants in Wistar rats submitted to repeated forced swim test. Prog Neuropsychopharmacol Biol Psychiatry 2015; 58: 15–21.
    1. Jacobsen JP, Mork A. The effect of escitalopram, desipramine, electroconvulsive seizures and lithium on brain-derived neurotrophic factor mRNA and protein expression in the rat brain and the correlation to 5-HT and 5-HIAA levels. Brain Res 2004; 1024: 183–192.
    1. Kozisek ME, Middlemas D, Bylund DB. Brain-derived neurotrophic factor and its receptor tropomyosin-related kinase B in the mechanism of action of antidepressant therapies. Pharmacol Ther 2008; 117: 30–51.
    1. Miro X, Perez-Torres S, Artigas F, Puigdomenech P, Palacios JM, Mengod G. Regulation of cAMP phosphodiesterase mRNAs expression in rat brain by acute and chronic fluoxetine treatment. An in situ hybridization study. Neuropharmacology 2002; 43: 1148–1157.
    1. Zetterstrom TS, Pei Q, Madhav TR, Coppell AL, Lewis L, Grahame-Smith DG. Manipulations of brain 5-HT levels affect gene expression for BDNF in rat brain. Neuropharmacology 1999; 38: 1063–1073.
    1. Alboni S, Benatti C, Capone G, Corsini D, Caggia F, Tascedda F et al. Time-dependent effects of escitalopram on brain derived neurotrophic factor (BDNF) and neuroplasticity related targets in the central nervous system of rats. Eur J Pharmacol 2010; 643: 180–187.
    1. Goekint M, Roelands B, Heyman E, Njemini R, Meeusen R. Influence of citalopram and environmental temperature on exercise-induced changes in BDNF. Neurosci Lett 2011; 494: 150–154.
    1. Shen Q, Lal R, Luellen BA, Earnheart JC, Andrews AM, Luscher B. gamma-Aminobutyric acid- type A receptor deficits cause hypothalamic-pituitary-adrenal axis hyperactivity and antidepressant drug sensitivity reminiscent of melancholic forms of depression. Biol Psychiatry 2010; 68: 512–520.
    1. Weber CC, Eckert GP, Muller WE. Effects of antidepressants on the brain/plasma distribution of corticosterone. Neuropsychopharmacology 2006; 31: 2443–2448.
    1. Rubio FJ, Ampuero E, Sandoval R, Toledo J, Pancetti F, Wyneken U. Long-term fluoxetine treatment induces input-specific LTP and LTD impairment and structural plasticity in the CA1 hippocampal subfield. Front Cell Neurosci 2013; 7: 66.
    1. Stewart CA, Reid IC. Repeated ECS and fluoxetine administration have equivalent effects on hippocampal synaptic plasticity. Psychopharmacology 2000; 148: 217–223.
    1. Branchi I. The double edged sword of neural plasticity: increasing serotonin levels leads to both greater vulnerability to depression and improved capacity to recover. Psychoneuroendocrinology 2011; 36: 339–351.
    1. Belsky J, Jonassaint C, Pluess M, Stanton M, Brummett B, Williams R. Vulnerability genes or plasticity genes? Mol Psychiatry 2009; 14: 746–754.
    1. Maya Vetencourt JF, Sale A, Viegi A, Baroncelli L, De Pasquale R, O'Leary OF et al. The antidepressant fluoxetine restores plasticity in the adult visual cortex. Science 2008; 320: 385–388.
    1. Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N, Agustsdottir A et al. Fear erasure in mice requires synergy between antidepressant drugs and extinction training. Science 2011; 334: 1731–1734.
    1. Brummett BH, Boyle SH, Siegler IC, Kuhn CM, Ashley-Koch A, Jonassaint CR et al. Effects of environmental stress and gender on associations among symptoms of depression and the serotonin transporter gene linked polymorphic region (5-HTTLPR). Behav Genet 2008; 38: 34–43.
    1. Eley TC, Sugden K, Corsico A, Gregory AM, Sham P, McGuffin P et al. Gene-environment interaction analysis of serotonin system markers with adolescent depression. Mol Psychiatry 2004; 9: 908–915.
    1. Mastronardi C, Paz-Filho GJ, Valdez E, Maestre-Mesa J, Licinio J, Wong ML. Long-term body weight outcomes of antidepressant-environment interactions. Mol Psychiatry 2011; 16: 265–272.
    1. Wong ML, Licinio J. Research and treatment approaches to depression. Nat Rev Neurosci 2001; 2: 343–351.
    1. Benatti C, Alboni S, Montanari C, Caggia F, Tascedda F, Brunello N et al. Central effects of a local inflammation in three commonly used mouse strains with a different anxious phenotype. Behav Brain Res 2011; 224: 23–34.
    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. Small SA, Schobel SA, Buxton RB, Witter MP, Barnes CA. A pathophysiological framework of hippocampal dysfunction in ageing and disease. Nat Rev Neurosci 2011; 12: 585–601.
    1. Santarelli L, Saxe M, Gross C, Surget A, Battaglia F, Dulawa S et al. Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science 2003; 301: 805–809.
    1. Navailles S, Hof PR, Schmauss C. Antidepressant drug-induced stimulation of mouse hippocampal neurogenesis is age-dependent and altered by early life stress. J Comp Neurol 2008; 509: 372–381.
    1. Cowen DS, Takase LF, Fornal CA, Jacobs BL. Age-dependent decline in hippocampal neurogenesis is not altered by chronic treatment with fluoxetine. Brain Res 2008; 1228: 14–19.
    1. Bremner JD, Narayan M, Anderson ER, Staib LH, Miller HL, Charney DS. Hippocampal volume reduction in major depression. Am J Psychiatry 2000; 157: 115–118.
    1. Gourley SL, Wu FJ, Kiraly DD, Ploski JE, Kedves AT, Duman RS et al. Regionally specific regulation of ERK MAP kinase in a model of antidepressant-sensitive chronic depression. Biol Psychiatry 2008; 63: 353–359.
    1. Duric V, Banasr M, Licznerski P, Schmidt HD, Stockmeier CA, Simen AA et al. A negative regulator of MAP kinase causes depressive behavior. Nat Med 2010; 16: 1328–1332.
    1. Chen AC, Shirayama Y, Shin KH, Neve RL, Duman RS. Expression of the cAMP response element binding protein (CREB) in hippocampus produces an antidepressant effect. Biol Psychiatry 2001; 49: 753–762.
    1. Thome J, Sakai N, Shin K, Steffen C, Zhang YJ, Impey S et al. cAMP response element-mediated gene transcription is upregulated by chronic antidepressant treatment. J Neurosci 2000; 20: 4030–4036.
    1. Shirayama Y, Chen AC, Nakagawa S, Russell DS, Duman RS. Brain-derived neurotrophic factor produces antidepressant effects in behavioral models of depression. J Neurosci 2002; 22: 3251–3261.
    1. Warner-Schmidt JL, Chen EY, Zhang X, Marshall JJ, Morozov A, Svenningsson P et al. A role for p11 in the antidepressant action of brain-derived neurotrophic factor. Biol Psychiatry 2010; 68: 528–535.
    1. Svenningsson P, Kim Y, Warner-Schmidt J, Oh YS, Greengard P. p11 and its role in depression and therapeutic responses to antidepressants. Nat Rev Neurosci 2013; 14: 673–680.
    1. Alboni S, Tascedda F, Corsini D, Benatti C, Caggia F, Capone G et al. Stress induces altered CRE/CREB pathway activity and BDNF expression in the hippocampus of glucocorticoid receptor-impaired mice. Neuropharmacology 2011; 60: 1337–1346.
    1. Holsboer F, Ising M. Stress hormone regulation: biological role and translation into therapy. Ann Rev Psychol 2010; 61: 81–109, C101-C111.
    1. Ising M, Horstmann S, Kloiber S, Lucae S, Binder EB, Kern N et al. Combined dexamethasone/corticotropin releasing hormone test predicts treatment response in major depression - a potential biomarker? Biol Psychiatry 2007; 62: 47–54.
    1. de Kloet ER, Fitzsimons CP, Datson NA, Meijer OC, Vreugdenhil E. Glucocorticoid signaling and stress-related limbic susceptibility pathway: about receptors, transcription machinery and microRNA. Brain Res 2009; 1293: 129–141.
    1. Datson NA, Speksnijder N, Mayer JL, Steenbergen PJ, Korobko O, Goeman J et al. The transcriptional response to chronic stress and glucocorticoid receptor blockade in the hippocampal dentate gyrus. Hippocampus 2012; 22: 359–371.
    1. Furay AR, Bruestle AE, Herman JP. The role of the forebrain glucocorticoid receptor in acute and chronic stress. Endocrinology 2008; 149: 5482–5490.
    1. Anacker C, Zunszain PA, Carvalho LA, Pariante CM. The glucocorticoid receptor: pivot of depression and of antidepressant treatment? Psychoneuroendocrinology 2011; 36: 415–425.
    1. Alme MN, Wibrand K, Dagestad G, Bramham CR. Chronic fluoxetine treatment induces brain region-specific upregulation of genes associated with BDNF-induced long-term potentiation. Neural Plasticity 2007; 2007: 26496.
    1. Kobayashi K, Ikeda Y, Asada M, Inagaki H, Kawada T, Suzuki H. Corticosterone facilitates fluoxetine-induced neuronal plasticity in the hippocampus. PLoS One 2013; 8: e63662.
    1. Wiles N, Thomas L, Abel A, Ridgway N, Turner N, Campbell J et al. Cognitive behavioural therapy as an adjunct to pharmacotherapy for primary care based patients with treatment resistant depression: results of the CoBalT randomised controlled trial. Lancet 2012; 381: 375–384.
    1. Marsden WN. Synaptic plasticity in depression: molecular, cellular and functional correlates. Prog Neuropsychopharmacol Biol Psychiatry 2013; 43: 168–184.
    1. Wang JW, David DJ, Monckton JE, Battaglia F, Hen R. Chronic fluoxetine stimulates maturation and synaptic plasticity of adult-born hippocampal granule cells. J Neurosci 2008; 28: 1374–1384.
    1. Svenningsson P, Tzavara ET, Witkin JM, Fienberg AA, Nomikos GG, Greengard P. Involvement of striatal and extrastriatal DARPP-32 in biochemical and behavioral effects of fluoxetine (Prozac). Proc NatlAcad Sci USA 2002; 99: 3182–3187.
    1. Lee HK, Barbarosie M, Kameyama K, Bear MF, Huganir RL. Regulation of distinct AMPA receptor phosphorylation sites during bidirectional synaptic plasticity. Nature 2000; 405: 955–959.
    1. Peng Y, Zhao J, Gu QH, Chen RQ, Xu Z, Yan JZ et al. Distinct trafficking and expression mechanisms underlie LTP and LTD of NMDA receptor-mediated synaptic responses. Hippocampus 2010; 20: 646–658.
    1. Joels M, Krugers HJ. LTP after stress: up or down? Neural Plasticity 2007; 2007: 93202.
    1. Duffy SN, Craddock KJ, Abel T, Nguyen PV. Environmental enrichment modifies the PKA- dependence of hippocampal LTP and improves hippocampus-dependent memory. Learn Memory 2001; 8: 26–34.
    1. Zucker RS. Short-term synaptic plasticity. Ann Rev Neurosci 1989; 12: 13–31.
    1. Lazarus RS, Folkman S. Stress, Appraisal, and Coping. Springer: New York, 1984.
    1. Kramer PD. Listening to Prozac: Psychiatrist Explores Antidepressant Drugs and the Remaking of the Self. Fourth Estate Ltd: London, 1994.
    1. Kraemer HC, Frank E, Kupfer DJ. Moderators of treatment outcomes: clinical, research, and policy importance. JAMA 2006; 296: 1286–1289.
    1. Belmaker RH, Agam G. Major depressive disorder. N Engl J Med 2008; 358: 55–68.
    1. Krishnan V, Nestler EJ. Linking molecules to mood: new insight into the biology of depression. Am J Psychiatry 2010; 167: 1305–1320.
    1. Kobayashi K, Ikeda Y, Sakai A, Yamasaki N, Haneda E, Miyakawa T et al. Reversal of hippocampal neuronal maturation by serotonergic antidepressants. Proc Natl Acad Sci USA 2010; 107: 8434–8439.
    1. Baroncelli L, Braschi C, Spolidoro M, Begenisic T, Sale A, Maffei L. Nurturing brain plasticity: impact of environmental enrichment. Cell Death Differ 2010; 17: 1092–1103.
    1. Ibarguen-Vargas Y, Surget A, Touma C, Palme R, Belzung C. Multifaceted strain-specific effects in a mouse model of depression and of antidepressant reversal. Psychoneuroendocrinology 2008; 33: 1357–1368.
    1. Surget A, Tanti A, Leonardo ED, Laugeray A, Rainer Q, Touma C et al. Antidepressants recruit new neurons to improve stress response regulation. Mol Psychiatry 2011; 16: 1177–1188.
    1. Tanti A, Westphal WP, Girault V, Brizard B, Devers S, Leguisquet AM et al. Region-dependent and stage-specific effects of stress, environmental enrichment, and antidepressant treatment on hippocampal neurogenesis. Hippocampus 2013; 23: 797–811.
    1. Jain FA, Hunter AM, Brooks JO 3rd, Leuchter AF. Predictive socioeconomic and clinical profiles of antidepressant response and remission. Depress Anxiety 2013; 30: 624–630.
    1. Cohen A, Houck PR, Szanto K, Dew MA, Gilman SE, Reynolds CF 3rd. Social inequalities in response to antidepressant treatment in older adults. Arch Gen psychiatry 2006; 63: 50–56.
    1. Cohen A, Gilman SE, Houck PR, Szanto K, Reynolds CF 3rd. Socioeconomic status and anxiety as predictors of antidepressant treatment response and suicidal ideation in older adults. Soc Psychiatry Psychiatr Epidemiol 2009; 44: 272–277.

Source: PubMed

3
订阅