1932

Abstract

Depression is an episodic form of mental illness characterized by mood state transitions with poorly understood neurobiological mechanisms. Antidepressants reverse the effects of stress and depression on synapse function, enhancing neurotransmission, increasing plasticity, and generating new synapses in stress-sensitive brain regions. These properties are shared to varying degrees by all known antidepressants, suggesting that synaptic remodeling could play a key role in depression pathophysiology and antidepressant function. Still, it is unclear whether and precisely how synaptogenesis contributes to mood state transitions. Here, we review evidence supporting an emerging model in which depression is defined by a distinct brain state distributed across multiple stress-sensitive circuits, with neurons assuming altered functional properties, synapse configurations, and, importantly, a reduced capacity for plasticity and adaptation. Antidepressants act initially by facilitating plasticity and enabling a functional reconfiguration of this brain state. Subsequently, synaptogenesis plays a specific role in sustaining these changes over time.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-neuro-110920-040422
2022-07-08
2024-04-19
Loading full text...

Full text loading...

/deliver/fulltext/neuro/45/1/annurev-neuro-110920-040422.html?itemId=/content/journals/10.1146/annurev-neuro-110920-040422&mimeType=html&fmt=ahah

Literature Cited

  1. Abdallah CG, Averill LA, Gueorguieva R, Goktas S, Purohit P et al. 2020. Modulation of the antidepressant effects of ketamine by the mTORC1 inhibitor rapamycin. Neuropsychopharmacology 45:6990–97
    [Google Scholar]
  2. Abdallah CG, De Feyter HM, Averill LA, Jiang L, Averill CL et al. 2018. The effects of ketamine on prefrontal glutamate neurotransmission in healthy and depressed subjects. Neuropsychopharmacology 43:102154–60
    [Google Scholar]
  3. Abdallah CG, Jackowski A, Salas R, Gupta S, Sato JR et al. 2017. The nucleus accumbens and ketamine treatment in major depressive disorder. Neuropsychopharmacology 42:81739–46
    [Google Scholar]
  4. Ali F, Gerhard DM, Sweasy K, Pothula S, Pittenger C et al. 2020. Ketamine disinhibits dendrites and enhances calcium signals in prefrontal dendritic spines. Nat. Commun. 11:172
    [Google Scholar]
  5. Ampuero E, Rubio FJ, Falcon R, Sandoval M, Diaz-Veliz G et al. 2010. Chronic fluoxetine treatment induces structural plasticity and selective changes in glutamate receptor subunits in the rat cerebral cortex. Neuroscience 169:198–108
    [Google Scholar]
  6. Andalman AS, Burns VM, Lovett-Barron M, Broxton M, Poole B et al. 2019. Neuronal dynamics regulating brain and behavioral state transitions. Cell 177:4970–85.e20
    [Google Scholar]
  7. Autry AE, Adachi M, Nosyreva E, Na ES, Los MF et al. 2011. NMDA receptor blockade at rest triggers rapid behavioural antidepressant responses. Nature 475:735491–95
    [Google Scholar]
  8. Barrett FS, Doss MK, Sepeda ND, Pekar JJ, Griffiths RR. 2020. Emotions and brain function are altered up to one month after a single high dose of psilocybin. Sci. Rep. 10:12214
    [Google Scholar]
  9. Bath KG, Jing DQ, Dincheva I, Neeb CC, Pattwell SS et al. 2012. BDNF Val66Met impairs fluoxetine-induced enhancement of adult hippocampus plasticity. Neuropsychopharmacology 37:51297–304
    [Google Scholar]
  10. Berman RM, Cappiello A, Anand A, Oren DA, Heninger GR et al. 2000. Antidepressant effects of ketamine in depressed patients. Biol. Psychiatry 47:4351–54
    [Google Scholar]
  11. Berton O, McClung CA, Dileone RJ, Krishnan V, Renthal W et al. 2006. Essential role of BDNF in the mesolimbic dopamine pathway in social defeat stress. Science 311:5762864–68
    [Google Scholar]
  12. Bessa JM, Ferreira D, Melo I, Marques F, Cerqueira JJ et al. 2009. The mood-improving actions of antidepressants do not depend on neurogenesis but are associated with neuronal remodeling. Mol. Psychiatry 14:8764–73
    [Google Scholar]
  13. Boldrini M, Santiago AN, Hen R, Dwork AJ, Rosoklija GB et al. 2013. Hippocampal granule neuron number and dentate gyrus volume in antidepressant-treated and untreated major depression. Neuropsychopharmacology 38:61068–77
    [Google Scholar]
  14. Cameron LP, Tombari RJ, Lu J, Pell AJ, Hurley ZQ et al. 2021. A non-hallucinogenic psychedelic analogue with therapeutic potential. Nature 589:7842474–79
    [Google Scholar]
  15. Carhart-Harris R, Giribaldi B, Watts R, Baker-Jones M, Murphy-Beiner A et al. 2021. Trial of psilocybin versus escitalopram for depression. N. Engl. J. Med. 384:151402–11
    [Google Scholar]
  16. Casarotto PC, Girych M, Fred SM, Kovaleva V, Moliner R et al. 2021. Antidepressant drugs act by directly binding to TRKB neurotrophin receptors. Cell 184:51299–313.e19
    [Google Scholar]
  17. Castrén E, Hen R. 2013. Neuronal plasticity and antidepressant actions. Trends Neurosci 36:5259–67
    [Google Scholar]
  18. Cerniauskas I, Winterer J, de Jong JW, Lukacsovich D, Yang H et al. 2019. Chronic stress induces activity, synaptic, and transcriptional remodeling of the lateral habenula associated with deficits in motivated behaviors. Neuron 104:5899–915.e8
    [Google Scholar]
  19. Chaudhury D, Walsh JJ, Friedman AK, Juarez B, Ku SM et al. 2013. Rapid regulation of depression-related behaviours by control of midbrain dopamine neurons. Nature 493:7433532–36
    [Google Scholar]
  20. Chen Z-Y, Jing D, Bath KG, Ieraci A, Khan T et al. 2006. Genetic variant BDNF (Val66Met) polymorphism alters anxiety-related behavior. Science 314:5796140–43
    [Google Scholar]
  21. Christoffel DJ, Golden SA, Dumitriu D, Robison AJ, Janssen WG et al. 2011. IκB kinase regulates social defeat stress-induced synaptic and behavioral plasticity. J. Neurosci. 31:1314–21
    [Google Scholar]
  22. Cole EJ, Stimpson KH, Bentzley BS, Gulser M, Cherian K et al. 2020. Stanford accelerated intelligent neuromodulation therapy for treatment-resistant depression. Am. J. Psychiatry 177:8716–26
    [Google Scholar]
  23. Conrad CD, Galea LAM, Kuroda Y, McEwen BS. 1996. Chronic stress impairs rat spatial memory on the Y maze, and this effect is blocked by tianeptine treatment. Behav. Neurosci. 110:61321–34
    [Google Scholar]
  24. DeLorenzo C, DellaGioia N, Bloch M, Sanacora G, Nabulsi N et al. 2015. In vivo ketamine-induced changes in [11C]ABP688 binding to metabotropic glutamate receptor subtype 5. Biol. Psychiatry 77:3266–75
    [Google Scholar]
  25. Diamond DM, Rose GM. 1994. Stress impairs LTP and hippocampal-dependent memory. Ann. N. Y. Acad. Sci. 746:411–14
    [Google Scholar]
  26. Dias-Ferreira E, Sousa JC, Melo I, Morgado P, Mesquita AR et al. 2009. Chronic stress causes frontostriatal reorganization and affects decision-making. Science 325:5940621–25
    [Google Scholar]
  27. Diazgranados N, Ibrahim L, Brutsche NE, Newberg A, Kronstein P et al. 2010. A randomized add-on trial of an N-methyl-d-aspartate antagonist in treatment-resistant bipolar depression. Arch. Gen. Psychiatry 67:8793–802
    [Google Scholar]
  28. Dong C, Ly C, Dunlap LE, Vargas MV, Sun J et al. 2021. Psychedelic-inspired drug discovery using an engineered biosensor. Cell 184:102779–92.e18
    [Google Scholar]
  29. Drevets WC, Price JL, Furey ML. 2008. Brain structural and functional abnormalities in mood disorders: implications for neurocircuitry models of depression. Brain Struct. Funct. 213:1–293–118
    [Google Scholar]
  30. Drevets WC, Videen TO, Price JL, Preskorn SH, Carmichael ST, Raichle ME. 1992. A functional anatomical study of unipolar depression. J. Neurosci. 12:93628–41
    [Google Scholar]
  31. Drysdale AT, Grosenick L, Downar J, Dunlop K, Mansouri F et al. 2017. Resting-state connectivity biomarkers define neurophysiological subtypes of depression. Nat. Med. 23:128–38
    [Google Scholar]
  32. Duman RS, Aghajanian GK. 2012. Synaptic dysfunction in depression: potential therapeutic targets. Science 338:610368–72
    [Google Scholar]
  33. Duman RS, Aghajanian GK, Sanacora G, Krystal JH. 2016. Synaptic plasticity and depression: new insights from stress and rapid-acting antidepressants. Nat. Med. 22:3238–49
    [Google Scholar]
  34. Esterlis I, DellaGioia N, Pietrzak RH, Matuskey D, Nabulsi N et al. 2018. Ketamine-induced reduction in mGluR5 availability is associated with an antidepressant response: an [11C]ABP688 and PET imaging study in depression. Mol. Psychiatry 23:4824–32
    [Google Scholar]
  35. Feldman S, Conforti N, Saphier D. 1990. The preoptic area and bed nucleus of the stria terminalis are involved in the effects of the amygdala on adrenocortical secretion. Neuroscience 37:3775–79
    [Google Scholar]
  36. Fox MD, Buckner RL, White MP, Greicius MD, Pascual-Leone A. 2012. Efficacy of transcranial magnetic stimulation targets for depression is related to intrinsic functional connectivity with the subgenual cingulate. Biol. Psychiatry 72:7595–603
    [Google Scholar]
  37. Fox ME, Chandra R, Menken MS, Larkin EJ, Nam H et al. 2020. Dendritic remodeling of D1 neurons by RhoA/Rho-kinase mediates depression-like behavior. Mol. Psychiatry 25:51022–34
    [Google Scholar]
  38. Francis TC, Chandra R, Gaynor A, Konkalmatt P, Metzbower SR et al. 2017. Molecular basis of dendritic atrophy and activity in stress susceptibility. Mol. Psychiatry 22:111512–19
    [Google Scholar]
  39. Gerhard DM, Pothula S, Liu R-J, Wu M, Li X-Y et al. 2020. GABA interneurons are the cellular trigger for ketamine's rapid antidepressant actions. J. Clin. Investig. 130:31336–49
    [Google Scholar]
  40. Goldwater DS, Pavlides C, Hunter RG, Bloss EB, Hof PR et al. 2009. Structural and functional alterations to rat medial prefrontal cortex following chronic restraint stress and recovery. Neuroscience 164:2798–808
    [Google Scholar]
  41. Gould E, McEwen BS, Tanapat P, Galea LA, Fuchs E. 1997. Neurogenesis in the dentate gyrus of the adult tree shrew is regulated by psychosocial stress and NMDA receptor activation. J. Neurosci. 17:72492–98
    [Google Scholar]
  42. Gould E, Tanapat P, McEwen BS, Flügge G, Fuchs E. 1998. Proliferation of granule cell precursors in the dentate gyrus of adult monkeys is diminished by stress. PNAS 95:63168–71
    [Google Scholar]
  43. Guo W, Machado-Vieira R, Mathew S, Murrough JW, Charney DS et al. 2018. Exploratory genome-wide association analysis of response to ketamine and a polygenic analysis of response to scopolamine in depression. Transl. Psychiatry 9:1108
    [Google Scholar]
  44. Halldorsdottir T, Binder EB. 2017. Gene x environment interactions: from molecular mechanisms to behavior. Annu. Rev. Psychol. 68:215–41
    [Google Scholar]
  45. Hamilton JP, Gotlib IH. 2008. Neural substrates of increased memory sensitivity for negative stimuli in major depression. Biol. Psychiatry 63:121155–62
    [Google Scholar]
  46. Hasler G, van der Veen JW, Tumonis T, Meyers N, Shen J, Drevets WC. 2007. Reduced prefrontal glutamate/glutamine and γ-aminobutyric acid levels in major depression determined using proton magnetic resonance spectroscopy. Arch. Gen. Psychiatry 64:2193–200
    [Google Scholar]
  47. Hayashi-Takagi A, Yagishita S, Nakamura M, Shirai F, Wu YI et al. 2015. Labelling and optical erasure of synaptic memory traces in the motor cortex. Nature 525:7569333–38
    [Google Scholar]
  48. Helm MS, Dankovich TM, Mandad S, Rammner B, Jähne S et al. 2021. A large-scale nanoscopy and biochemistry analysis of postsynaptic dendritic spines. Nat. Neurosci. 24:81151–62
    [Google Scholar]
  49. Hesselgrave N, Troppoli TA, Wulff AB, Cole AB, Thompson SM. 2021. Harnessing psilocybin: antidepressant-like behavioral and synaptic actions of psilocybin are independent of 5-HT2R activation in mice. PNAS 118:17e2022489118
    [Google Scholar]
  50. Hikida T, Kimura K, Wada N, Funabiki K, Nakanishi S. 2010. Distinct roles of synaptic transmission in direct and indirect striatal pathways to reward and aversive behavior. Neuron 66:6896–907
    [Google Scholar]
  51. Hirschfeld RM. 2000. History and evolution of the monoamine hypothesis of depression. J. Clin. Psychiatry 61:Suppl. 64–6
    [Google Scholar]
  52. Holmes SE, Scheinost D, Finnema SJ, Naganawa M, Davis MT et al. 2019. Lower synaptic density is associated with depression severity and network alterations. Nat. Commun. 10:11529
    [Google Scholar]
  53. Homayoun H, Moghaddam B. 2007. NMDA receptor hypofunction produces opposite effects on prefrontal cortex interneurons and pyramidal neurons. J. Neurosci. 27:4311496–500
    [Google Scholar]
  54. Howard DM, Adams MJ, Clarke T-K, Hafferty JD, Gibson J et al. 2019. Genome-wide meta-analysis of depression identifies 102 independent variants and highlights the importance of the prefrontal brain regions. Nat. Neurosci. 22:3343–52
    [Google Scholar]
  55. Hultman R, Ulrich K, Sachs BD, Blount C, Carlson DE et al. 2018. Brain-wide electrical spatiotemporal dynamics encode depression vulnerability. Cell 173:1166–80.e14
    [Google Scholar]
  56. Iñiguez SD, Aubry A, Riggs LM, Alipio JB, Zanca RM et al. 2016. Social defeat stress induces depression-like behavior and alters spine morphology in the hippocampus of adolescent male C57BL/6 mice. Neurobiol. Stress 5:54–64
    [Google Scholar]
  57. Joëls M, Pu Z, Wiegert O, Oitzl MS, Krugers HJ. 2006. Learning under stress: How does it work?. Trends Cogn. Sci. 10:4152–58
    [Google Scholar]
  58. Johansen-Berg H, Gutman DA, Behrens TEJ, Matthews PM, Rushworth MFS et al. 2008. Anatomical connectivity of the subgenual cingulate region targeted with deep brain stimulation for treatment-resistant depression. Cereb. Cortex 18:61374–83
    [Google Scholar]
  59. Kang HJ, Adams DH, Simen A, Simen BB, Rajkowska G et al. 2007. Gene expression profiling in postmortem prefrontal cortex of major depressive disorder. J. Neurosci. 27:4813329–40
    [Google Scholar]
  60. Kang HJ, Voleti B, Hajszan T, Rajkowska G, Stockmeier CA et al. 2012. Decreased expression of synapse-related genes and loss of synapses in major depressive disorder. Nat. Med. 18:91413–17
    [Google Scholar]
  61. Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N et al. 2011. Fear erasure in mice requires synergy between antidepressant drugs and extinction training. Science 334:60631731–34
    [Google Scholar]
  62. Karst H, Berger S, Turiault M, Tronche F, Schütz G, Joëls M. 2005. Mineralocorticoid receptors are indispensable for nongenomic modulation of hippocampal glutamate transmission by corticosterone. PNAS 102:5219204–7
    [Google Scholar]
  63. Kasai H, Matsuzaki M, Noguchi J, Yasumatsu N, Nakahara H. 2003. Structure-stability-function relationships of dendritic spines. Trends Neurosci 26:7360–68
    [Google Scholar]
  64. Kavalali ET, Monteggia LM. 2020. Targeting homeostatic synaptic plasticity for treatment of mood disorders. Neuron 106:5715–26
    [Google Scholar]
  65. Kessler RC, Berglund P, Demler O, Jin R, Koretz D et al. 2003. The epidemiology of major depressive disorder: results from the National Comorbidity Survey Replication (NCS-R). JAMA 289:233095–105
    [Google Scholar]
  66. Kim J-W, Autry AE, Na ES, Adachi M, Björkholm C et al. 2021. Sustained effects of rapidly acting antidepressants require BDNF-dependent MeCP2 phosphorylation. Nat. Neurosci. 24:81100–9
    [Google Scholar]
  67. Kim S, Webster MJ. 2011. Integrative genome-wide association analysis of cytoarchitectural abnormalities in the prefrontal cortex of psychiatric disorders. Mol. Psychiatry 16:4452–61
    [Google Scholar]
  68. Kirkby LA, Luongo FJ, Lee MB, Nahum M, Van Vleet TM et al. 2018. An amygdala-hippocampus subnetwork that encodes variation in human mood. Cell 175:61688–700.e14
    [Google Scholar]
  69. Kishimoto T, Chawla JM, Hagi K, Zarate CA, Kane JM et al. 2016. Single-dose infusion ketamine and non-ketamine N-methyl-d-aspartate receptor antagonists for unipolar and bipolar depression: a meta-analysis of efficacy, safety and time trajectories. Psychol. Med. 46:71459–72
    [Google Scholar]
  70. Krishnan V, Han M-H, Graham DL, Berton O, Renthal W et al. 2007. Molecular adaptations underlying susceptibility and resistance to social defeat in brain reward regions. Cell 131:2391–404
    [Google Scholar]
  71. Labonté B, Engmann O, Purushothaman I, Menard C, Wang J et al. 2017. Sex-specific transcriptional signatures in human depression. Nat. Med. 23:91102–11
    [Google Scholar]
  72. Lakshminarasimhan H, Chattarji S. 2012. Stress leads to contrasting effects on the levels of brain derived neurotrophic factor in the hippocampus and amygdala. PLOS ONE 7:1e30481
    [Google Scholar]
  73. Lecca S, Pelosi A, Tchenio A, Moutkine I, Lujan R et al. 2016. Rescue of GABAB and GIRK function in the lateral habenula by protein phosphatase 2A inhibition ameliorates depression-like phenotypes in mice. Nat. Med. 22:254–61
    [Google Scholar]
  74. LeGates TA, Kvarta MD, Tooley JR, Francis TC, Lobo MK et al. 2018. Reward behaviour is regulated by the strength of hippocampus-nucleus accumbens synapses. Nature 564:7735258–62
    [Google Scholar]
  75. Lepack AE, Fuchikami M, Dwyer JM, Banasr M, Duman RS. 2014. BDNF release is required for the behavioral actions of ketamine. Int. J. Neuropsychopharmacol. 18:1pyu033
    [Google Scholar]
  76. Levey DF, Stein MB, Wendt FR, Pathak GA, Zhou H et al. 2021. Bi-ancestral depression GWAS in the Million Veteran Program and meta-analysis in >1.2 million individuals highlight new therapeutic directions. Nat. Neurosci. 24:954–63
    [Google Scholar]
  77. Li B, Piriz J, Mirrione M, Chung C, Proulx CD et al. 2011. Synaptic potentiation onto habenula neurons in the learned helplessness model of depression. Nature 470:7335535–39
    [Google Scholar]
  78. Li K, Zhou T, Liao L, Yang Z, Wong C et al. 2013. βCaMKII in lateral habenula mediates core symptoms of depression. Science 341:61491016–20
    [Google Scholar]
  79. Li N, Lee B, Liu R-J, Banasr M, Dwyer JM et al. 2010. mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science 329:5994959–64
    [Google Scholar]
  80. Lim BK, Huang KW, Grueter BA, Rothwell PE, Malenka RC. 2012. Anhedonia requires MC4R-mediated synaptic adaptations in nucleus accumbens. Nature 487:7406183–89
    [Google Scholar]
  81. Lin P-Y, Ma ZZ, Mahgoub M, Kavalali ET, Monteggia LM. 2021. A synaptic locus for TrkB signaling underlying ketamine rapid antidepressant action. Cell Rep 36:7109513
    [Google Scholar]
  82. Liston C, Chen AC, Zebley BD, Drysdale AT, Gordon R et al. 2014. Default mode network mechanisms of transcranial magnetic stimulation in depression. Biol. Psychiatry 76:517–26
    [Google Scholar]
  83. Liston C, Cichon JM, Jeanneteau F, Jia Z, Chao MV, Gan W-B. 2013. Circadian glucocorticoid oscillations promote learning-dependent synapse formation and maintenance. Nat. Neurosci. 16:6698–705
    [Google Scholar]
  84. Liston C, Gan W-B. 2011. Glucocorticoids are critical regulators of dendritic spine development and plasticity in vivo. PNAS 108:3816074–79
    [Google Scholar]
  85. Liston C, McEwen BS, Casey BJ. 2009. Psychosocial stress reversibly disrupts prefrontal processing and attentional control. PNAS 106:3912–17
    [Google Scholar]
  86. Liston C, Miller MM, Goldwater DS, Radley JJ, Rocher AB et al. 2006. Stress-induced alterations in prefrontal cortical dendritic morphology predict selective impairments in perceptual attentional set-shifting. J. Neurosci. 26:307870–74
    [Google Scholar]
  87. Liu R-J, Duman C, Kato T, Hare B, Lopresto D et al. 2017. GLYX-13 produces rapid antidepressant responses with key synaptic and behavioral effects distinct from ketamine. Neuropsychopharmacology 42:61231–42
    [Google Scholar]
  88. Lu J, Tjia M, Mullen B, Cao B, Lukasiewicz K et al. 2021. An analog of psychedelics restores functional neural circuits disrupted by unpredictable stress. Mol. Psychiatry. https://doi.org/10.1038/s41380-021-01159-1
    [Crossref] [Google Scholar]
  89. Lui S, Wu Q, Qiu L, Yang X, Kuang W et al. 2011. Resting-state functional connectivity in treatment-resistant depression. Am. J. Psychiatry 168:6642–48
    [Google Scholar]
  90. Lupien SJ, McEwen BS, Gunnar MR, Heim C. 2009. Effects of stress throughout the lifespan on the brain, behaviour and cognition. Nat. Rev. Neurosci. 10:6434–45
    [Google Scholar]
  91. Magariños AM, McEwen BS, Flügge G, Fuchs E. 1996. Chronic psychosocial stress causes apical dendritic atrophy of hippocampal CA3 pyramidal neurons in subordinate tree shrews. J. Neurosci. 16:103534–40
    [Google Scholar]
  92. Maras PM, Baram TZ. 2012. Sculpting the hippocampus from within: stress, spines, and CRH. Trends Neurosci 35:5315–24
    [Google Scholar]
  93. Martin DA, Nichols CD. 2018. The effects of hallucinogens on gene expression. Curr. Top. Behav. Neurosci. 36:137–58
    [Google Scholar]
  94. Matsumoto M, Hikosaka O. 2007. Lateral habenula as a source of negative reward signals in dopamine neurons. Nature 447:71481111–15
    [Google Scholar]
  95. Matsuzaki M, Ellis-Davies GC, Nemoto T, Miyashita Y, Iino M, Kasai H. 2001. Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1 pyramidal neurons. Nat. Neurosci. 4:111086–92
    [Google Scholar]
  96. Mayberg HS, Lozano AM, Voon V, McNeely HE, Seminowicz D et al. 2005. Deep brain stimulation for treatment-resistant depression. Neuron 45:5651–60
    [Google Scholar]
  97. McEwen BS. 1998. Protective and damaging effects of stress mediators. N. Engl. J. Med. 338:3171–79
    [Google Scholar]
  98. McEwen BS. 2007. Physiology and neurobiology of stress and adaptation: central role of the brain. Physiol. Rev. 87:3873–904
    [Google Scholar]
  99. McGaugh JL. 2004. The amygdala modulates the consolidation of memories of emotionally arousing experiences. Annu. Rev. Neurosci. 27:1–28
    [Google Scholar]
  100. McGirr A, LeDue J, Chan AW, Xie Y, Murphy TH. 2017. Cortical functional hyperconnectivity in a mouse model of depression and selective network effects of ketamine. Brain 140:82210–25
    [Google Scholar]
  101. Milak MS, Proper CJ, Mulhern ST, Parter AL, Kegeles LS et al. 2016. A pilot in vivo proton magnetic resonance spectroscopy study of amino acid neurotransmitter response to ketamine treatment of major depressive disorder. Mol. Psychiatry 21:3320–27
    [Google Scholar]
  102. Milak MS, Rashid R, Dong Z, Kegeles LS, Grunebaum MF et al. 2020. Assessment of relationship of ketamine dose with magnetic resonance spectroscopy of Glx and GABA responses in adults with major depression: a randomized clinical trial. JAMA Netw. Open 3:8e2013211
    [Google Scholar]
  103. Moda-Sava RN, Murdock MH, Parekh PK, Fetcho RN, Huang BS et al. 2019. Sustained rescue of prefrontal circuit dysfunction by antidepressant-induced spine formation. Science 364:6436aat8078
    [Google Scholar]
  104. Moghaddam B, Adams B, Verma A, Daly D. 1997. Activation of glutamatergic neurotransmission by ketamine: a novel step in the pathway from NMDA receptor blockade to dopaminergic and cognitive disruptions associated with the prefrontal cortex. J. Neurosci. 17:82921–27
    [Google Scholar]
  105. Mühlen K, Ockenfels H. 1968. Morphologische Veränderungen im Diencephalon und Telencephalon nach Störungen des Regelkreises Adenohypophyse-Nebennierenrinde. Z. Zellforsch. Mikrosk. Anat. 93:126–41
    [Google Scholar]
  106. Muir J, Tse YC, Iyer ES, Biris J, Cvetkovska V et al. 2020. Ventral hippocampal afferents to nucleus accumbens encode both latent vulnerability and stress-induced susceptibility. Biol. Psychiatry 88:11843–54
    [Google Scholar]
  107. Murrough JW, Iosifescu DV, Chang LC, Al Jurdi RK, Green CE et al. 2013. Antidepressant efficacy of ketamine in treatment-resistant major depression: a two-site randomized controlled trial. Am. J. Psychiatry 170:101134–42
    [Google Scholar]
  108. Nestler EJ, Hyman SE. 2010. Animal models of neuropsychiatric disorders. Nat. Neurosci. 13:101161–69
    [Google Scholar]
  109. Ofer N, Berger DR, Kasthuri N, Lichtman JW, Yuste R. 2021. Ultrastructural analysis of dendritic spine necks reveals a continuum of spine morphologies. Dev. Neurobiol. 81:5746–57
    [Google Scholar]
  110. Otis JM, Namboodiri VMK, Matan AM, Voets ES, Mohorn EP et al. 2017. Prefrontal cortex output circuits guide reward seeking through divergent cue encoding. Nature 543:7643103–7
    [Google Scholar]
  111. Pare CM, Sandler M. 1959. A clinical and biochemical study of a trial of iproniazid in the treatment of depression. J. Neurol. Neurosurg. Psychiatry 22:247–51
    [Google Scholar]
  112. Phoumthipphavong V, Barthas F, Hassett S, Kwan AC. 2016. Longitudinal effects of ketamine on dendritic architecture in vivo in the mouse medial frontal cortex. eNeuro 3:2ENEURO.0133–15.2016
    [Google Scholar]
  113. Pizzagalli DA, Holmes AJ, Dillon DG, Goetz EL, Birk JL et al. 2009. Reduced caudate and nucleus accumbens response to rewards in unmedicated individuals with major depressive disorder. Am. J. Psychiatry 166:6702–10
    [Google Scholar]
  114. Poo M-M. 2001. Neurotrophins as synaptic modulators. Nat. Rev. Neurosci. 2:124–32
    [Google Scholar]
  115. Post RM, Denicoff KD, Leverich GS, Altshuler LL, Frye MA et al. 2003. Morbidity in 258 bipolar outpatients followed for 1 year with daily prospective ratings on the NIMH life chart method. J. Clin. Psychiatry 64:6680–90
    [Google Scholar]
  116. Radley JJ, Anderson RM, Hamilton BA, Alcock JA, Romig-Martin SA. 2013. Chronic stress-induced alterations of dendritic spine subtypes predict functional decrements in an hypothalamo-pituitary-adrenal-inhibitory prefrontal circuit. J. Neurosci. 33:3614379–91
    [Google Scholar]
  117. Radley JJ, Rocher AB, Miller M, Janssen WG, Liston C et al. 2006. Repeated stress induces dendritic spine loss in the rat medial prefrontal cortex. Cereb. Cortex 16:3313–20
    [Google Scholar]
  118. Radley JJ, Sisti HM, Hao J, Rocher AB, McCall T et al. 2004. Chronic behavioral stress induces apical dendritic reorganization in pyramidal neurons of the medial prefrontal cortex. Neuroscience 125:11–6
    [Google Scholar]
  119. Rajkowska G, Miguel-Hidalgo JJ, Wei J, Dilley G, Pittman SD et al. 1999. Morphometric evidence for neuronal and glial prefrontal cell pathology in major depression. Biol. Psychiatry 45:91085–98
    [Google Scholar]
  120. Rajkowska G, O'Dwyer G, Teleki Z, Stockmeier CA, Miguel-Hidalgo JJ 2007. GABAergic neurons immunoreactive for calcium binding proteins are reduced in the prefrontal cortex in major depression. Neuropsychopharmacology 32:2471–82
    [Google Scholar]
  121. Rosenkranz JA, Venheim ER, Padival M. 2010. Chronic stress causes amygdala hyperexcitability in rodents. Biol. Psychiatry 67:121128–36
    [Google Scholar]
  122. Russo SJ, Nestler EJ. 2013. The brain reward circuitry in mood disorders. Nat. Rev. Neurosci. 14:9609–25
    [Google Scholar]
  123. Salas R, Baldwin P, de Biasi M, Montague PR. 2010. BOLD responses to negative reward prediction errors in human habenula. Front. Hum. Neurosci. 4:36
    [Google Scholar]
  124. Sanacora G, Gueorguieva R, Epperson CN, Wu Y-T, Appel M et al. 2004. Subtype-specific alterations of gamma-aminobutyric acid and glutamate in patients with major depression. Arch. Gen. Psychiatry 61:7705–13
    [Google Scholar]
  125. Sartorius A, Kiening KL, Kirsch P, von Gall CC, Haberkorn U et al. 2010. Remission of major depression under deep brain stimulation of the lateral habenula in a therapy-refractory patient. Biol. Psychiatry 67:2e9–11
    [Google Scholar]
  126. Scangos KW, Makhoul GS, Sugrue LP, Chang EF, Krystal AD 2021. State-dependent responses to intracranial brain stimulation in a patient with depression. Nat. Med. 27:2229–31
    [Google Scholar]
  127. Shabel SJ, Proulx CD, Piriz J, Malinow R. 2014. GABA/glutamate co-release controls habenula output and is modified by antidepressant treatment. Science 345:62031494–98
    [Google Scholar]
  128. Shao L-X, Liao C, Gregg I, Davoudian PA, Savalia NK et al. 2021. Psilocybin induces rapid and persistent growth of dendritic spines in frontal cortex in vivo. Neuron 109:162535–44.e4
    [Google Scholar]
  129. Sheline YI, Barch DM, Donnelly JM, Ollinger JM, Snyder AZ, Mintun MA. 2001. Increased amygdala response to masked emotional faces in depressed subjects resolves with antidepressant treatment: an fMRI study. Biol. Psychiatry 50:9651–58
    [Google Scholar]
  130. Sheline YI, Wang PW, Gado MH, Csernansky JG, Vannier MW. 1996. Hippocampal atrophy in recurrent major depression. PNAS 93:93908–13
    [Google Scholar]
  131. Shirayama Y, Chen AC-H, Nakagawa S, Russell DS, Duman RS. 2002. Brain-derived neurotrophic factor produces antidepressant effects in behavioral models of depression. J. Neurosci. 22:83251–61
    [Google Scholar]
  132. Snyder JS, Soumier A, Brewer M, Pickel J, Cameron HA. 2011. Adult hippocampal neurogenesis buffers stress responses and depressive behaviour. Nature 476:7361458–61
    [Google Scholar]
  133. Soliman F, Glatt CE, Bath KG, Levita L, Jones RM et al. 2010. A genetic variant BDNF polymorphism alters extinction learning in both mouse and human. Science 327:5967863–66
    [Google Scholar]
  134. Spellman T, Svei M, Kaminsky J, Manzano-Nieves G, Liston C. 2021. Prefrontal deep projection neurons enable cognitive flexibility via persistent feedback monitoring. Cell 184:102750–66.e17
    [Google Scholar]
  135. Suzuki K, Nosyreva E, Hunt KW, Kavalali ET, Monteggia LM. 2017. Effects of a ketamine metabolite on synaptic NMDAR function. Nature 546:7659E1–3
    [Google Scholar]
  136. Tye KM, Mirzabekov JJ, Warden MR, Ferenczi EA, Tsai H-C et al. 2013. Dopamine neurons modulate neural encoding and expression of depression-related behaviour. Nature 493:7433537–41
    [Google Scholar]
  137. Vialou V, Feng J, Robison AJ, Nestler EJ. 2013. Epigenetic mechanisms of depression and antidepressant action. Annu. Rev. Pharmacol. Toxicol. 53:59–87
    [Google Scholar]
  138. Vialou V, Robison AJ, Laplant QC, Covington HE 3rd, Dietz DM et al. 2010. ΔFosB in brain reward circuits mediates resilience to stress and antidepressant responses. Nat. Neurosci. 13:6745–52
    [Google Scholar]
  139. Vyas A, Bernal S, Chattarji S. 2003. Effects of chronic stress on dendritic arborization in the central and extended amygdala. Brain Res 965:1–2290–94
    [Google Scholar]
  140. Vyas A, Mitra R, Rao BSS, Chattarji S. 2002. Chronic stress induces contrasting patterns of dendritic remodeling in hippocampal and amygdaloid neurons. J. Neurosci. 22:156810–18
    [Google Scholar]
  141. Vyas A, Pillai AG, Chattarji S. 2004. Recovery after chronic stress fails to reverse amygdaloid neuronal hypertrophy and enhanced anxiety-like behavior. Neuroscience 128:4667–73
    [Google Scholar]
  142. Wang J-W, David DJ, Monckton JE, Battaglia F, Hen R. 2008. Chronic fluoxetine stimulates maturation and synaptic plasticity of adult-born hippocampal granule cells. J. Neurosci. 28:61374–84
    [Google Scholar]
  143. Warden MR, Selimbeyoglu A, Mirzabekov JJ, Lo M, Thompson KR et al. 2012. A prefrontal cortex-brainstem neuronal projection that controls response to behavioural challenge. Nature 492:7429428–32
    [Google Scholar]
  144. Watanabe Y, Gould E, McEwen BS. 1992. Stress induces atrophy of apical dendrites of hippocampal CA3 pyramidal neurons. Brain Res 588:2341–45
    [Google Scholar]
  145. Wellman CL. 2001. Dendritic reorganization in pyramidal neurons in medial prefrontal cortex after chronic corticosterone administration. J. Neurobiol. 49:3245–53
    [Google Scholar]
  146. Whitfield-Gabrieli S, Ford JM. 2012. Default mode network activity and connectivity in psychopathology. Annu. Rev. Clin. Psychol. 8:49–76
    [Google Scholar]
  147. Wilke SA, Lavi K, Byeon S, Donohue KC, Sohal VS. 2022. Convergence of clinically relevant manipulations on dopamine-regulated prefrontal activity underlying stress coping responses. Biol. Psychiatry 91:981020
    [Google Scholar]
  148. Williams LM. 2016. Precision psychiatry: a neural circuit taxonomy for depression and anxiety. Lancet Psychiatry 3:5472–80
    [Google Scholar]
  149. Woolley CS, Gould E, McEwen BS. 1990. Exposure to excess glucocorticoids alters dendritic morphology of adult hippocampal pyramidal neurons. Brain Res 531:1–2225–31
    [Google Scholar]
  150. Wu M, Minkowicz S, Dumrongprechachan V, Hamilton P, Kozorovitskiy Y. 2021a. Ketamine rapidly enhances glutamate-evoked dendritic spinogenesis in medial prefrontal cortex through dopaminergic mechanisms. Biol. Psychiatry 89:111096–105
    [Google Scholar]
  151. Wu M, Minkowicz S, Dumrongprechachan V, Hamilton P, Xiao L, Kozorovitskiy Y. 2021b. Attenuated dopamine signaling after aversive learning is restored by ketamine to rescue escape actions. eLife 10:e64041
    [Google Scholar]
  152. Xia CH, Ma Z, Ciric R, Gu S, Betzel RF et al. 2018. Linked dimensions of psychopathology and connectivity in functional brain networks. Nat. Commun. 9:13003
    [Google Scholar]
  153. Yang Y, Cui Y, Sang K, Dong Y, Ni Z et al. 2018. Ketamine blocks bursting in the lateral habenula to rapidly relieve depression. Nature 554:7692317–22
    [Google Scholar]
  154. Yuen EY, Wei J, Liu W, Zhong P, Li X, Yan Z 2012. Repeated stress causes cognitive impairment by suppressing glutamate receptor expression and function in prefrontal cortex. Neuron 73:5962–77
    [Google Scholar]
  155. Yuste R, Bonhoeffer T. 2001. Morphological changes in dendritic spines associated with long-term synaptic plasticity. Annu. Rev. Neurosci. 24:1071–89
    [Google Scholar]
  156. Zanos P, Moaddel R, Morris PJ, Georgiou P, Fischell J et al. 2016. NMDAR inhibition-independent antidepressant actions of ketamine metabolites. Nature 533:7604481–86
    [Google Scholar]
  157. Zarate CA Jr., Singh JB, Carlson PJ, Brutsche NE, Ameli R et al. 2006. A randomized trial of an N-methyl-d-aspartate antagonist in treatment-resistant major depression. Arch. Gen. Psychiatry 63:8856–64
    [Google Scholar]
  158. Zhang J-Y, Liu T-H, He Y, Pan H-Q, Zhang W-H et al. 2019. Chronic stress remodels synapses in an amygdala circuit-specific manner. Biol. Psychiatry 85:3189–201
    [Google Scholar]
/content/journals/10.1146/annurev-neuro-110920-040422
Loading
/content/journals/10.1146/annurev-neuro-110920-040422
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error