Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Oct 19.
Published in final edited form as: Curr Drug Targets. 2013 Oct;14(11):1225–1236. doi: 10.2174/13894501113149990156

ASTROCYTE PATHOLOGY IN MAJOR DEPRESSIVE DISORDER: INSIGHTS FROM HUMAN POSTMORTEM BRAIN TISSUE

Grazyna Rajkowska a, Craig A Stockmeier a,b
PMCID: PMC3799810  NIHMSID: NIHMS517148  PMID: 23469922

Abstract

The present paper reviews astrocyte pathology in major depressive disorder (MDD) and proposes that reductions in astrocytes and related markers are key features in the pathology of MDD. Astrocytes are the most numerous and versatile of all types of glial cells. They are crucial to the neuronal microenvironment by regulating glucose metabolism, neurotransmitter uptake (particularly for glutamate), synaptic development and maturation and the blood brain barrier. Pathology of astrocytes has been consistently noted in MDD as well as in rodent models of depressive-like behavior. This review summarizes evidence from human postmortem tissue showing alterations in the expression of protein and mRNA for astrocyte markers such as glial fibrillary acidic protein (GFAP), gap junction proteins (connexin 40 and 43), the water channel aquaporin-4 (AQP4), a calcium-binding protein S100B and glutamatergic markers including the excitatory amino acid transporters 1 and 2 (EAAT1, EAAT2) and glutamine synthetase. Moreover, preclinical studies are presented that demonstrate the involvement of GFAP and astrocytes in animal models of stress and depressive-like behavior and the influence of different classes of antidepressant medications on astrocytes. In light of the various astrocyte deficits noted in MDD, astrocytes may be novel targets for the action of antidepressant medications. Possible functional consequences of altered expression of astrocytic markers in MDD are also discussed. Finally, the unique pattern of cell pathology in MDD, characterized by prominent reductions in the density of astrocytes and in the expression of their markers without obvious neuronal loss, is contrasted with that found in other neuropsychiatric and neurodegenerative disorders.

Keywords: Glia, Fronto-limbic, Depression, Glutamate, Postmortem

What is Major Depressive Disorder?

Major depressive disorder (MDD) (also known as clinical depression or unipolar depression) is a chronic, recurrent and debilitating mental illness that affects the lives of millions of people worldwide. Depression is the 3rd leading contributor to the global burden of disease with the prevalence of depression among adults in the United States at 6.7 percent [1, 2]. MDD is characterized by core symptoms such as depressed mood, loss of interest or pleasure, changes in weight, changes in sleep, fatigue or loss of energy, feeling of worthlessness, concentration difficulties and thoughts of death or suicide [3]. Although the neurobiology of MDD has been intensely studied for several decades, its underlying etiopathology is still not fully understood and only about half of individuals with depression respond to currently available treatments [4, 5]. Cellular and molecular abnormalities arising from genetic and environmental factors are believed to be critical in the pathology of depression [6]. A consistent observation of cell pathology in MDD has been reductions in glial cells, and astrocytes, in particular [7].

Introduction to Astrocytes

Of the three kinds of glial cells (i.e., astrocytes, oligodendrocytes, microglia) of the central nervous system (CNS), astrocytes or astroglia are the most numerous and versatile type of glial cells. Rather than serving as static or inert “brain glue” as formerly thought, astrocytes, outnumbering neurons by over fivefold, are now recognized as playing many active roles in the CNS [8-10]. The two major types of astrocytes, protoplasmic and fibrous, are distinguished by their morphology, biochemistry, development and location within CNS [11]. Protoplasmic astrocytes are found in the gray matter. Their numerous processes extend radially from the cell body which contains a spherical nucleus, and give rise to many smaller branches. Some of these processes contact blood vessels to form perivascular endfeet while others cover neuronal membranes. Fibrous astrocytes, in contrast to protoplasmic astrocytes, are present in the white matter. They have ovoid nuclei and fewer, longer and relatively thin processes with very few branches. Their cell bodies are often arranged in rows between the axon bundles and some of their processes reach the perinodal spaces of adjacent axons. Other processes from fibrous astrocytes contact blood vessels to also form perivascular endfeet [12]. Astrocytes are crucial to the neuronal microenvironment by regulating such things as glucose metabolism, neurotransmitter uptake (particularly glutamate), synaptic development and maturation and the blood brain barrier [13]. Studies revealing unexpected pathology in astrocytes in MDD and in rodent models of depressive-like behavior are outlined below. This pathology is unique and differs from reactive astrogliosis described in cases of brain injury, tumor or neurodegenerative disorders.

Human Postmortem Studies

Histopathological studies of postmortem brain tissue reveal prominent glial pathology in MDD. Of the three types of CNS glia, astrocytes are most often implicated as a source of glial pathology in MDD. Several cell counting studies report decreases in the packing density or number of the general, Nissl-stained populations of glial cells in subjects diagnosed with MDD as compared to non-psychiatric control subjects [14-20]. Such changes were observed in fronto-limbic brain regions including the dorsolateral prefrontal cortex [15, 17, 19], orbitofrontal cortex [15], subgenual cortex [14] anterior cingulate cortex [16, 20] and amygdala [18]. However, Khundakar et al. [21, 22] noted no change in glial density in the orbitofrontal cortex or anterior cingulate cortex in elderly subjects with MDD.

In addition to alterations in glial cell density and number, the average size of nuclei of glial cells was increased in the gray matter of dorsolateral prefrontal cortex in MDD [15]. In contrast, one other study in the dorsolateral prefrontal cortex reported no change in glial size in MDD [17]. A recent detailed analysis of astrocytes stained with the Golgi method revealed hypertrophy of astrocytic cell bodies and processes in the white matter of the anterior cingulate cortex in depressed subjects dying by suicide [23]. The authors interpret astrocytic hypertrophy as a reflection of local inflammation in support of the neuroinflammatory theory of depression [24].

Markers of Astrocytes

Astrocytes have been localized in brain tissue by antibodies to a number of proteins. These astrocytic markers include glial fibrillary acidic protein (GFAP), gap junctions proteins such as connexin 40 and 43, the water channel aquaporin-4 (AQP4), a calcium-binding protein S100B and the glutamatergic markers including the excitatory amino acid transporters 1 and 2 (EAAT1, EAAT2) and glutamine synthetase. Evidence is presented below from human postmortem studies demonstrating changes in each of these astrocytic markers in MDD.

GFAP

Glial fibrillary acidic protein (GFAP) is the principle component of cytoskeletal intermediate filaments that is strongly expressed in the CNS by mature and reactive astrocyte cells [25, 26]. Eight isoforms of GFAP have been identified to mark specific subpopulations of astrocytes in the human brain during development, aging and disease, however the function of the specific isoforms is not fully understood [26]. GFAP is expressed in astrocyte processes and cell bodies and it is thought this protein helps astrocytes to maintain mechanical strength and shape [27, 28]. Moreover, GFAP is involved in processes related to cell movement and structure and has been proposed to play a role in astrocyte-neuron communication [29, 30]. Antibodies to GFAP, an astrocyte-specific protein, can be used to immunohistochemically distinguish astrocytes from other types of glial cells [10]. However, antibodies to GFAP only identify about 15–20 percent of astrocytes expressed in the cortex of mature animals [26].

The expression of GFAP in gray matter has been quantified in depression by assessing GFAP-immunoreactive (IR) astrocyte density or the area covered by GFAP-IR cell bodies and processes, so called area fraction. There was a significant decrease in the density of GFAP-IR astrocytes and GFAP area fraction in gray matter of the dorsolateral prefrontal cortex in younger depressed subjects (<60 years old) as compared to age-matched non-psychiatric control subjects [31]. In addition, GFAP-IR area fraction is significantly decreased in the gray matter of the orbitofrontal cortex in a mixture of younger and older subjects with MDD [32]. In contrast, older subjects with late-onset depression showed increases in GFAP-IR area fraction and cell density in the gray matter of dorsolateral prefrontal cortex [31, 33], which may reflect a compensatory reaction to neuronal damage reported in older subjects with MDD [34]. Thus, the pattern of astrocyte pathology in cortical gray matter is unique in younger vs. older subjects with depression [7, 35, 36]. In a semi-quantitative study, Muller et al. [37] detected a significant decrease in GFAP-IR astrocytes in the CA1 and CA2 subregions of the hippocampus in depression. A similar decrease in GFAP-IR astrocytes was noted in subjects that had been treated with steroids, suggesting that elevated glucocorticoid hormones acting at glucocorticoid receptors on astrocytes may have contributed to the reduction in GFAP expression in astrocytes [37, 38]. In a 3-dimensional quantitative study, a significant reduction in the density of GFAP-IR astrocytes was recently observed in the hilus of the hippocampus in subjects with MDD not treated with antidepressant medications (Stockmeier et al., personal communication). The observations of pathological changes in GFAP in MDD in gray matter of dorsolateral prefrontal cortex and orbitofrontal cortex have been extended to white matter in other cortical regions. GFAP immunolabeling was decreased in MDD in white matter of the anterior cingulate cortex [39] and the orbitofrontal cortex (Rajkowska et al., unpublished observations).

Expression of GFAP protein and mRNA has also been examined in MDD. Levels of GFAP protein were decreased in gray matter from the dorsolateral prefrontal and orbitofrontal cortex in MDD [32, 40]. An earlier proteomic study reported decreases in four isoforms of GFAP in the frontal cortex of subjects with MDD [41]. In agreement with protein expression studies, GFAP mRNA was also under-expressed in the white matter of the anterior cingulate cortex in MDD [42]. We recently observed a decrease in the expression of mRNA for GFAP in white and gray matter of the orbitofrontal cortex in subjects with MDD (Newton and Rajkowska, unpublished observations). In summary, reductions in the density and area fraction of GFAP-IR astrocytes and in the levels of GFAP protein and mRNA reveal dysfunctional astrocytes in MDD in fronto-limbic cortical regions. Moreover, reduced density of astrocytes may account, at least in part, for volumetric reductions reported by neuroimaging studies in the prefrontal cortex of depressed patients [43-47].

Subcortical brain regions also show a reduction in the expression of astrocyte-associated genes and protein. In locus coeruleus, limbic thalamic nuclei, putamen and the internal capsule from subjects with MDD, there was a reduction in glia-associated genes expression [48, 49]. Moreover, Chandley et al. [50] recently observed a reduction in expression of GFAP mRNA and protein and a decrease in the density of GFAP-IR astrocytes in the locus coeruleus in MDD. Other astrocyte-associated genes were also down-regulated in MDD in astrocytes that were laser-dissected from locus coeruleus [50, 51]. Finally, significant decreases in GFAP protein expression and GFAP-IR astrocyte density were noted in the cerebellum [52] and amygdala [53] in MDD. Thus, a number of subcortical brain areas in addition to fronto-limbic cortical regions reveal pathology of astrocytes in MDD.

Preclinical studies provide evidence for the involvement of GFAP and astrocytes in animal models of depressive-like behavior. Various types of stress cause reductions in measures of GFAP-IR astrocytes. For example, the stress of separating juveniles from their family diminished the density of GFAP-IR astrocytes in the rodent medial prefrontal cortex [54]. The stress of chronic social defeat reduced the number and soma volume of GFAP-IR astrocytes in the hippocampus of tree shrews [55] and decreased the level of GFAP protein in rat hippocampus [56]. Early life stress also resulted in a reduced density of GFAP-IR astrocytes in adult rats in various prefrontal and frontal cortical regions, hippocampus and the basolateral amygdala [57]. Furthermore, chronic mild stress significantly decreased levels of GFAP mRNA in rat medial prefrontal cortex [58]. Interestingly, infusion of L-α aminoadipic acid in rodent prefrontal cortex, thought to selectively lesion glial cells including GFAP-IR astrocytes but not neurons, induced depressive-like behaviors [59, 60]. These two lesion studies appear to support the hypothesis that the loss of glia contributes to the pathology of depression [7], although this conclusion rests on the specificity of the toxin to glia. There is also correlative support for a role of GFAP-IR astrocytes in depressive-like behavior in Wistar-Kyoto rats, a strain of rats that is genetically predisposed to anxiety-like and depressive-like behavior [61]. Significant reductions in the density of GFAP-IR astrocytes but not NeuN-IR neurons were observed in the prefrontal cortex, anterior cingulate cortex, amygdala and hippocampus in Wistar-Kyoto rats as compared to control Spraque-Dawley rats [62]. Thus, specific astrocytic deficits in the expression of GFAP in corticolimbic circuits are associated with depressive-like behavior.

Astrocytes have been suggested as a target for therapeutic interventions in depression. Several animal studies reveal an influence of different classes of antidepressant medications on astrocytes. Treatment with fluoxetine, a serotonin-selective reuptake inhibitor (SSRI), prevented the psychosocial stress-induced reduction in astrocyte number [55]. Riluzole, a glutamate modulating drug, also prevented the chronic, mild stress-induced reduction in the expression of GFAP mRNA in the rat prefrontal cortex [58]. The beneficial effects of the SSRI antidepressants citalopram and fluoxetine may involve their ability to induce calcium signals in astrocytes in the prefrontal cortex [63]. However, not all studies show reversibility of number of astrocytes or GFAP level by antidepressants. For example, a four-week treatment with citalopram, also an SSRI, did not restore the social defeat-induced reduction in GFAP protein in the rat hippocampus, although the behavior of the animals was normalized within this treatment period [56]. Likewise, imipramine, a tricyclic antidepressant drug, could not reverse the effects of learned helplessness on hippocampal astrocytes [64].

Other effective treatments for depression such as electroconvulsive therapy (ECT) and transcranial magnetic stimulation can also alter the expression of GFAP in animal models. Repeated treatment of rats with electroconvulsive shock (ECS) increased the expression of GFAP but not neuron specific enolase in limbic brain regions including the piriform cortex, amygdala and hippocampus [65]. Chronic ECS in rats also increases proliferation of astrocytes and changes the morphology of astrocytes and other types of CNS glia in limbic brain regions including hippocampus, piriform cortex and prefrontal cortex [66]. A single ECS or transcranial magnetic stimulation up-regulated mRNA for GFAP selectively in the dentate gyrus of the mouse hippocampus [67, 68]. Thus, ECT and transcranial magnetic stimulation therapy can modulate expression of the gene for GFAP and therefore may account for some features of astrocytic plasticity in the action of antidepressant treatments of depression.

In summary, models of chronic stress in experimental animals significantly diminish cortical and hippocampal astrocytes as measured with GFAP while lesions of cortical glia, including astrocytes, yield behavioral deficits comparable to those seen following chronic stress. The effects of chronic stress on GFAP-IR astrocytes can be reversed by chronic treatment with antidepressant medications. Upregulation in GFAP mRNA can also be induced by ECS or transcranial magnetic stimulation. Thus, in light of astrocyte deficits noted in MDD and stress being a risk factor for depression, astrocytes may be potential novel targets for the action of antidepressant medications [69].

Connexins

Connexin 30 and connexin 43 are two other proteins located on astrocytic endfeet that implicate these glial cells in the pathology of depression. Connexin 30 and connexin 43 are gap junction-forming membrane proteins that allow communication between astrocytes [70]. Decreases in the gene and protein expression of connexin 30 and connexin 43 have been observed in the dorsolateral prefrontal cortex of suicide completers, some of whom were diagnosed with MDD [71]. In the orbitofrontal cortex, a decrease in the expression of connexin 43 protein was noted in MDD (Miguel-Hidalgo, personal communication). In addition, down-regulation of genes for connexin 43 and 40 was reported in the locus coeruleus of subjects with MDD [49]. Moreover, rats exposed to chronic unpredictable stress exhibited depressive-like behavior and abnormal morphology and functions of gap junctions as well as a significant decrease in the expression of connexin 43 [72]. These stress-induced changes were reversed by chronic treatment with the antidepressant medications fluoxetine or duloxetine. In another study, connexin 43 was also up-regulated in rat prefrontal cortex after chronic exposure to fluoxetine [73]. Taken together, these animal studies support the involvement of connexin in depression and the therapeutic action of antidepressant medications. The consequences of decreased expression of connexin 30 and connexin 43 may alter calcium wave propagation and communication between astrocytes [64, 74]. Finally, mice lacking connexin 30 and connexin 43 had a weakened blood-brain barrier, in which astrocytes are a crucial component [75].

AQP4

A recent study has implicated another protein, aquaporin-4 (AQP4), in the pathophysiology of MDD [76]. AQP4 is located predominantly in astrocytic endfeet that are in contact with blood vessels [77, 78]. AQP4 is a water channel that regulates water and ion homeostasis in the brain and is an integral part of the neurovascular unit [79-81]. Reduced coverage of blood vessels by astrocytic endfeet that are immunopositive for AQP4 has been observed in the orbitofrontal cortex in subjects with MDD as compared to psychiatrically-normal control subjects [76]. In addition, a decrease in the expression of mRNA for AQP4 was identified in locus coeruleus in MDD [49]. These decreases in AQP4 in MDD could affect many brain functions as AQP4, in addition to its role in water redistribution, also regulates cerebral blood flow [82, 83], glucose transport and metabolism [84], integrity of the blood brain barrier [85, 86], glutamate turnover [87] and synaptic plasticity [88]. Thus, AQP4 may serve as a marker of astrocytic pathology in MDD.

S100B

Another marker of astrocytes is the calcium binding protein S100B [89]. In the gray matter, this protein is predominantly expressed and secreted by astrocytes, while oligodendrocytes expressed S100B in white matter [90-92]. Levels of expression of S100B mRNA and protein, both intracellular and extracellular, may signify astrocyte reaction or death in several types of brain injury [93-96]. Altered expression of S100B may also be present in mood disorders. A transcript for S100B was down-regulated in the ventral prefrontal cortex of depressed suicide victims [97].

Damage to astrocytes causes leakage of S100B into the extracellular compartment and cerebrospinal fluid, continuing to the bloodstream [98]. In vivo support for damage to astrocytes in depression is provided by reports of elevated levels of S100B in the cerebrospinal fluid and serum during major depressive or manic episodes [98-103]. In contrast, six weeks of successful antidepressant treatment lowered serum level of S100B [103]. In the study by Schroeter et al. [103], both the astroglial marker S100B and neuron-specific enolase, a marker specific for neurons, were measured. Only S100B was elevated with no change noted in neuron-specific enolase in mood disorders. Thus, the in vivo studies on S100B support our original hypothesis based on postmortem brain tissue that depression is characterized by prominent glial and only subtle neuronal pathology [7].

Glutamate Markers

We extensively review studies of depression which are specifically related to glutamate markers located on or in astrocytes. While astrocytes have transporters and receptors for other neurotransmitters, the role of astrocytes in glutamate neurotransmission has been studied much more extensively in depression using postmortem brain tissue than the role of astrocytes in other neurotransmitter systems. Thus, our review is focused on the role of astrocytes in glutamate neurotransmission in depression.

Astrocytes are actively involved in the uptake, metabolism and recycling of glutamate. Extracellular levels of glutamate are regulated by removal of this neurotransmitter from the synaptic cleft via specialized transporters located on astrocytic processes [104]. In human brain these glutamate transporters include the excitatory amino acid transporter-1 and -2 (EAAT1 and EAAT2) which in rodents are known as glutamate–aspartate transporter (GLAST) and glutamate transporter 1 (GLT1), respectively [105, 106]. The internalized glutamate is subsequently converted within astrocytes into glutamine by the enzyme, glutamine synthetase [107]. Glutamine leaves astrocytes to be taken up by neurons where it can be reconverted into glutamate or GABA. Thus, astrocytes play a critical role in several aspects of glutamate neurotransmission.

Glutamate transporters and glutamine synthetase associated with astrocytes may serve as markers of astrocyte function and these markers appear to be dysregulated in postmortem brain tissue from subjects with MDD. For example, reduced expression of mRNA for EAAT1, EAAT2 and glutamine synthetase was noted in the anterior cingulate and dorsolateral prefrontal cortex in postmortem brain samples from subjects with MDD [108]. Expression of the mRNA for glutamine synthetase was also down-regulated in the dorsolateral prefrontal cortex, premotor cortex and the amygdala of depressed suicide victims [109]. Moreover, the expression of EAAT1, EAAT2 and glutamine synthetase protein was also reduced in the orbitofrontal cortex in immunohistochemical and Western blotting studies of subjects with MDD [32]. Finally, glutamate signaling and astrocyte-associated genes were under-expressed in locus coeruleus in MDD [49, 50, 51], suggesting more global dysfunction of glutamate signaling and astrocyte pathology in MDD. Support for disease-specific astroglial pathology in MDD comes from Bernard et al. [49] demonstrating that these changes in glutamate-related gene expression do not occur in neurons. Other evidence supporting a role for dysregulated astrocytic glutamate uptake in depression comes from rat studies where the pharmacological blockade of glutamate uptake into astrocytes in the amygdala [110], ventral tegmental area [111] or in the prefrontal cortex [112] is sufficient to decrease sucrose consumption, a behavioral marker thought to be related to anhedonia, a core symptom of depression. Finally, animal studies reveal that astrocytic GFAP plays a key role in the trafficking of glutamate transporters and protecting the brain against glutamate-mediated excitotoxicity [113, 114].

Neuroimaging studies reveal alterations in glutamate in MDD that may be related to cellular changes detected in postmortem brain tissue. Levels of glutamate and glutamine or combined glutamate/glutamine (Glx) are significantly decreased in living subjects with MDD, as determined in plasma [115] and several brain regions including prefrontal [116, 117] and anterior cingulate cortex [118-120], and the combined region of amygdala plus anterior hippocampus [121, 122]. However, one imaging study and one study in postmortem brain tissue reported increases in glutamate levels in the occipital cortex and frontal cortex, respectively, in MDD [123, 124]. These differing reports of glutamate levels in depression may be due to variations in the brain region examined, relative involvement of the hypothalamic-pituitary-adrenal axis, and the age of the subject and chronicity of depression.

The sequence of events relating glial pathology with changes in glutamate levels is not clearly understood. Based on studies outlined above, one hypothesis suggests that elevated levels of glucocorticoids, which significantly decrease expression of mRNA for GFAP [125], reduce the density of astrocytes early in the course of depression [7]. As a result, extracellular glutamate accumulates due to fewer glutamate transporters EAAT1 and EAAT2 located on astrocytes. These initially high levels of extracellular glutamate may also be toxic to GABA neurons [126, 127]. Diminished astrocytic uptake of glutamate, and its reduced subsequent conversion to glutamine and eventually to glutamate, may be related to reduced density of glutamatergic pyramidal neurons in elderly depressed subjects [34] and be responsible for depleted cortical levels of glutamate as depression progresses.

Functions of Astrocytes and Consequences of Their Pathology in MDD

In addition to astrocyte functions described above, astrocytes are involved in other higher brain functions. For example, astrocytes actively control neuronal activity and synaptic neurotransmission They are an active component of the tripartite synapse, including 1) the presynaptic terminal, 2) the postsynaptic neuronal membrane and 3). the surrounding astrocyte [128, 129]. Within the synapse, astrocytes respond to neuronal activity by elevating their internal Ca2+ concentration that triggers the release of glial transmitters which, in turn, regulate neuronal activity [128, 129]. Astrocytes are also crucial in promoting the development and modeling of synapses [130-133]. Correspondingly, the density of synapses and expression of several synapse-related genes were significantly decreased in the dorsolateral prefrontal cortex in the same depressed subjects in which astrocyte density was decreased [31, 134].

Astrocytes also modulate glucose metabolism, as well as neuronal activity and synaptic density. Astrocytic endfeet enveloping cerebral microvasculature are enriched with glucose transporters. Glucose, the main energy substrate for neurons and glia, is taken up by the astrocytic glucose transporter. In astrocytes, glucose undergoes glycolysis and oxidative phosphorylation and this is believed to provide the observed signal in functional magnetic resonance imaging [135, 136]. In positron emission tomography (PET), the uptake of radioactive 18F-deoxyglucose by astrocytes is associated with neuronal activity [137]. A well-replicated finding in the PET literature is that of reduced glucose metabolism in prefrontal cortex in MDD (reviewed in [138]). Astrocytes are also able to control blood flow by inducing local vasoconstriction or vasodilation [139-141]. Thus, defective or missing astrocytes, as reported in studies of postmortem brain tissue from subjects with MDD, may significantly contribute to the changes in cerebral blood flow in depressed patients as noted with PET and functional magnetic resonance imaging (fMRI).

Astrocytes regulate synaptic and intracellular levels of neurotransmitters via transporters located on their processes. In addition to the glutamate transporters described above, astrocytes also have transporters for GABA [142-144], serotonin [145, 146], dopamine [147, 148], norepinephrine [147-149] and histamine [150] and they contain the enzyme monoamine oxidase (MAO) [151], which further regulates the concentrations of the monoamine neurotransmitters. Abnormally elevated levels of MAO-A have been reported in patients with MDD [152-154]. In postmortem prefrontal cortex from subjects with MDD, an elevated expression of MAO-A protein is coupled with a reduction in expression of R1, a novel protein that represses expression of MAO-A [155]. Thus, astrocytes may be a significant cellular substrate for the above mentioned pathology of MAO in depression. Astrocytes express nearly all of the receptors and ion channels found in neurons [156-165], although no specific pathologies of these astrocytic receptors have been linked to depression.

Astrocytes also play a role in inflammatory and neurodegenerative processes in the brain. Reactive astrocytes, together with microglia and oligodendrocytes, are both source and target of various inflammatory cytokines such as interleukins and tumor necrosis factor. These cytokines are expressed under pathological conditions and involved in the regulation of neuroinflammation, immune processes and tissue repair [166]. Peripheral cytokines may serve as biomarkers of MDD and the response to antidepressant treatment, and their entry into the brain may induce production of cytokines by astrocytes [167]. For example, in MDD, pro-inflammatory markers in plasma such as C-reactive protein (CRP), interleukin-1 (IL-1), interleukin-6 (IL-6) and tumor necrosis factor alpha (TNF-α) were increased in depressed patients as compared to control subjects [168-171]. Protein and mRNA levels of IL-1, IL-6 and TNF-α in prefrontal cortex were also significantly increased in teenage suicide victims with a mood and/or psychoactive substance use disorder [172]. Treatment with antidepressant medications (SSRIs) reduced plasma levels of inflammatory cytokines in patients with MDD [168, 173-177].

Astrocytes, in addition to microglia, also respond to neuronal injury. The astrocytic reaction to injury involves increasing their number, size and expression of GFAP [178, 179]. However, in direct contrast, astrocytic pathology in depression is characterized by a decrease in astrocyte density and expression of GFAP. Thus, MDD is viewed primarily as a disease of disrupted neuroplasticity and cellular resilience rather than cell loss [180, 181]. Evidence for impaired neuroplasticity in depression is provided by studies using postmortem brain tissue and by preclinical models reporting a reduction in protein and mRNA expression of neurotrophic factors and their receptors such as brain-derived neurotrophic factor (BDNF) and fibroblast growth factor 2 (FGF2) [182-190]. Conversely, repeated treatment with antidepressant medication up-regulates expression of BDNF and FGF2 factors in humans and rodents [183, 191-193]. The reduction in neurotrophic factors in depression may be occurring in astrocytes, in addition to neurons, as astrocytes also secrete neurotrophic factors [194]. Cultured cortical astrocytes treated with SSRIs showed an increase in expression of BDNF and other neurotrophic factors [195]. Thus, astrocytic pathology may significantly contribute to the disrupted neuroplasticity of depression.

Another nerve growth factor implicated in depression and the mechanism of action of antidepressant medications is a glial cell line-derived neurotrophic factor (GDNF). Several studies reveal that this protein was under-expressed in the peripheral blood of depressed patients [196-198], while GDNF levels in hippocampus were significantly decreased in the rodent model of chronic unpredictable stress [199]. In rodents, the chronic administration of electroconvulsive shock but not antidepressant medications increased levels of the GDNF receptors (GFRalpha-1 and GFRalpha-2) in hippocampus, while repeated chlomipramine treatment reversed the chronic stress-induced decrease in hippocampal GDNF [199, 200]. In subjects with MDD, serum levels of GDNF were significantly increased by chronic treatment with antidepressant medications or in responders to electroconvulsive therapy [196, 201]. One study has examined GDNF directly in postmortem brain tissue in a small number of subjects with recurrent or reactive depression [202]. Several cortico-limbic regions were examined, but there was a paradoxical increase in the expression of GDNF protein in only the parietal cortex. Additional studies need to be performed on GDNF and its receptors in postmortem human brain of subjects experiencing recurrent depression or major depressive disorder to determine whether brain levels correspond to plasma levels of GDNF in depression.

Unique Pattern of Glial and Neuronal Cell Pathology in MDD

As described above, MDD is a disorder with a prominent pathology of astroglia. However, the pattern of astrocyte pathology in MDD is very different from that observed in several neurological and neurodegenerative disorders such as CNS injury, brain tumors, CNS inflammation, stroke, epilepsy, amyotrophic lateral sclerosis, Huntington’s disease, Parkinson’s disease or Alzheimer’s disease. In these disorders, reactive astrogliosis, glial scar formation and neuronal degeneration take place [10]. In contrast to neurodegenerative disorders where there is astrogliosis and an increased in expression of GFAP, in MDD there is no astrogliosis, the expression of GFAP and other markers of astrocytes is decreased and no prominent neuronal pathology is observed. Thus, the neuropathology of depression is significantly different from that observed in brain injury or neurodegeneration.

There are several lines of evidence pointing to the preservation of neurons in MDD. Most cell counting studies in postmortem brain tissue did not find reductions in neuronal density or total number of neurons in subjects with MDD [14, 16, 17, 21, 22, 203-205]. Rather, smaller sizes of neuronal cell bodies or reductions in dendritic branching are reported [15-17, 19, 203, 206, 207] suggesting neuronal atrophy rather than neuronal loss in MDD. In addition, in vivo studies measuring brain metabolite N-acetylaspartate (NAA), a putative marker of neurons revealed no changes in the concentration of this marker in the frontal lobes and basal ganglia in depressed patients further suggesting no neuronal loss in depression [208]. In another study measuring serum levels of neuron-specific enolase, a specific marker of neurons in the CNS, there were no changes in patients with MDD compared to control patients [103, 209]. In the same study of depressed patients, there was an increase in levels of serum S100B, a marker of astrocytes. The expression of mRNA for the neuronal markers neurofilament and enolase2 was not significantly changed in the locus coeruleus of subjects with MDD, whereas mRNA for glial markers (glial glutamate transporters and glutamine synthetase) were significantly reduced as compared to control subjects [49]. Interestingly, a neuron-specific toxin injected into the rodent prefrontal cortex caused no change in behavior while ablation of glia with a glial-specific toxin led to depressive-like behaviors [59]. In summary, cellular pathology in MDD appears to be selective for glia without obvious neuronal loss. It is noteworthy that such pathology appears to apply mostly to younger and middle age subjects with MDD (<60 years-of-age) [31, 35, 40]. In older subjects with MDD (>60 years-of-age), neuronal pathology consisted of prominent reductions in the density of pyramidal glutamatergic neurons in the orbitofrontal cortex [34, 35]. In contrast, astrocytes density and GFAP levels are unaltered in elderly depressed compared to age-matched control subjects [31, 33, 40]. In light of such age-related cell pathology in MDD, we hypothesized that glial (and astrocytic, in particular) reductions take place early in depression and neuronal pathology occurs later in the progression of the disease [7]. Neuronal pathology may be induced by excitotoxicity due to an excess of extracellular glutamate building up in synaptic cleft due to reduced numbers of astrocytes and astrocytic glutamate transporters. Later in the course of depression, astrocytes may react to neuronal pathology by maintaining their density with age in the face of decreased neuronal density. Support for this hypothesis comes from our observation of no significant reductions in astrocyte density or expression of GFAP in elderly subjects with depression (for further details on this hypothesis see [7]).

Astrocyte Pathology in Other Neuropsychiatric Disorders

Although glial pathology has been noted in fronto-limbic cortex in several psychiatric disorders, the specific features of that pathology are unique in each disorder. In postmortem brain tissue of subjects diagnosed with MDD, a large number of markers of astrocytes are consistently altered in fronto-limbic cortical areas and related subcortical regions. Astrocytic proteins such as GFAP, AQP4, connexins, glutamate transporters (EAAT1 and EAAT2) and glutamine synthetase are all decreased in MDD.

In bipolar disorder, however, studies of postmortem brain tissue reveal mixed results with respect to GFAP and glutamine synthetase. One study reported a decrease in the optical density of GFAP-IR across all layers of the orbitofrontal cortex in bipolar disorder [210]. In contrast, another study found no significant changes in the density of astrocytes immunoreactive for GFAP in the gray or white matter of the subgenual cingulate cortex [211], while the expression of mRNA for GFAP was decreased in the adjacent dorsal portion of the anterior cingulate cortex in subjects with bipolar disorder [42]. However, expression of glutamine synthetase, mostly localized to astrocytes, was not significantly changed in the dorsolateral and orbitofrontal cortex in bipolar subjects [210] while the expression of mRNA and protein for glutamine synthetase was significantly decreased in these two regions in MDD [32,108].

In schizophrenia, relatively few studies have examined astrocytes and have yielded inconsistent results again as with bipolar disorder. The density of astrocytes immunoreactive for GFAP was not significantly changed in schizophrenia in few cortical and subcortical regions [212-214], while other studies noted a significant decrease in the density of these cells or mRNA expression for GFAP in the dorsolateral prefrontal, orbitofrontal and subgenual cingulate cortex [42, 210, 211, 215]. In contrast, Toro et al. [210] reported a significant increase in the GFAP-IR in dorsolateral prefrontal cortex in schizophrenia. In contrast to MDD, immunoreactivity for glutamine synthetase was not significantly affected in dorsolateral prefrontal or orbitofrontal cortex in schizophrenia [32, 210].

Alcohol dependence is another psychiatric disorder, whether comorbid with MDD or not, that is characterized by reductions in GFAP immunoreactivity and density of GFAP-IR astrocytes [216, 217]. However, in contrast to reductions in MDD, immunoreactivity for glutamatergic markers in astrocytes (glutamine synthetase, EAAT1 and EAAT2) was not significantly affected in orbitofrontal cortex in alcohol dependence [32].

The pattern of astrocyte cell pathology differs between neurological and depression disorders. In neurological disorders there is a loss of neurons with an accompanying proliferation of astroglia. On the other hand, in MDD there is prominent reduction in astroglial density. Thus, MDD should not be considered as a neurodegenerative disorder but rather a disorder of disrupted neuroplasticity and cellular resilience.

Concluding Remarks

Astrocyte pathology in MDD is well documented by a number of quantitative studies on postmortem fronto-limbic brain regions. There are consistent reductions in the density of astrocyte cell bodies immunoreactive for GFAP protein in MDD. Proteins expressed by astrocytes, such as GFAP, AQP4, connexins, glutamate transporters (EAAT1 and EAAT2) and glutamine synthetase are also decreased in MDD. Although the extent of astrocytic pathology in MDD has been examined for several years, accessible molecular targets to mitigate their obvious dysfunction have not been identified. The induction of depressive-like behaviors in rodents by lesions to astrocytes in prefrontal cortex may be evidence that astrocytic pathology in humans is crucial in the development of depression. Astrocyte pathology similar to that observed in MDD was noted in rodent models of chronic stress and depressive-like behaviors and repeated treatment with antidepressant medications reversed the morphological and behavioral changes observed after exposure to chronic stress. Thus, astrocytes may be a therapeutic target for novel and more effective antidepressant treatments. Further studies on the reciprocal communication between astrocytes and neurons are needed to better understand the effects of astrocyte pathology on neurons in MDD.

Acknowledgements

The authors acknowledge support from the National Institutes of Health (NCRR RR17701). The authors report no biomedical, financial interests or potential conflicts of interest.

References

  • [1].World Health Organization . The Global Burden of Disease 2004 Update. [Google Scholar]
  • [2].Kessler RC, Chiu WT, Demler O, Walters EE. Prevalence, severity, and comorbidity of twelve-month DSM-IV disorders in the National Comorbidity Survey Replication (NCS-R) Archives of General Psychiatry. 2005;62:617–27. doi: 10.1001/archpsyc.62.6.617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].American Psychiatric Association . Diagnostic and Statistical Manual of Mental Disorders. 4th ed. Text Revision American Psychiatric Association; Washington, DC: 2000. [Google Scholar]
  • [4].Trivedi MH, Rush AJ, Wisniewski SR, et al. STAR*D Study Team 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. doi: 10.1176/appi.ajp.163.1.28. [DOI] [PubMed] [Google Scholar]
  • [5].Krishnan V, Nestler EJ. Linking molecules to mood: new insight into the biology of depression. Am J Psychiatry. 2010;167:1305–20. doi: 10.1176/appi.ajp.2009.10030434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Krishnan V, Nestler EJ. The molecular neurobiology of depression. Nature. 2008;455:894–902. doi: 10.1038/nature07455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Rajkowska G, Miguel-Hidalgo JJ. Gliogenesis and glial pathology in depression. CNS Neurol Disord Drug Targets. 2007;6:219–33. doi: 10.2174/187152707780619326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Volterra A, Meldolesi J. Astrocytes, from brain glue to communication elements: the revolution continues. Nature Rev Neurosci. 2005;6:626–40. doi: 10.1038/nrn1722. [DOI] [PubMed] [Google Scholar]
  • [9].Allen NJ, Barres BA. Neuroscience: glia—more than just brain glue. Nature. 2009;457:675–77. doi: 10.1038/457675a. [DOI] [PubMed] [Google Scholar]
  • [10].Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119:7–35. doi: 10.1007/s00401-009-0619-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Miller RH, Raff MC. Fibrous and protoplasmic astrocytes are biochemically and developmentally distinct. J Neurosci. 1984;4:585–92. doi: 10.1523/JNEUROSCI.04-02-00585.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Reichenbach A, Wolburg H. In: Neuroglia. 2nd ed Kettenmann H, Ransom BR, editors. Oxford University Press; New York: 2005. pp. 19–35. [Google Scholar]
  • [13].Kettenmann H, Ranson BR. Neuroglia. 2nd ed. Oxford University Press; New York: 2005. [Google Scholar]
  • [14].Ongür D, Drevets WC, Price JL. Glial reduction in the subgenual prefrontal cortex in mood disorders. Proc Natl Acad Sci USA. 1998;95:13290–5. doi: 10.1073/pnas.95.22.13290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Rajkowska G, Miguel-Hidalgo JJ, Wei J, et al. Morphometric evidence for neuronal and glial prefrontal cell pathology in major depression. Biol Psychiatry. 1999;45:1085–98. doi: 10.1016/s0006-3223(99)00041-4. [DOI] [PubMed] [Google Scholar]
  • [16].Cotter D, Mackay D, Landau S, Kerwin R, Everall I. Reduced glial cell density and neuronal size in the anterior cingulate cortex in major depressive disorder. Arch Gen Psychiatry. 2001;58:545–53. doi: 10.1001/archpsyc.58.6.545. [DOI] [PubMed] [Google Scholar]
  • [17].Cotter D, Mackay D, Chana G, Beasley C, Landau S, Everall IP. Reduced neuronal size and glial cell density in area 9 of the dorsolateral prefrontal cortex in subjects with major depressive disorder. Cereb Cortex. 2002;12:386–94. doi: 10.1093/cercor/12.4.386. [DOI] [PubMed] [Google Scholar]
  • [18].Bowley MP, Drevets WC, Ongür D, Price JL. Low glial numbers in the amygdala in major depressive disorder. Biol Psychiatry. 2002;52:404–12. doi: 10.1016/s0006-3223(02)01404-x. [DOI] [PubMed] [Google Scholar]
  • [19].Torres-Platas D, Mackay D, Chana G, Beasley C, Landau S, Everall IP. Reduced neuronal size and glial cell density in area 9 of the dorsolateral prefrontal cortex in subjects with major depressive disorder. Cereb Cortex. 2002;12:386–94. doi: 10.1093/cercor/12.4.386. [DOI] [PubMed] [Google Scholar]
  • [20].Gittins RA, Harrison PJ. A morphometric study of glia and neurons in the anterior cingulate cortex in mood disorder. J Affect Disord. 2011;133:328–32. doi: 10.1016/j.jad.2011.03.042. [DOI] [PubMed] [Google Scholar]
  • [21].Khundakar A, Morris C, Oakley A, Thomas AJ. A morphometric examination of neuronal and glial cell pathology in the orbitofrontal cortex in late-life depression. Int Psychogeriatr. 2011;23:132–40. doi: 10.1017/S1041610210000700. [DOI] [PubMed] [Google Scholar]
  • [22].Khundakar AA, Morris CM, Oakley AE, Thomas AJ. Cellular pathology within the anterior cingulate cortex of patients with late-life depression: a morphometric study. Psychiatry Res. 2011;194:184–9. doi: 10.1016/j.pscychresns.2011.04.008. [DOI] [PubMed] [Google Scholar]
  • [23].Torres-Platas SG, Hercher C, Davoli MA, et al. Astrocytic hypertrophy in anterior cingulate white matter of depressed suicides. Neuropsychopharmacology. 2011;36:2650–8. doi: 10.1038/npp.2011.154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Maes M, Yirmyia R, Noraberg J, et al. The inflammatory & neurodegenerative (I&ND) hypothesis of depression: leads for future research and new drug developments in depression. Metab Brain Dis. 2009;24:27–53. doi: 10.1007/s11011-008-9118-1. [DOI] [PubMed] [Google Scholar]
  • [25].Jacque CM, Vinner C, Kujas M, Raoul M, Racadot J, Baumann NA. Determination of glial fibrillary acidic protein (GFAP) in human brain tumors. J Neurol Sci. 1978;35:147–55. doi: 10.1016/0022-510x(78)90107-7. [DOI] [PubMed] [Google Scholar]
  • [26].Middeldorp J, Hol EM. GFAP in health and disease. Prog Neurobiol. 2011;93:421–43. doi: 10.1016/j.pneurobio.2011.01.005. [DOI] [PubMed] [Google Scholar]
  • [27].Simard M, Arcuino G, Takano T, Liu QS, Nedergaard M. Signaling at the gliovascular interface. J Neurosci. 2003;23:9254–62. doi: 10.1523/JNEUROSCI.23-27-09254.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Cullen DK, Simon CM, Laplaca MC. Strain rate-dependent induction of reactive astrogliosis and cell death in three-dimensional neuronal-astrocytic co-cultures. Brain Res. 2007;1158:103–15. doi: 10.1016/j.brainres.2007.04.070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Nedergaard M, Ransom B, Goldman SA. New roles for astrocytes: redefining the functional architecture of the brain. Trends Neurosci. 2003;26:523–30. doi: 10.1016/j.tins.2003.08.008. [DOI] [PubMed] [Google Scholar]
  • [30].Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119:7–35. doi: 10.1007/s00401-009-0619-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Miguel-Hidalgo JJ, Baucom C, Dilley G, et al. Glial fibrillary acidic protein immunoreactivity in the prefrontal cortex distinguishes younger from older adults in major depressive disorder. Biol Psychiatry. 2000;48:861–73. doi: 10.1016/s0006-3223(00)00999-9. [DOI] [PubMed] [Google Scholar]
  • [32].Miguel-Hidalgo JJ, Waltzer R, Whittom AA, Rajkowska G, Stockmeier CA. Glial and glutamatergic markers in depression, alcoholism, and their comorbidity. J Affect Disord. 2010;127:230–40. doi: 10.1016/j.jad.2010.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Davis S, Thomas A, Perry R, Oakley A, Kalaria RN, O’Brien JT. Glial fibrillary acidic protein in late life major depressive disorder: an immunocytochemical study. J Neurol Neurosurg Psychiatry. 2002;73:556–60. doi: 10.1136/jnnp.73.5.556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Rajkowska G, Miguel-Hidalgo JJ, Dubey P, Stockmeier CA, Krishnan KR. Prominent reduction in pyramidal neurons density in the orbitofrontal cortex of elderly depressed patients. Biol Psychiatry. 2005;58:297–306. doi: 10.1016/j.biopsych.2005.04.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Khundakar AA, Thomas AJ. Morphometric changes in early- and late-life major depressive disorder: evidence from postmortem studies. Int Psychogeriatr. 2009;21:844–54. doi: 10.1017/S104161020999007X. [DOI] [PubMed] [Google Scholar]
  • [36].Paradise MB, Naismith SL, Norrie LM, Graeber MB, Hickie IB. The role of glia in late-life depression. Int Psychogeriatr. 2012;24:1878–90. doi: 10.1017/S1041610212000828. [DOI] [PubMed] [Google Scholar]
  • [37].Müller MB, Lucassen PJ, Yassouridis A, Hoogendijk WJ, Holsboer F, Swaab DF. Neither major depression nor glucocorticoid treatment affects the cellular integrity of the human hippocampus. Eur J Neurosci. 2001;14:1603–12. doi: 10.1046/j.0953-816x.2001.01784.x. [DOI] [PubMed] [Google Scholar]
  • [38].Wang Q, Van Heerikhuize J, Aronica E, et al. Glucocorticoid receptor protein expression in human hippocampus; stability with age. Neurobiology of Aging. 2013 doi: 10.1016/j.neurobiolaging.2012.11.019. in press. [DOI] [PubMed] [Google Scholar]
  • [39].Gittins RA, Harrison PJ. A morphometric study of glia and neurons in the anterior cingulate cortex in mood disorder. J Affect Disord. 2011;133:328–32. doi: 10.1016/j.jad.2011.03.042. [DOI] [PubMed] [Google Scholar]
  • [40].Si X, Miguel-Hidalgo JJ, O’Dwyer G, Stockmeier CA, Rajkowska G. Age-dependent reductions in the level of glial fibrillary acidic protein in the prefrontal cortex in major depression. Neuropsychopharmacology. 2004;29:2088–96. doi: 10.1038/sj.npp.1300525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Johnston-Wilson NL, Sims CD, Hofmann JP, et al. The Stanley Neuropathology Consortium Disease-specific alterations in frontal cortex brain proteins in schizophrenia, bipolar disorder, and major depressive disorder. Mol Psychiatry. 2000;5:142–9. doi: 10.1038/sj.mp.4000696. [DOI] [PubMed] [Google Scholar]
  • [42].Webster MJ, O’Grady J, Kleinman JE, Weickert CS. Glial fibrillary acidic protein mRNA levels in the cingulate cortex of individuals with depression, bipolar disorder and schizophrenia. Neuroscience. 2005;133:453–61. doi: 10.1016/j.neuroscience.2005.02.037. [DOI] [PubMed] [Google Scholar]
  • [43].Drevets WC. Functional anatomical abnormalities in limbic and prefrontal cortical structures in major depression. Prog Brain Res. 2000;126:413–31. doi: 10.1016/S0079-6123(00)26027-5. [DOI] [PubMed] [Google Scholar]
  • [44].Bremner JD, Vythilingam M, Vermetten E, et al. Reduced volume of orbitofrontal cortex in major depression. Biol Psychiatry. 2002;51:273–9. doi: 10.1016/s0006-3223(01)01336-1. [DOI] [PubMed] [Google Scholar]
  • [45].Malykhin NV, Carter R, Hegadoren KM, Seres P, Coupland NJ. Fronto-limbic volumetric changes in major depressive disorder. J Affect Disord. 2011;36:1104–13. doi: 10.1016/j.jad.2011.10.038. [DOI] [PubMed] [Google Scholar]
  • [46].Salvadore G, Nugent AC, Lemaitre H, et al. Prefrontal cortical abnormalities in currently depressed versus currently remitted patients with major depressive disorder. NeuroImage. 2011;54:2643–51. doi: 10.1016/j.neuroimage.2010.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Lim HK, Jung WS, Ahn KJ, et al. Regional cortical thickness and subcortical volume changes are associated with cognitive impairments in the drug-naïve patients with late-onset depression. Neuropsychopharmacology. 2012;37:838–49. doi: 10.1038/npp.2011.264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Barley K, Dracheva S, Byne W. Subcortical oligodendrocyte- and astrocyte-associated gene expression in subjects with schizophrenia, major depression and bipolar disorder. Schizophr Res. 2009;112:54–64. doi: 10.1016/j.schres.2009.04.019. [DOI] [PubMed] [Google Scholar]
  • [49].Bernard R, Kerman IA, Thompson RC, et al. Altered expression of glutamate signaling, growth factor, and glia genes in the locus coeruleus of patients with major depression. Mol Psychiatry. 2011;16:634–46. doi: 10.1038/mp.2010.44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Chandley M, Szebeni K, Szebeni A, et al. Gene expression deficits in pontine locus coeruleus astrocytes in man with major depressive disorder. Journal of Psychiatry and Neurosc. 2012 doi: 10.1503/jpn.120110. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Ordway GA, Szebeni A, Chandley MJ, et al. Low gene expression of bone morphogenetic protein 7 in brainstem astrocytes in major depression. Int J Neuropsychopharmacol. 2012;15:855–68. doi: 10.1017/S1461145711001350. [DOI] [PubMed] [Google Scholar]
  • [52].Fatemi SH, Laurence JA, Araghi-Niknam M, et al. Glial fibrillary acidic protein is reduced in cerebellum of subjects with major depression, but not schizophrenia. Schizophr Res. 2004;69:317–23. doi: 10.1016/j.schres.2003.08.014. [DOI] [PubMed] [Google Scholar]
  • [53].Altshuler LL, Abulseoud OA, Foland-Ross L, et al. Amygdala astrocyte reduction in subjects with major depressive disorder but not bipolar disorder. Bipolar Disord. 2010;12:541–9. doi: 10.1111/j.1399-5618.2010.00838.x. [DOI] [PubMed] [Google Scholar]
  • [54].Braun K, Antemano R, Helmeke C, Büchner M, Poeggel G. Juvenile separation stress induces rapid region- and layer-specific changes in S100ss- and glial fibrillary acidic protein-immunoreactivity in astrocytes of the rodent medial prefrontal cortex. Neuroscience. 2009;160:629–38. doi: 10.1016/j.neuroscience.2009.02.074. [DOI] [PubMed] [Google Scholar]
  • [55].Czéh B, Simon M, Schmelting B, Hiemke C, Fuchs E. Astroglial plasticity in the hippocampus is affected by chronic psychosocial stress and concomitant fluoxetine treatment. Neuropsychopharmacology. 2006;31:1616–26. doi: 10.1038/sj.npp.1300982. [DOI] [PubMed] [Google Scholar]
  • [56].Araya-Callís C, Hiemke C, Abumaria N, Flugge G. Chronic psychosocial stress and citalopram modulate the expression of the glial proteins GFAP and NDRG2 in the hippocampus. Psychopharmacology (Berl) 2012;224:209–22. doi: 10.1007/s00213-012-2741-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Leventopoulos M, Rüedi-Bettschen D, Knuesel I, Feldon J, Pryce CR, Opacka-Juffry J. Long-term effects of early life deprivation on brain glia in Fischer rats. Brain Res. 2007;1142:119–26. doi: 10.1016/j.brainres.2007.01.039. [DOI] [PubMed] [Google Scholar]
  • [58].Banasr M, Chowdhury GM, Terwilliger R, et al. Glial pathology in an animal model of depression: reversal of stress-induced cellular, metabolic and behavioral deficits by the glutamate-modulating drug riluzole. Mol Psychiatry. 2010;15:501–11. doi: 10.1038/mp.2008.106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Banasr M, Duman RS. Glial loss in the prefrontal cortex is sufficient to induce depressive-like behaviors. Biol Psychiatry. 2008;64:863–70. doi: 10.1016/j.biopsych.2008.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [60].Lee Y, Son H, Kim G, et al. Glutamine deficiency in the prefrontal cortex increases depressive-like behaviours in male mice. J Psychiatry Neurosci. 2012;37:120024. doi: 10.1503/jpn.120024. doi: 10.1503/jpn.120024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [61].Will CC, Aird F, Redei EE. Selectively bred Wistar-Kyoto rats: an animal model of depression and hyper-responsiveness to antidepressants. Mol Psychiatry. 2003;8:925–32. doi: 10.1038/sj.mp.4001345. [DOI] [PubMed] [Google Scholar]
  • [62].Gosselin RD, Gibney S, O’Malley D, Dinan TG, Cryan JF. Region specific decrease in glial fibrillary acidic protein immunoreactivity in the brain of a rat model of depression. Neuroscience. 2009;159:915–25. doi: 10.1016/j.neuroscience.2008.10.018. [DOI] [PubMed] [Google Scholar]
  • [63].Schipke CG, Heuser I, Peters O. Antidepressants act on glial cells: SSRIs and serotonin elicit astrocyte calcium signaling in the mouse prefrontal cortex. J Psychiatr Res. 2011;45:242–8. doi: 10.1016/j.jpsychires.2010.06.005. [DOI] [PubMed] [Google Scholar]
  • [64].Iwata M, Shirayama Y, Ishida H, Hazama GI, Nakagome K. Hippocampal astrocytes are necessary for antidepressant treatment of learned helplessness rats. Hippocampus. 2011;21:877–84. doi: 10.1002/hipo.20803. [DOI] [PubMed] [Google Scholar]
  • [65].Kragh J, Bolwig TG, Woldbye DP, Jørgensen OS. Electroconvulsive shock and lidocaine-induced seizures in the rat activate astrocytes as measured by glial fibrillary acidic protein. Biol Psychiatry. 1993;33:794–800. doi: 10.1016/0006-3223(93)90020-e. [DOI] [PubMed] [Google Scholar]
  • [66].Jansson L, Wennström M, Johanson A, Tingström A. Glial cell activation in response to electroconvulsive seizures. Prog Neuropsychopharmacol Biol Psychiatry. 2009;33:1119–28. doi: 10.1016/j.pnpbp.2009.06.007. [DOI] [PubMed] [Google Scholar]
  • [67].Steward O. Electroconvulsive seizures upregulate astroglial gene expression selectively in the dentate gyrus. Brain Res Mol Brain Res. 1994;25:217–24. doi: 10.1016/0169-328x(94)90156-2. [DOI] [PubMed] [Google Scholar]
  • [68].Fujiki M, Steward O. High frequency transcranial magnetic stimulation mimics the effects of ECS in upregulating astroglial gene expression in the murine CNS. Brain Res Mol Brain Res. 1997;44:301–8. doi: 10.1016/s0169-328x(96)00232-x. [DOI] [PubMed] [Google Scholar]
  • [69].Czéh B, Di Benedetto B. Antidepressants act directly on astrocytes: Evidences and functional consequences. Eur Neuropsychopharmacol. 2013;23(3):171–85. doi: 10.1016/j.euroneuro.2012.04.017. [DOI] [PubMed] [Google Scholar]
  • [70].Giaume C, Theis M. Pharmacological and genetic approaches to study connexin-mediated channels in glial cells of the central nervous system. Brain Res Rev. 2010;63:160–76. doi: 10.1016/j.brainresrev.2009.11.005. [DOI] [PubMed] [Google Scholar]
  • [71].Ernst C, Nagy C, Kim S, et al. Dysfunction of astrocyte connexins 30 and 43 in dorsal lateral prefrontal cortex of suicide completers. Biol Psychiatry. 2011;70:312–9. doi: 10.1016/j.biopsych.2011.03.038. [DOI] [PubMed] [Google Scholar]
  • [72].Sun JD, Liu Y, Yuan YH, Li J, Chen NH. Gap junction dysfunction in the prefrontal cortex induces depressive-like behaviors in rats. Neuropsychopharmacology. 2012;37:1305–20. doi: 10.1038/npp.2011.319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Fatemi SH, Folsom TD, Reutiman TJ, et al. Chronic psychotropic drug treatment causes differential expression of connexin 43 and GFAP in frontal cortex of rats. Schizophr Res. 2008;104:127–34. doi: 10.1016/j.schres.2008.05.016. [DOI] [PubMed] [Google Scholar]
  • [74].Blomstrand F, Aberg ND, Eriksson PS, Hansson E, Rönnbäck L. Extent of intercellular calcium wave propagation is related to gap junction permeability and level of connexin-43 expression in astrocytes in primary cultures from four brain regions. Neuroscience. 1999;92:255–65. doi: 10.1016/s0306-4522(98)00738-6. [DOI] [PubMed] [Google Scholar]
  • [75].Ezan P, André P, Cisternino S, et al. Deletion of astroglial connexins weakens the blood-brain barrier. J Cereb Blood Flow Metab. 2012;32:1457–67. doi: 10.1038/jcbfm.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [76].Rajkowska G, Hughes J, Stockmeier C, Miguel-Hidalgo JJ, Maciag D. Coverage of blood vessels by astrocytic endfeet is reduced in major depressive disorder. Biol Psychiatry. 2012 doi: 10.1016/j.biopsych.2012.09.024. Epub ahead of print. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Nielsen S, Nagelhus EA, Amiry-Moghaddam M, Bourque C, Agre P, Ottersen OP. Specialized membrane domains for water transport in glial cells: high-resolution immunogold cytochemistry of aquaporin-4 in rat brain. J Neurosci. 1997;17:171–80. doi: 10.1523/JNEUROSCI.17-01-00171.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Nagelhus EA, Veruki ML, Torp R, et al. Aquaporin-4 water channel protein in the rat retina and optic nerve: polarized expression in Müller cells and fibrous astrocytes. J Neurosci. 1998;18:2506–19. doi: 10.1523/JNEUROSCI.18-07-02506.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Amiry-Moghaddam M, Ottersen OP. The molecular basis of water transport in the brain. Nat Rev Neurosci. 2003;4:991–1001. doi: 10.1038/nrn1252. [DOI] [PubMed] [Google Scholar]
  • [80].Leybaert L. Neurobarrier coupling in the brain: a partner of neurovascular and neurometabolic coupling? J Cereb Blood Flow Metab. 2005;25:2–16. doi: 10.1038/sj.jcbfm.9600001. [DOI] [PubMed] [Google Scholar]
  • [81].Haydon PG, Carmignoto G. Astrocyte control of synaptic transmission and neurovascular coupling. Physiol Rev. 2006;86:1009–31. doi: 10.1152/physrev.00049.2005. [DOI] [PubMed] [Google Scholar]
  • [82].Paulson OB, Newman EA. Does the release of potassium from astrocyte endfeet regulate cerebral blood flow? Science. 1987;237:896–98. doi: 10.1126/science.3616619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Koehler RC, Roman RJ, Harder DR. Astrocytes and the regulation of cerebral blood flow. Trends Neurosc. 2009;32:160–9. doi: 10.1016/j.tins.2008.11.005. [DOI] [PubMed] [Google Scholar]
  • [84].Kimelberg HK. Water homeostasis in the brain: basic concepts. Neuroscience. 2004;129:851–60. doi: 10.1016/j.neuroscience.2004.07.033. [DOI] [PubMed] [Google Scholar]
  • [85].Nico B, Frigeri A, Nicchia GP, et al. Role of aquaporin-4 water channel in the development and integrity of the blood-brain barrier. J Cell Sci. 2001;114:1297–07. doi: 10.1242/jcs.114.7.1297. [DOI] [PubMed] [Google Scholar]
  • [86].Meshorer E, Biton IE, Ben-Shaul Y, et al. Chronic cholinergic imbalances promote brain diffusion and transport abnormalities. FASEB J. 2005;19:910–22. doi: 10.1096/fj.04-2957com. [DOI] [PubMed] [Google Scholar]
  • [87].Zeng XN, Sun XL, Gao L, Fan Y, Ding JH, Hu G. Aquaporin-4 deficiency down-regulates glutamate uptake and GLT-1 expression in astrocytes. Mol Cell Neurosci. 2007;34:34–9. doi: 10.1016/j.mcn.2006.09.008. [DOI] [PubMed] [Google Scholar]
  • [88].Li YK, Wang F, Wang W, et al. Aquaporin-4 Deficiency Impairs Synaptic Plasticity and Associative Fear Memory in the Lateral Amygdala: Involvement of Downregulation of Glutamate Transporter-1 Expression. Neuropsychopharmacology. 2012;37:1867–78. doi: 10.1038/npp.2012.34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Donato R. Functional roles of S100 proteins, calcium-binding proteins of the EF-hand type. Biochem Biophys Acta. 1999;1450:191–31. doi: 10.1016/s0167-4889(99)00058-0. [DOI] [PubMed] [Google Scholar]
  • [90].Pinto SS, Gottfried C, Mendez A, et al. Immunocontent and secretion of S100B in astrocyte cultures from different brain regions in relation to morphology. FEBS Lett. 2000;486:203–7. doi: 10.1016/s0014-5793(00)02301-2. [DOI] [PubMed] [Google Scholar]
  • [91].Marenholz I, Heizmann CW, Fritz G. S100 proteins in mouse and man: from evolution to function and pathology. Biochem Biophys Res Commun. 2004;322:1111–22. doi: 10.1016/j.bbrc.2004.07.096. [DOI] [PubMed] [Google Scholar]
  • [92].Steiner J, Bernstein HG, Bielau H, et al. Evidence for a wide extra-astrocytic distribution of S100B in human brain. BMC Neurosci. 2007;8:2. doi: 10.1186/1471-2202-8-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Marks A, O’Hanlon D, Lei M, Percy ME, Becker LE. Accumulation of S100 beta mRNA and protein in cerebellum during infancy in Down syndrome and control subjects. Brain Res Mol Brain Res. 1996;36:343–8. doi: 10.1016/0169-328x(95)00293-2. [DOI] [PubMed] [Google Scholar]
  • [94].Rothermundt M, Peters M, Prehn JH, Arolt V. S100B in brain damage and neurodegeneration. Microsc Res Tech. 2003;60:614–32. doi: 10.1002/jemt.10303. [DOI] [PubMed] [Google Scholar]
  • [95].Kleindienst A, Ross Bullock M. A critical analysis of the role of the neurotrophic protein S100B in acute brain injury. J Neurotrauma. 2006;2:1185–2000. doi: 10.1089/neu.2006.23.1185. [DOI] [PubMed] [Google Scholar]
  • [96].Gonçalves CA, Leite MC, Nardin P. Biological and methodological features of the measurement of S100B, a putative marker of brain injury. Clin Biochem. 2008;41:755–63. doi: 10.1016/j.clinbiochem.2008.04.003. [DOI] [PubMed] [Google Scholar]
  • [97].Klempan TA, Sequeira A, Canetti L, et al. Altered expression of genes involved in ATP biosynthesis and GABAergic neurotransmission in the ventral prefrontal cortex of suicides with and without major depression. Mol Psychiatry. 2009;14:175–89. doi: 10.1038/sj.mp.4002110. [DOI] [PubMed] [Google Scholar]
  • [98].Rothermundt M, Arolt V, Wiesmann M, et al. S-100B is increased in melancholic but not in non-melancholic major depression. J Affect Disord. 2001;66:89–93. doi: 10.1016/s0165-0327(00)00321-9. [DOI] [PubMed] [Google Scholar]
  • [99].Grabe HJ, Ahrens N, Rose HJ, Kessler C, Freyberger HJ. Neurotrophic factor S100beta in major depression. Neuropsychobiology. 2001;44:88–90. doi: 10.1159/000054922. [DOI] [PubMed] [Google Scholar]
  • [100].Schroeter ML, Abdul-Khaliq H, Diefenbacher A, Blasig IE. S100B is increased in mood disorders and may be reduced by antidepressive treatment. Neuroreport. 2002;13:1675–78. doi: 10.1097/00001756-200209160-00021. [DOI] [PubMed] [Google Scholar]
  • [101].Arolt V, Peters M, Erfurth A, et al. S100B and response to treatment in major depression: a pilot study. Eur Neuropsychopharmacol. 2003;13:235–39. doi: 10.1016/s0924-977x(03)00016-6. [DOI] [PubMed] [Google Scholar]
  • [102].Hetzel G, Moeller O, Evers S, et al. The astroglial protein S100B and visually evoked event-related potentials before and after antidepressant treatment. Psychopharmacology. 2005;178:161–6. doi: 10.1007/s00213-004-1999-z. [DOI] [PubMed] [Google Scholar]
  • [103].Schroeter ML, Abdul-Khaliq H, Krebs M, Diefenbacher A, Blasig IE. Serum markers support disease-specific glial pathology in major depression. J Affect Disord. 2008;111:271–80. doi: 10.1016/j.jad.2008.03.005. [DOI] [PubMed] [Google Scholar]
  • [104].Anderson CM, Swanson RA. Astrocyte glutamate transport: review of properties, regulation, and physiological functions. Glia. 2000;32:1–14. [PubMed] [Google Scholar]
  • [105].Bezzi P, Gundersen V, Galbete JL, et al. Astrocytes contain a vesicular compartment that is competent for regulated exocytosis of glutamate. Nat Neurosci. 2004;7:613–20. doi: 10.1038/nn1246. [DOI] [PubMed] [Google Scholar]
  • [106].Furuta A, Takashima S, Yokoo H, Rothstein JD, Wada K, Iwaki T. Expression of glutamate transporter subtypes during normal human corticogenesis and type II lissencephaly. Brain Res Dev Brain Res. 2005;155:155–64. doi: 10.1016/j.devbrainres.2005.01.005. [DOI] [PubMed] [Google Scholar]
  • [107].Toro CT, Hallak JE, Dunham JS, Deakin JF. Glial fibrillary acidic protein and glutamine synthetase in subregions of prefrontal cortex in schizophrenia and mood disorder. Neurosci Lett. 2006;404(3):276–81. doi: 10.1016/j.neulet.2006.05.067. [DOI] [PubMed] [Google Scholar]
  • [108].Choudary PV, Molnar M, Evans SJ, et al. Altered cortical glutamatergic and GABAergic signal transmission with glial involvement in depression. Proc Natl Acad Sci USA. 2005;102:15653–58. doi: 10.1073/pnas.0507901102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Sequeira A, Mamdani F, Ernst C, et al. Global brain gene expression analysis links glutamatergic and GABAergic alterations to suicide and major depression. PLoS One. 2009;4:e6585. doi: 10.1371/journal.pone.0006585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Lee Y, Gaskins D, Anand A, Shekhar A. Glia mechanisms in mood regulation: a novel model of mood disorders. Psychopharmacology. 2007;191:55–65. doi: 10.1007/s00213-006-0652-4. [DOI] [PubMed] [Google Scholar]
  • [111].Herberg LJ, Rose IC. Excitatory amino acid pathways in brain-stimulation reward. Behav Brain Res. 1990;39:230–39. doi: 10.1016/0166-4328(90)90029-e. [DOI] [PubMed] [Google Scholar]
  • [112].John CS, Smith KL, Van’t Veer A, et al. Blockade of astrocytic glutamate uptake in the prefrontal cortex induces anhedonia. Neuropsychopharmacol. 2012;37:2467–75. doi: 10.1038/npp.2012.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Hughes EG, Maguire JL, McMinn MT, Scholz RE, Sutherland ML. Loss of glial fibrillary acidic protein results in decreased glutamate transport and inhibition of PKA-induced EAAT2 cell surface trafficking. Brain Res Mol Brain Res. 2004;124:114–23. doi: 10.1016/j.molbrainres.2004.02.021. [DOI] [PubMed] [Google Scholar]
  • [114].Sullivan SM, Lee A, Björkman ST, et al. Cytoskeletal anchoring of GLAST determines susceptibility to brain damage: an identified role for GFAP. J Biol Chem. 2007;282:29414–23. doi: 10.1074/jbc.M704152200. [DOI] [PubMed] [Google Scholar]
  • [115].Altamura C, Maes M, Dai J, Meltzer HY. Plasma concentrations of excitatory amino acids, serine, glycine, taurine and histidine in major depression. Eur Neuropsychopharmacol. 1995;5:71–5. doi: 10.1016/0924-977x(95)00033-l. [DOI] [PubMed] [Google Scholar]
  • [116].Yildiz-Yesiloglu A, Ankerst DP. Review of 1H magnetic resonance spectroscopy findings in major depressive disorder: a meta-analysis. Psychiatry Res. 2006;147:1–25. doi: 10.1016/j.pscychresns.2005.12.004. [DOI] [PubMed] [Google Scholar]
  • [117].Hasler G, van der Veen JW, Tumonis T, Meyers N, Shen J, Drevets WC. Reduced prefrontal glutamate/glutamine and gamma-aminobutyric acid levels in major depression determined using proton magnetic resonance spectroscopy. Arch Gen Psychiatry. 2007;64:193–200. doi: 10.1001/archpsyc.64.2.193. [DOI] [PubMed] [Google Scholar]
  • [118].Auer DP, Putz B, Kraft E, Lipinski B, Schill J, Holsboer F. Reduced glutamate in the anterior cingulate cortex in depression: an in vivo proton magnetic resonance spectroscopy study. Biol Psychiatry. 2000;47:305–13. doi: 10.1016/s0006-3223(99)00159-6. [DOI] [PubMed] [Google Scholar]
  • [119].Mirza Y, Tang J, Russell A, Banerjee SP, Bhandari R, Ivey J, Rose M, Moore GJ, Rosenberg DR. Reduced anterior cingulate cortex glutamatergic concentrations in childhood major depression. J Am Acad Child Adolesc Psychiatry. 2004;43:341–48. doi: 10.1097/00004583-200403000-00017. [DOI] [PubMed] [Google Scholar]
  • [120].Rosenberg DR, Macmaster FP, Mirza Y, et al. Reduced anterior cingulate glutamate in pediatric major depression: a magnetic resonance spectroscopy study. Biol Psychiatry. 2005;58:700–4. doi: 10.1016/j.biopsych.2005.05.007. [DOI] [PubMed] [Google Scholar]
  • [121].Pfleiderer B, Michael N, Erfurth A, et al. Effective electroconvulsive therapy reverses glutamate/glutamine deficit in the left anterior cingulum of unipolar depressed patients. Psychiatry Res. 2003;122:185–92. doi: 10.1016/s0925-4927(03)00003-9. [DOI] [PubMed] [Google Scholar]
  • [122].Michael N, Erfurth A, Ohrmann P, Arolt V, Heindel W, Pfleiderer B. Neurotrophic effects of electroconvulsive therapy: a proton magnetic resonance study of the left amygdalar region in patients with treatment-resistant depression. Neuropsychopharmacology. 2003;28:720–25. doi: 10.1038/sj.npp.1300085. [DOI] [PubMed] [Google Scholar]
  • [123].Sanacora G, Gueorguieva R, Epperson CN, et al. Subtype-specific alterations of gamma-aminobutyric acid and glutamate in patients with major depression. Arch Gen Psychiatry. 2004;61:705–13. doi: 10.1001/archpsyc.61.7.705. [DOI] [PubMed] [Google Scholar]
  • [124].Hashimoto K, Sawa A, Iyo M. Increased levels of glutamate in brains from patients with mood disorders. Biol Psychiatry. 2007;62:1310–6. doi: 10.1016/j.biopsych.2007.03.017. [DOI] [PubMed] [Google Scholar]
  • [125].Nichols NR, Osterburg HH, Masters JN, Millar SL, Finch CE. Messenger RNA for glial fibrillary acidic protein is decreased in rat brain following acute and chronic corticosterone treatment. Brain Res Mol Brain Res. 1990;7:1–7. doi: 10.1016/0169-328x(90)90066-m. [DOI] [PubMed] [Google Scholar]
  • [126].Rajkowska G, O’Dwyer G, Teleki Z, Stockmeier CA, Miguel-Hidalgo JJ. GABAergic neurons immunoreactive for calcium binding proteins are reduced in the prefrontal cortex in major depression. Neuropsychopharmacology. 2007;32:471–82. doi: 10.1038/sj.npp.1301234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Oh DH, Son H, Hwang S, Kim SH. Neuropathological abnormalities of astrocytes, GABAergic neurons, and pyramidal neurons in the dorsolateral prefrontal cortices of patients with major depressive disorder. Eur Neuropsychopharmacol. 2012;22:330–8. doi: 10.1016/j.euroneuro.2011.09.001. [DOI] [PubMed] [Google Scholar]
  • [128].Araque A, Parpura V, Sanzgiri RP, Haydon PG. Tripartite synapses: glia, the unacknowledged partner. Trends Neurosci. 1999;22:208–15. doi: 10.1016/s0166-2236(98)01349-6. [DOI] [PubMed] [Google Scholar]
  • [129].Nedergaard M, Verkhratsky A. Artifact versus reality--how astrocytes contribute to synaptic events. Glia. 2012;60:1013–23. doi: 10.1002/glia.22288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Pfrieger FW, Barres BA. Synaptic efficacy enhanced by glial cells in vitro. Science. 1997;277:1684–7. doi: 10.1126/science.277.5332.1684. [DOI] [PubMed] [Google Scholar]
  • [131].Nägler K, Mauch DH, Pfrieger FW. Glia-derived signals induce synapse formation in neurones of the rat central nervous system. J Physiol. 2001;533:665–79. doi: 10.1111/j.1469-7793.2001.00665.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [132].Ullian EM, Sapperstein SK, Christopherson KS, Barres BA. Control of synapse number by glia. Science. 2001;291:657–61. doi: 10.1126/science.291.5504.657. [DOI] [PubMed] [Google Scholar]
  • [133].Pfrieger FW. In: The tripartite synapse. Glia in synaptic transmission. Volterra A, Magistretti PJ, Haydon PG, editors. Oxford University Press; New York: 2002. pp. 24–34. [Google Scholar]
  • [134].Kang HJ, Voleti B, Hajszan T, et al. Decreased expression of synapse-related genes and loss of synapses in major depressive disorder. Nat Med. 2012;18:1413–7. doi: 10.1038/nm.2886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [135].Raichle ME. Cognitive neuroscience. Bold insights. Nature. 2001;412:128–30. doi: 10.1038/35084300. [DOI] [PubMed] [Google Scholar]
  • [136].Rossi DJ. Another BOLD role for astrocytes: coupling blood flow to neural activity. Nat Neurosci. 2006;9:159–61. doi: 10.1038/nn0206-159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [137].Magistretti PJ, Pellerin L. Cellular mechanisms of brain energy metabolism and their relevance to functional brain imaging. Philos Trans R Soc Lond B Biol Sci. 1999;354:1155–63. doi: 10.1098/rstb.1999.0471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [138].Hundal Ø . Major depressive disorder viewed as a dysfunction in astroglial bioenergetics. Med Hypotheses. 2007;68:370–7. doi: 10.1016/j.mehy.2006.06.050. [DOI] [PubMed] [Google Scholar]
  • [139].Zonta M, Angulo MC, Gobbo S, et al. Neuron-to-astrocyte signaling is central to the dynamic control of brain microcirculation. Nat Neurosci. 2003;6:43–50. doi: 10.1038/nn980. [DOI] [PubMed] [Google Scholar]
  • [140].Parri R, Crunelli V. An astrocyte bridge from synapse to blood flow. Nat Neurosci. 2003;6:5–6. doi: 10.1038/nn0103-5. [DOI] [PubMed] [Google Scholar]
  • [141].Takano T, Tian GF, Peng W, et al. Astrocyte-mediated control of cerebral blood flow. Nat Neurosci. 2006;9:260–7. doi: 10.1038/nn1623. [DOI] [PubMed] [Google Scholar]
  • [142].Hertz L, Wu PH, Schousboe A. Evidence for net uptake of GABA into mouse astrocytes in primary cultures--its sodium dependence and potassium independence. Neurochem Res. 1978;3:313–23. doi: 10.1007/BF00965577. [DOI] [PubMed] [Google Scholar]
  • [143].Minelli A, DeBiasi S, Brecha NC, Zuccarello LV, Conti F. GAT-3, a high-affinity GABA plasma membrane transporter, is localized to astrocytic processes, and it is not confined to the vicinity of GABAergic synapses in the cerebral cortex. J Neurosci. 1996;16:6255–64. doi: 10.1523/JNEUROSCI.16-19-06255.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Wu Q, Wada M, Shimada A, Yamamoto A, Fujita T. Functional characterization of Zn2(+)-sensitive GABA transporter expressed in primary cultures of astrocytes from rat cerebral cortex. Brain Res. 2006;1075:100–9. doi: 10.1016/j.brainres.2005.12.109. [DOI] [PubMed] [Google Scholar]
  • [145].Bal N, Figueras G, Vilaró MT, Suñol C, Artigas F. Antidepressant drugs inhibit a glial 5-hydroxytryptamine transporter in rat brain. Eur J Neurosci. 1997;9:1728–38. doi: 10.1111/j.1460-9568.1997.tb01530.x. [DOI] [PubMed] [Google Scholar]
  • [146].Hirst WD, Price GW, Rattray M, Wilkin GP. Serotonin transporters in adult rat brain astrocytes revealed by [3H]5-HT uptake into glial plasmalemmal vesicles. Neurochem Int. 1998;33:11–22. doi: 10.1016/s0197-0186(05)80003-8. [DOI] [PubMed] [Google Scholar]
  • [147].Pelton EW, 2nd, Kimelberg HK, Shipherd SV, Bourke RS. Dopamine and norepinephrine uptake and metabolism by astroglial cells in culture. Life Sci. 1981;28:1655–63. doi: 10.1016/0024-3205(81)90322-2. [DOI] [PubMed] [Google Scholar]
  • [148].Takeda H, Inazu M, Matsumiya T. Astroglial dopamine transport is mediated by norepinephrine transporter. Naunyn Schmiedebergs Arch Pharmacol. 2002;366:620–3. doi: 10.1007/s00210-002-0640-0. [DOI] [PubMed] [Google Scholar]
  • [149].Inazu M, Takeda H, Matsumiya T. Functional expression of the norepinephrine transporter in cultured rat astrocytes. J Neurochem. 2003;84:136–44. doi: 10.1046/j.1471-4159.2003.01514.x. [DOI] [PubMed] [Google Scholar]
  • [150].Inazu M, Takeda H, Matsumiya T. Expression and functional characterization of the extraneuronal monoamine transporter in normal human astrocytes. J Neurochem. 2003;84:43–52. doi: 10.1046/j.1471-4159.2003.01566.x. [DOI] [PubMed] [Google Scholar]
  • [151].Fitzgerald LW, Kaplinsky L, Kimelberg HK. Serotonin metabolism by monoamine oxidase in rat primary astrocyte cultures. J Neurochem. 1990;55:2008–14. doi: 10.1111/j.1471-4159.1990.tb05789.x. [DOI] [PubMed] [Google Scholar]
  • [152].Meyer JH, Ginovart N, Boovariwala A, et al. Elevated monoamine oxidase a levels in the brain: an explanation for the monoamine imbalance of major depression. Arch Gen Psychiatry. 2006;63:1209–16. doi: 10.1001/archpsyc.63.11.1209. [DOI] [PubMed] [Google Scholar]
  • [153].Meyer JH, Wilson AA, Sagrati S, et al. Brain monoamine oxidase A binding in major depressive disorder: relationship to selective serotonin reuptake inhibitor treatment, recovery, and recurrence. Arch Gen Psychiatry. 2009;66:1304–12. doi: 10.1001/archgenpsychiatry.2009.156. [DOI] [PubMed] [Google Scholar]
  • [154].Sacher J, Wilson AA, Houle S, et al. Elevated brain monoamine oxidase A binding in the early postpartum period. Arch Gen Psychiatry. 2010;67:468–74. doi: 10.1001/archgenpsychiatry.2010.32. [DOI] [PubMed] [Google Scholar]
  • [155].Johnson S, Stockmeier CA, Meyer JH, et al. The reduction of R1, a novel repressor protein for monoamine oxidase A, in major depressive disorder. Neuropsychopharmacology. 2011;36:2139–48. doi: 10.1038/npp.2011.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Steinhäuser C, Jabs R, Kettenmann H. Properties of GABA and glutamate responses in identified glial cells of the mouse hippocampal slice. Hippocampus. 1994;4:19–35. doi: 10.1002/hipo.450040105. [DOI] [PubMed] [Google Scholar]
  • [157].Azmitia EC, Gannon PJ, Kheck NM, Whitaker-Azmitia PM. Cellular localization of the 5-HT1A receptor in primate brain neurons and glial cells. Neuropsychopharmacology. 1996;14:35–46. doi: 10.1016/S0893-133X(96)80057-1. [DOI] [PubMed] [Google Scholar]
  • [158].Carson MJ, Thomas EA, Danielson PE, Sutcliffe JG. The 5HT5A serotonin receptor is expressed predominantly by astrocytes in which it inhibits cAMP accumulation: a mechanism for neuronal suppression of reactive astrocytes. Glia. 1996;17:317–26. doi: 10.1002/(SICI)1098-1136(199608)17:4<317::AID-GLIA6>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
  • [159].Steinhäuser C, Gallo V. News on glutamate receptors in glial cells. Trends Neurosci. 1996;19:339–45. doi: 10.1016/0166-2236(96)10043-6. [DOI] [PubMed] [Google Scholar]
  • [160].Hirst WD, Cheung NY, Rattray M, Price GW, Wilkin GP. Cultured astrocytes express messenger RNA for multiple serotonin receptor subtypes, without functional coupling of 5-HT1 receptor subtypes to adenylyl cyclase. Brain Res Mol Brain Res. 1998;61:90–9. doi: 10.1016/s0169-328x(98)00206-x. [DOI] [PubMed] [Google Scholar]
  • [161].Wu C, Singh SK, Dias P, Kumar S, Mann DM. Activated astrocytes display increased 5-HT2a receptor expression in pathological states. Exp Neurol. 1999;158:529–33. doi: 10.1006/exnr.1999.7105. [DOI] [PubMed] [Google Scholar]
  • [162].Verkhratsky A, Steinhäuser C. Ion channels in glial cells. Brain Res Brain Res Rev. 2000;32:380–412. doi: 10.1016/s0165-0173(99)00093-4. [DOI] [PubMed] [Google Scholar]
  • [163].Azmitia EC. Modern views on an ancient chemical: serotonin effects on cell proliferation, maturation, and apoptosis. Brain Res Bull. 2001;56:413–24. doi: 10.1016/s0361-9230(01)00614-1. [DOI] [PubMed] [Google Scholar]
  • [164].Khan ZU, Koulen P, Rubinstein M, Grandy DK, Goldman-Rakic PS. An astroglia-linked dopamine D2-receptor action in prefrontal cortex. Proc Natl Acad Sci USA. 2001;98:1964–9. doi: 10.1073/pnas.98.4.1964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [165].Kettenmann H, Steinhauser C. In: Neuroglia. 2nd ed Kettenmann H, Ransom BR, editors. Oxford University Press; New York: 2005. pp. 131–45. [Google Scholar]
  • [166].Aloisi F. In: Neuroglia. 2nd ed Kettenmann H, Ransom BR, editors. Oxford University Press; New York: 2005. pp. 285–301. [Google Scholar]
  • [167].Schmidt HD, Shelton RC, Duman RS. Functional biomarkers of depression: diagnosis, treatment, and pathophysiology. Neuropsychopharmacology. 2011;36:2375–94. doi: 10.1038/npp.2011.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [168].Crnković D, Buljan D, Karlović D, Krmek M. Connection between inflammatory markers, antidepressants and depression. Acta Clin Croat. 2012;51:25–33. [PubMed] [Google Scholar]
  • [169].Karlović D, Serretti A, Vrkić N, Martinac M, Marčinko D. Serum concentrations of CRP, IL-6, TNF-α and cortisol in major depressive disorder with melancholic or atypical features. Psychiatry Res. 2012;198:74–80. doi: 10.1016/j.psychres.2011.12.007. [DOI] [PubMed] [Google Scholar]
  • [170].Maes M, Song C, Yirmiya R. Targeting IL-1 in depression. Expert Opin Ther Targets. 2012;16:1097–112. doi: 10.1517/14728222.2012.718331. [DOI] [PubMed] [Google Scholar]
  • [171].Vogelzangs N, Duivis HE, Beekman AT, et al. Association of depressive disorders, depression characteristics and antidepressant medication with inflammation. Transl Psychiatry. 2012;2:e79. doi: 10.1038/tp.2012.8. doi: 10.1038/tp.2012.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [172].Pandey GN, Rizavi HS, Ren X, et al. Proinflammatory cytokines in the prefrontal cortex of teenage suicide victims. J Psychiatr Res. 2012;46:57–63. doi: 10.1016/j.jpsychires.2011.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [173].Kagaya A, Kugaya A, Takebayashi M, et al. Plasma concentrations of interleukin-1beta, interleukin-6, soluble interleukin-2 receptor and tumor necrosis factor alpha of depressed patients in Japan. Neuropsychobiology. 2001;43:59–62. doi: 10.1159/000054867. [DOI] [PubMed] [Google Scholar]
  • [174].Tuglu C, Kara SH, Caliyurt O, Vardar E, Abay E. Increased serum tumor necrosis factor-alpha levels and treatment response in major depressive disorder. Psychopharmacology (Berl) 2003;170:429–33. doi: 10.1007/s00213-003-1566-z. [DOI] [PubMed] [Google Scholar]
  • [175].Basterzi AD, Aydemir C, Kisa C, et al. IL-6 levels decrease with SSRI treatment in patients with major depression. Hum Psychopharmacol. 2005;20:473–76. doi: 10.1002/hup.717. [DOI] [PubMed] [Google Scholar]
  • [176].Leo R, Di Lorenzo G, Tesauro M, et al. Association between enhanced soluble CD40 ligand and proinflammatory and prothrombotic states in major depressive disorder: pilot observations on the effects of selective serotonin reuptake inhibitor therapy. J Clin Psychiatry. 2006;67:1760–6. doi: 10.4088/jcp.v67n1114. [DOI] [PubMed] [Google Scholar]
  • [177].Yoshimura R, Hori H, Ikenouchi-Sugita A, Umene-Nakano W, Ueda N, Nakamura J. Higher plasma interleukin-6 (IL-6) level is associated with SSRI- or SNRI-refractory depression. Prog Neuropsychopharmacol Biol Psychiatry. 2009;33:722–26. doi: 10.1016/j.pnpbp.2009.03.020. [DOI] [PubMed] [Google Scholar]
  • [178].Laping NJ, Teter B, Nichols NR, Rozovsky I, Finch CE. Glial fibrillary acidic protein: regulation by hormones, cytokines, and growth factors. Brain Pathol. 1994;4:259–75. doi: 10.1111/j.1750-3639.1994.tb00841.x. [DOI] [PubMed] [Google Scholar]
  • [179].Mongin AA, Kimelberg HK. In: Neuroglia. 2nd ed Kettenmann H, Ransom BR, editors. Oxford University Press; New York: 2005. pp. 550–62. [Google Scholar]
  • [180].Manji HK, Moore GJ, Rajkowska G, Chen G. Neuroplasticity and cellular resilience in mood disorders. Mol Psychiatry. 2000;5:578–93. doi: 10.1038/sj.mp.4000811. [DOI] [PubMed] [Google Scholar]
  • [181].Pittenger C, Duman RS. Stress, depression, and neuroplasticity: a convergence of mechanisms. Neuropsychopharmacology. 2008;33:88–109. doi: 10.1038/sj.npp.1301574. [DOI] [PubMed] [Google Scholar]
  • [182].Dwivedi Y, Rizavi HS, Conley RR, Roberts RC, Tamminga CA, Pandey GN. Altered gene expression of brain-derived neurotrophic factor and receptor tyrosine kinase B in postmortem brain of suicide subjects. Arch Gen Psychiatry. 2003;60:804–15. doi: 10.1001/archpsyc.60.8.804. [DOI] [PubMed] [Google Scholar]
  • [183].Evans SJ, Choudary PV, Neal CR, et al. Dysregulation of the fibroblast growth factor system in major depression. Proc Natl Acad Sci USA. 2004;101:15506–11. doi: 10.1073/pnas.0406788101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [184].Duman RS, Monteggia LM. A neurotrophic model for stress-related mood disorders. Biol Psychiatry. 2006;59:1116–27. doi: 10.1016/j.biopsych.2006.02.013. [DOI] [PubMed] [Google Scholar]
  • [185].Gaughran F, Payne J, Sedgwick PM, Cotter D, Berry M. Hippocampal FGF-2 and FGFR1 mRNA expression in major depression, schizophrenia and bipolar disorder. Brain Res Bull. 2006;70:221–7. doi: 10.1016/j.brainresbull.2006.04.008. [DOI] [PubMed] [Google Scholar]
  • [186].Kang HJ, Adams DH, Simen A, et al. Gene expression profiling in postmortem prefrontal cortex of major depressive disorder. J Neurosci. 2007;27:13329–40. doi: 10.1523/JNEUROSCI.4083-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [187].Schmidt HD, Duman RS. The role of neurotrophic factors in adult hippocampal neurogenesis, antidepressant treatments and animal models of depressive-like behavior. Behav Pharmacol. 2007;18:391–418. doi: 10.1097/FBP.0b013e3282ee2aa8. [DOI] [PubMed] [Google Scholar]
  • [188].Ernst C, Deleva V, Deng X, et al. Alternative splicing, methylation state, and expression profile of tropomyosin-related kinase B in the frontal cortex of suicide completers. Arch Gen Psychiatry. 2009;66:22–32. doi: 10.1001/archpsyc.66.1.22. [DOI] [PubMed] [Google Scholar]
  • [189].Kerman IA. New insights into BDNF signaling: relevance to major depression and antidepressant action. Am J Psychiatry. 2012;169:1137–40. doi: 10.1176/appi.ajp.2012.12081053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [190].Tripp A, Oh H, Guilloux JP, Martinowich K, Lewis DA, Sibille E. Brain-derived neurotrophic factor signaling and subgenual anterior cingulate cortex dysfunction in major depressive disorder. Am J Psychiatry. 2012;169:1194–202. doi: 10.1176/appi.ajp.2012.12020248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [191].Mallei A, Shi B, Mocchetti I. Antidepressant treatments induce the expression of basic fibroblast growth factor in cortical and hippocampal neurons. Mol Pharmacol. 2002;61:1017–24. doi: 10.1124/mol.61.5.1017. [DOI] [PubMed] [Google Scholar]
  • [192].Bachis A, Mallei A, Cruz MI, Wellstein A, Mocchetti I. Chronic antidepressant treatments increase basic fibroblast growth factor and fibroblast growth factor-binding protein in neurons. Neuropharmacology. 2008;55:1114–20. doi: 10.1016/j.neuropharm.2008.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [193].Elsayed M, Banasr M, Duric V, Fournier NM, Licznerski P, Duman RS. Antidepressant effects of fibroblast growth factor-2 in behavioral and cellular models of depression. Biol Psychiatry. 2012;72:258–65. doi: 10.1016/j.biopsych.2012.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [194].Reuss B, Unsicker K. In: Neuroglia. 2nd ed Kettenmann H, Ransom BR, editors. Oxford University Press; New York: 2005. pp. 377–88. [Google Scholar]
  • [195].Allaman I, Fiumelli H, Magistretti PJ, Martin JL. Fluoxetine regulates the expression of neurotrophic/growth factors and glucose metabolism in astrocytes. Psychopharmacology (Berl) 2011;216:75–84. doi: 10.1007/s00213-011-2190-y. [DOI] [PubMed] [Google Scholar]
  • [196].Zhang X, Zhang Z, Xie C, et al. Effect of treatment on serum glial cell line-derived neurotrophic factor in depressed patients. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32:886–90. doi: 10.1016/j.pnpbp.2008.01.004. [DOI] [PubMed] [Google Scholar]
  • [197].Otsuki K, Uchida S, Watanuki T, et al. Altered expression of neurotrophic factors in patients with major depression. J Psychiatr Res. 2008;42:1145–53. doi: 10.1016/j.jpsychires.2008.01.010. [DOI] [PubMed] [Google Scholar]
  • [198].Takebayashi M, Hisaoka K, Nishida A, et al. Decreased levels of whole blood glial cell line-derived neurotrophic factor (GDNF) in remitted patients with mood disorders. Int J Neuropsychopharmacol. 2006;9:607–12. doi: 10.1017/S1461145705006085. [DOI] [PubMed] [Google Scholar]
  • [199].Liu Q, Zhu HY, Li B, Wang YQ, Yu J, Wu GC. Chronic clomipramine treatment restores hippocampal expression of glial cell line-derived neurotrophic factor in a rat model of depression. J Affect Disord. 2012;141:367–72. doi: 10.1016/j.jad.2012.03.018. [DOI] [PubMed] [Google Scholar]
  • [200].Chen AC, Eisch AJ, Sakai N, Takahashi M, Nestler EJ, Duman RS. Regulation of GFRalpha-1 and GFRalpha-2 mRNAs in rat brain by electroconvulsive seizure. Synapse. 2001;39:42–50. doi: 10.1002/1098-2396(20010101)39:1<42::AID-SYN6>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  • [201].Zhang X, Zhang Z, Sha W, et al. Electroconvulsive therapy increases glial cell-line derived neurotrophic factor (GDNF) serum levels in patients with drug-resistant depression. Psychiatry Res. 2009;170:273–5. doi: 10.1016/j.psychres.2009.01.011. [DOI] [PubMed] [Google Scholar]
  • [202].Michel TM, Frangou S, Camara S, et al. Altered glial cell line-derived neurotrophic factor (GDNF) concentrations in the brain of patients with depressive disorder: a comparative post-mortem study. Eur Psychiatry. 2008;23:413–20. doi: 10.1016/j.eurpsy.2008.06.001. [DOI] [PubMed] [Google Scholar]
  • [203].Chana G, Landau S, Beasley C, Everall IP, Cotter D. Two-dimensional assessment of cytoarchitecture in the anterior cingulate cortex in major depressive disorder, bipolar disorder, and schizophrenia: evidence for decreased neuronal somal size and increased neuronal density. Biol Psychiatry. 2003;53:1086–98. doi: 10.1016/s0006-3223(03)00114-8. [DOI] [PubMed] [Google Scholar]
  • [204].Van Otterloo E, O’Dwyer G, Stockmeier CA, Steffens DC, Krishnan RR, Rajkowska G. Reductions in neuronal density in elderly depressed are region specific. Int J Geriatr Psychiatry. 2009;24:856–64. doi: 10.1002/gps.2281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [205].Cobb JA, Simpson J, Mahajan GJ, et al. Hippocampal volume and total cell numbers in major depressive disorder. J Psychiatr Res. 2013;47(3):299–306. doi: 10.1016/j.jpsychires.2012.10.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [206].Stockmeier CA, Mahajan GJ, Konick LC, et al. Cellular changes in the postmortem hippocampus in major depression. Biol Psychiatry. 2004;56:640–50. doi: 10.1016/j.biopsych.2004.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [207].Hercher C, Canetti L, Turecki G, Mechawar N. Anterior cingulate pyramidal neurons display altered dendritic branching in depressed suicides. J Psychiatr Res. 2010;44:286–93. doi: 10.1016/j.jpsychires.2009.08.011. [DOI] [PubMed] [Google Scholar]
  • [208].Yildiz-Yesiloglu A, Ankerst DP. Review of 1H magnetic resonance spectroscopy findings in major depressive disorder: a meta-analysis. Psychiatry Res. 2006;147:1–25. doi: 10.1016/j.pscychresns.2005.12.004. [DOI] [PubMed] [Google Scholar]
  • [209].Schroeter ML, Abdul-Khaliq H, Sacher J, Steiner J, Blasig IE, Mueller K. Mood disorders are glial disorders: evidence from in vivo studies. Cardiovasc Psychiatry Neurol. 2010;2010:780645. doi: 10.1155/2010/780645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [210].Toro CT, Hallak JE, Dunham JS, Deakin JF. Glial fibrillary acidic protein and glutamine synthetase in subregions of prefrontal cortex in schizophrenia and mood disorder. Neurosci Lett. 2006;404:276–81. doi: 10.1016/j.neulet.2006.05.067. [DOI] [PubMed] [Google Scholar]
  • [211].Williams MR, Hampton T, Pearce RK, et al. Astrocyte decrease in the subgenual cingulate and callosal genu in schizophrenia. Eur Arch Psychiatry Clin Neurosci. 2013;263(1):41–52. doi: 10.1007/s00406-012-0328-5. [DOI] [PubMed] [Google Scholar]
  • [212].Roberts GW, Colter N, Lofthouse R, Bogerts B, Zech M, Crow TJ. Gliosis in schizophrenia: a survey. Biol Psychiatry. 1986;21:1043–50. doi: 10.1016/0006-3223(86)90285-4. [DOI] [PubMed] [Google Scholar]
  • [213].Roberts GW, Colter N, Lofthouse R, Johnstone EC, Crow TJ. Is there gliosis in schizophrenia? Investigation of the temporal lobe. Biol Psychiatry. 1987;22:1459–68. doi: 10.1016/0006-3223(87)90104-1. [DOI] [PubMed] [Google Scholar]
  • [214].Falkai P, Honer WG, David S, Bogerts B, Majtenyi C, Bayer TA. No evidence for astrogliosis in brains of schizophrenic patients. A post-mortem study. Neuropathol Appl Neurobiol. 1999;25:48–53. doi: 10.1046/j.1365-2990.1999.00162.x. [DOI] [PubMed] [Google Scholar]
  • [215].Rajkowska G, Miguel-Hidalgo JJ, Makkos Z, Meltzer H, Overholser J, Stockmeier C. Layer-specific reductions in GFAP-reactive astroglia in the dorsolateral prefrontal cortex in schizophrenia. Schizophr Res. 2002;57:127–38. doi: 10.1016/s0920-9964(02)00339-0. [DOI] [PubMed] [Google Scholar]
  • [216].Miguel-Hidalgo JJ, Wei J, Andrew M, Overholser JC, Jurjus G, Stockmeier CA, Rajkowska G. Glia pathology in the prefrontal cortex in alcohol dependence with and without depressive symptoms. Biol Psychiatry. 2002;52:1121–33. doi: 10.1016/s0006-3223(02)01439-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [217].Miguel-Hidalgo JJ, Rajkowska G. Comparison of prefrontal cell pathology between depression and alcohol dependence. J Psychiatr Res. 2003;37:411–20. doi: 10.1016/s0022-3956(03)00049-9. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES