Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jan 1.
Published in final edited form as: Neurobiol Dis. 2011 Feb 17;45(1):37–47. doi: 10.1016/j.nbd.2011.01.025

Mitochondria, oligodendrocytes and inflammation in bipolar disorder: evidence from transcriptome studies points to intriguing parallels with multiple sclerosis

Christine Konradi 1,2,3,*, Stephanie E Daws 4, Hayley B Clay 4
PMCID: PMC3117935  NIHMSID: NIHMS274867  PMID: 21310238

Abstract

Gene expression studies of bipolar disorder (BPD) have shown changes in transcriptome profiles in multiple brain regions. Here we summarize the most consistent findings in the scientific literature, and compare them to data from schizophrenia (SZ) and major depressive disorder (MDD). The transcriptome profiles of all three disorders overlap, making the existence of a BPD-specific profile unlikely.

Three groups of functionally related genes are consistently expressed at altered levels in BPD, SZ and MDD. Genes involved in energy metabolism and mitochondrial function are downregulated, genes involved in immune response and inflammation are upregulated, and genes expressed in oligodendrocytes are downregulated. Experimental paradigms for multiple sclerosis demonstrate a tight link between energy metabolism, inflammation and demyelination. These studies also show variabilities in the extent of oligodendrocyte stress, which can vary from a downregulation of oligodendrocyte genes, such as observed in psychiatric disorders, to cell death and brain lesions seen in multiple sclerosis. We conclude that experimental models of multiple sclerosis could be of interest for the research of BPD, SZ and MDD.

Keywords: mood disorders, transcriptome, microarrays, myelin, white matter, oligodendrocytes, mitochondria, inflammation, schizophrenia, multiple sclerosis

Introduction

A subset of symptoms of bipolar disorder (BPD) overlaps with other psychiatric disorders such as major depressive disorder (MDD) and schizophrenia (SZ) (Berrettini, 2003; Moller, 2003). In concordance with the shared clinical symptoms, molecular and genetic findings show significant overlap across these disorders (Green et al., 2009; Huang et al., 2010; Knight et al., 2009). Thus, if we accept that any gene expression profile found in BPD might also be found in selected other psychiatric disorders and is not exclusive for BPD, a number of themes emerge from transcriptome studies. Given the complexities of psychiatric disorders, the overlap of findings in BPD, MDD, and SZ could be useful to correlate gene expression patterns with specific symptoms such as depression or psychosis, and advance the development of symptom-specific medications.

After an initial discussion of experimental limitations, essential for the proper consideration of the available literature, we will address the most consistent transcriptome findings in BPD. These findings point to abnormal bioenergetic function, myelin deficiencies, and increased activity of the immune system. Interestingly, this triad of pathological abnormalities is also the focus of multiple sclerosis (MS) research. We conclude that certain experimental models of MS are of potential use for the study of BPD and related psychiatric disorders.

Tissue sources for gene expression profiling

Human brain samples are the most proximal tissue for any kind of brain disorder, but are also the most difficult to obtain. Samples are collected post-mortem, and the lack of experimental control before samples enter a brain bank presents a number of limitations. These include significant lags between time of death and proper storage of tissue, heterogeneity in cause of death, adverse pre-mortem events, various medical interventions, advanced age of patients, comorbidity with other disorders, exposure to prescription drugs, dietary preferences, smoking habits, and potential addictions. Additionally, post-mortem samples cannot be collected in large numbers, limiting the statistical control for the influence of these complex factors on experimental outcomes.

Alternative tissue types that might be easier to collect on larger scales, and might permit tighter experimental control, are usually further removed from the brain. Blood samples and cell lines derived from various tissues such as lymphocytes or fibroblasts are the most frequent alternatives. These samples are collected from live patients at comparatively younger ages, are easier to harvest, and can be kept under defined culture conditions with ample time to overcome (‘wash out’) the effects of drug exposure and dietary influences. Peripheral cells can be transformed into cell lines, which can be multiplied and propagated, and the effects of various experimental manipulations such as drug treatments can be studied in parallel cultures of the same sample. Aside from valid questions regarding the comparability of gene expression patterns in peripheral tissues with gene expression patterns in brain tissue, general tissue culture issues have to be considered as well. These include the consequences of transforming cells programmed naturally for a finite number of mitotic divisions into cells that divide continuously, and the fact that over time cells with the fastest rate of division overgrow all other cells, creating a monoclonal culture that might not be representative of the original sample. Primary cultures, in which cells are not transformed, have a finite life span and limited quantities, but they can be helpful to circumvent the unwanted effects of cell immortalization.

The nature of psychiatric disorders makes them a uniquely human affliction, as these disorders affect complex behaviors that we are most adept to diagnose in humans. Animal models can be employed to determine the biological targets of therapeutic agents used to treat BPD and to study the downstream effects of pathological conditions on biochemical pathways and behaviors. However, post-mortem gene expression profiling is essential to identify these pathological changes modeled in animals.

Although there is no optimal approach to gene expression profiling in BPD, by combining different approaches and by considering their shortcomings in the interpretation of the data, we have obtained novel hypotheses about the pathophysiological mechanisms of BPD.

Studies of the transcriptome

The sequencing of the human genome and the associated coding sequences, together with the rapid progression of biological technologies have led to new hopes for a better understanding of psychiatric disorders (Horvath et al.; Mirnics et al., 2006). ‘Transcriptome’ refers to the entirety of messenger ribonucleic acids (mRNAs) in a particular tissue. It is assumed that an altered expression level of an mRNA in a disease is an indication for an abnormality in either the encoded protein, or a biologically connected structure or function. In BPD, transcriptomes have been most thoroughly studied with microarrays, though methodological approaches are constantly evolving. Levels of almost all known transcripts of the genome can be examined on a single microarray, a particularly attractive option for the study of polygenic disorders (Konradi, 2005). More recently, sequence-based approaches have begun to replace microarray studies. These ‘deep sequencing’ approaches sequence the entire transcriptome multiple times and collect information on sequence variants and frequency of transcripts (Wang et al., 2009).

Owing to their large scale, transcriptome studies yield a fair number of false-positive and false-negative findings. A number of corrective algorithms have been developed to deal with the problem of multiple testing such as the commonly used Benjamini-Hochberg correction (Benjamini and Hochberg, 1995). Large-scale gene expression approaches are not well suited for the study of a small number of genes of interest. The strength of transcriptome studies lies in the ability to examine expression levels of groups of functionally or anatomically related genes, as the likelihood that an entire group of related genes is concordantly regulated by chance is exceedingly small (Konradi, 2005). Programs have been developed that use existing databases to determine the expected versus observed rates of regulated genes mapped to predefined groups of related genes (see for example Dennis et al., 2003; Huang et al., 2009; Salomonis et al., 2007; Subramanian et al., 2005; Zhang et al., 2005).

While analysis of the entire transcriptome represented a shift from the conventional hypothesis-driven approaches and was criticized by some as ‘high-priced fishing expeditions’, these studies have extend the boundaries of current knowledge and generated novel hypotheses (Gibson, 2003).

Transcriptome studies have the potential to yield valuable clues to the underlying pathology in BPD, and they can help to define novel research strategies (Konradi, 2005). For example, a downregulation of mRNA transcripts encoding mitochondrial genes could indicate a loss of mitochondria or a loss of a cell type that has particularly large numbers of mitochondria. Alternatively, the same molecular profile could arise from dysregulation of a genetic ‘master switch’ that regulates a group of genes expressed in mitochondria, or from dysregulation of epigenetic mechanisms such as histone modifications and DNA methylation (Petronis, 2010). Initial findings can be followed up with additional experimental approaches to clarify and extend our knowledge base.

Major gene groups of interest in bipolar disorder: Rounding up the usual suspects

For many psychiatric disorders, including BPD, a number of neurotransmitter systems were initially suspected of playing a role in the disease mechanism. Early studies were guided by available technology and pharmacological observations, with considerable elements of serendipity. For example, compounds that were initially used as anesthetics, or as solvents for drugs, had surprising effects on symptoms of psychiatric disorders (Cade, 1949; Lopez-Munoz et al., 2005). These observations led to testable hypotheses but they also set boundaries and created narratives in which what could not be imagined or measured remained unexplored. Discovery-driven studies and novel technologies changed that picture and created interesting new models.

Initial studies focused on known neurotransmitter systems such as biogenic amines (Cousins et al., 2009; Manji and Lenox, 2000; Schildkraut, 1974), glutamate, and GABA (Krystal et al., 2002; Post, 1990), and on altered regulation of glucocorticoids (Daban et al., 2005; Watson et al., 2004), with implications of an involvement of each system. Studies with lithium led to the hypothesis of a role of the glycogen synthase kinase 3 beta pathway (Ikonomov and Manji, 1999; Klein and Melton, 1996), while studies with valproic acid introduced a role for epigenetic mechanisms in BPD (Phiel et al., 2001). Serotonin and norepinephrine became a focus because of their involvement in the mechanism of action of antidepressants (Owens, 2004). Overall, many different systems seemed to be implicated in BPD, with supportive evidence for each one.

Going beyond the usual suspects: Mitochondria, myelin and the immune system in BPD

In gene expression studies, groups of regulated genes are examined for shared biological, molecular, or structural functions as well as for mutual mechanisms of regulation. These studies are quite comprehensive and in the past led to novel discoveries, novel hypotheses and re-discovery of old hypotheses. Herein, we will focus on three gene groups that have raised particular attention in transcriptome studies: mitochondria, myelin and immune function (tables 1, 2). These systems are important throughout the brain and could explain the widespread pathological findings.

Table 1.

Publications of gene expression profiles in the brain of subjects with BPD.

Reference group # of
samples
mean age mean PMI gender tissue markers of energy
function
oligodendrocyte
markers
markers of immune
response/stress
collection
site
Platform
(Tkachev et al., 2003) BPD 15 42.3 32.5 9M, 6F PFC - BA 9 decreased levels
(BPD): OLIG2,
SOX10, GALC,
MAG, PLP1,
CLDN11, MOG,
ERBB3, TF


increased levels
(SZ): MPZL1;
decreased levels
(SZ): OLIG1,
OLIG2, SOX10,
GALC, MBP,
MAG, PLP1,
MOBP, CLDN11,
MOG, ERBB3, TF
Stanley
Foundation
Brain
Collection
Affymetrix
HG-U133A
SZ 15 44.2 33.7 9M, 6F
control 15 48.1 23.7 9M, 6F

(Konradi et al., 2004) BPD 9 65.4 ± 5.8 18.4 /− 2.3 * 4 M, 5 F Hippocampus decreased levels:
OXCT1, ATP5G3,
ATP5H, ATP5J2,
ATP5C1, ATP5O,
ATP6V1D, ATP6V1A,
COX7A2, COX7B,
DNM1L, FH, MRPL3,
NDUFAB1, SSBP1,
SLC25A4, UQCRQ,
UGP2, VDAC1P1,
VDAC1, VDAC2
Harvard
Brain
Tissue
Resource
Center
Affymetrix
HG-U95Av2
SZ 8 62 ± 6.4 18.9 ± 1.8 * 4 M, 4 F
control 10 66.1 ± 4.8 19 ± 1.4 * 5 M, 5 F

(Iwamoto et al., 2005) BPD 33 43.8 36.4 16 M, 17 F PFC - BA 46 a general
downregulation of
mitochondrial genes
was observed in BPD
and SZ
Stanley
Foundation
Array
Collection
Affymetrix
HG-U133A
SZ 35 42.6 31.4 26 M, 9 F
control 34 44.1 29.6 25 M, 9 F

(Ryan et al., 2006) BPD 10 43.5 ± 10.7 ** 33.8 ± 18.5 ** 5 M, 5 F OFC - BA 11 increased levels: AZU1,
C8A, CD7, CD72, CRIP1,
GBX2, HLA-E, IL1R2,
IL2RA, IL4, IL4R, IL5,
IL8RB, KIR2DS1, KNG1,
MS4A2, ORM1/ORM2,
PFC, PRLR, TNFRSF17;
decreased levels:
SOCS5, WWP1
Stanley
Foundation
Brain
Collection
Affymetrix
HG-U133A
control 11 47.8 ± 11 ** 23.1 ± 9.6 ** 7 M, 4 F

BPD 30 44.5 ± 10.7 ** 37.2 ± 17.7 ** 16 M, 14 F PFC - BA 9 no significant changes
detected
control 31 43.8 ± 7.3 ** 29.1 ± 13.1 ** 24 M, 7 F

(Sun et al., 2006) BPD 35 45 ± 11 ** 38 ± 18 ** 17 M, 18 F DLPFC (no BA
listed)
increased levels:
MMAA, HMGCS2,
MCEE, CLPX;
decreased levels:
COX6C, NDUFS8,
NDUFS7, COX5A,
ATP5C1, ATP5J,
ATP5G3, UQCRC2,
SLC25A16, GPX4,
PRSS25, DIABLO,
MPST, MSRB2,
TOMM40, MRPS12,
MCCC1, COQ7

Stanley
Foundation
Array
Collection
19K cDNA
Array
(University
Health
Network
Microarray
Centre,
Toronto)
control 35 44 ± 8 ** 29 ± 13 ** 26 M, 9 F

(Shao and Vawter, 2008) BPD 29 45.3 ± 9.8 ** 39.1 ± 17.9 ** 15 M, 12 F PFC - BA 46 increased levels: ERBB2,
IL17RB, IL2RA, JARID2,
LGALS3, NFATC1,
PBX4, PPARA, TXNIP;
decreased levels: NMU,
PVR
Stanley
Foundation
Array
Collection
Codelink 20K
oligonucleotide
microarrays
SZ 32 42.9 ± 8.6 ** 30.5 ± 15.1 ** 23 M, 9 F
control 27 44.4 ± 6.5 ** 28.9 ± 12.7 ** 23 M, 6 F

(Rao et al., 2010) BPD 10 49 ± 7.2 21 ± 3.0 not listed PFC (no BA
listed)
increased mRNA and
protein: IL-1beta, IL-1R,
myeloid differentiation
factor 88, NFkappaB
subunits
Harvard
Brain
Tissue
Resource
Center
western blots,
QPCR,
immunohisto-
chemistry
control 10 43 ± 3.5 27 ± 1.5
*

SEM;

**

SD;

BA = Brodmann area; PFC = prefrontal cortex; DLPFC = dorsolateral PFC; OFC = orbitofrontal cortex; for abbreviations see ‘list of abbreviations’

Table 2.

Publications of gene expression profiles in peripheral tissue of subjects with BPD.

Reference group # of
samples
mean age gender tissue markers of energy function markers of immune
response/stress
collection site Platform
(Washizuka et al., 2005) BPD I 13 52.5 ± 12.4 7 M, 6 F Lymphoblastoids decreased levels: NDUFA1,
NDUFV3, NDUFA6, NDUFS7,
COX6C
Brain Science
Institute, RIKEN,
Saitama
BPD II 8 57.0 ± 10.7 2 M, 6 F
control 11 51.1 ± 10.0 8 M, 3 F

(Naydenov et al., 2007) BPD 21 40.5 ± 2.3 10 M, 11 F Glucose-deprived
primary
lymphocytes in
culture
decreased levels in low
glucose: ATP5G2, ATP5L,
ATP5S, COX11, COX15,
COX4I1, COX7A2L, COX7C,
NDUFA5, NDUFA6, NDUFB1,
UQCRB, UQCRC2, UQCRH
McLean Hospital Affymetrix
HG-U133A
2.0
control 21 39.3 ± 2.9 12 M, 9F

(Padmos et al., 2008) BPD 42 42 16 M, 26 F Monocytes increased levels: PDE4B,
IL1B, IL6, TNF, TNFAIP3,
PTGS2, PTX3, BCL2A1,
EMP1


increased levels: PDE4B,
IL1B, IL6, TNF, TNFAIP3,
PTGS2, PTX3, BCL2A1,
EMP1
Erasmus Med.
Center,
University
Utrecht
Affymetrix
HG-U95Av2
(5 BPD
subjects, 6
controls); Q-
PCR (all
subjects)
control 25 40 11 M, 14 F
BPD offspring 54 18 26 M, 28 F
control-children 70 16 33 M, 37 F

(Padmos et al., 2009) BPD - twin
MZ/concordant
6 pairs 36 4 M, 8 F Monocytes increased levels: PDE4B,
IL1B, IL6, TNF, TNFAIP3,
PTGS2, PTX3, BCL2A1,
EMP1
Erasmus Med.
Center,
University
Utrecht
QPCR
BPD - twin
DZ/discordant
12 pairs 41 6 M, 18 F
BPD - twin
DZ/concordant
4 pairs 43 2 M, 6 F
BPD - twin
DZ/discordant
19 pairs 43 14 M, 24 F
control - twin
MZ
18 pairs 40 6 M, 30 F
control - twin
DZ
16 pairs 44 9 M, 23 F

Abnormal patterns of mitochondrial genes in BPD

Gene and protein expression profiles in BPD show a decrease of mRNA and proteins involved in mitochondrial functions such as oxidative phosphorylation (OXPHOS) (Iwamoto et al., 2005; Konradi et al., 2004; Pennington et al., 2008; Washizuka et al., 2005). Downregulations were observed in the prefrontal cortex (PFC; Brodmann areas 9 and 46, table 1), (Iwamoto et al., 2005; Pennington et al., 2008; Sun et al., 2006), hippocampus (Konradi et al., 2004) and in lymphoblastoid cell lines (table 2), (Washizuka et al., 2005). Under compromised OXPHOS, ATP levels decrease and the activity of ATP-dependent ion pumps such as the Na+/K+-ATPase is slowed down, causing the membrane potential to shift toward hypopolarization. Neurotransmitters are released, and their uptake is delayed. The consequences of ATP loss have been studied in models of ischemia and ensuing glutamate toxicity, providing insights into the pathological mechanisms that in extreme cases lead to cell death (Calabresi et al., 1995; Nieber, 1999; Santos et al., 1996).

A confound in post-mortem studies of mRNA levels for proteins involved in mitochondrial function is a causality dilemma between degradation of mRNA under lower pH levels (Vawter et al., 2006; Washizuka et al., 2005), and reduced mRNA levels for mitochondrial proteins reducing pH levels due to compromised OXPHOS (figure 1). Samples with low pH are considered to be affected by adverse agonal events and are excluded from RNA studies, which inadvertently leads to the exclusion of samples with genuine molecular abnormalities affecting bioenergetic function. Specifically, lower mRNA and protein levels of the electron transfer chain cause a shift from OXPHOS to glycolysis to retain ATP production, which leads to the accumulation of lactic acid, the acidification of tissue, and a decrease in pH (Clay et al., 2010). By excluding samples with lower pH, patients with pathologically reduced mitochondrial function are excluded. A correlation between mRNA levels of mitochondrial genes and pH has been demonstrated (Vawter et al., 2006), though cause and effect remain to be elucidated. Since low pH levels are also a function of agonal stress and should be observed throughout the brain (Vawter et al., 2006), we chose to match samples for pH levels in the cerebellum, while examining mRNA profiles in the hippocampus (Konradi et al., 2004). We reasoned that decreased mitochondrial mRNA levels in the hippocampus could not be due to agonal stress if pH levels in the cerebellum were in the normal range. By this approach, we found a large decrease of mRNA levels for proteins involved in OXPHOS in BPD.

Figure 1.

Figure 1

Relationship between electron transfer mRNA transcripts (ETTs), glycolysis and pH. A: In the normal functioning living cell, glycolysis feeds into the tricarboxylic acid (TCA) cycle followed by the electron transfer chain (gray boxes). Oxygen is consumed and water is produced. B: In a living cell with reduced expression of ETTs, the electron transfer chain and OXPHOS get compromised, causing a shift toward glycolysis. Lactic acid accumulates in the tissue and the pH drops. C: Agonal stress prior to death leads to a drop in pH of the brain tissue, which accelerates RNA degradation in the post-mortem tissue. Reduced levels of ETTs found in post-mortem brain could thus be either a function of agonal events, or an indication of inadequate synthesis of ETTs. ETTs - electron transfer mRNA transcripts; TCA - tricarboxylic acid cycle

In a transcriptome study of samples matched for pH, the downregulation of mitochondrial genes was still observed in the post-mortem PFC in BPD and SZ (Iwamoto et al., 2005). A study in cultured lymphocytes, which was not confounded by agonal stress, showed that cells from subjects with BPD responded to glucose deprivation with a downregulation of mRNA transcripts for the mitochondrial electron transfer chain, whereas cells from normal controls responded with a robust upregulation (Naydenov et al., 2007). This study showed that cells from BPD patients are unable to properly adjust bioenergetic output during glucose shortages, indicating problems with the molecular regulation of OXPHOS gene expression.

Corroborating in vivo evidence for mitochondrial involvement in BPD comes from magnetic resonance spectroscopy (MRS), an imaging technique that allows visualization of energy-related metabolite levels and pH in the brain. Kato et al. found reduced frontal lobe pH in medicated and unmedicated BPD patients (Kato et al., 1998; Kato et al., 1993), and two-dimensional proton echo-planar spectroscopic imaging of medication-free BPD patients showed a shift from OXPHOS to glycolysis (Dager et al., 2004), (figure 1).

Mitochondrial diseases are frequently comorbid with psychotic symptoms and misdiagnosed as BPD or SZ (Campos et al., 2001; Fattal et al., 2006; Grover et al., 2006; Mancuso et al., 2008; Prayson and Wang, 1998). Patients exhibiting classical mitochondrial disease symptoms in muscle tissue, such as ragged red fibers and electron transport chain deficiency, also commonly present with psychosis or depression (Fattal et al., 2006; Grover et al., 2006; Mancuso et al., 2008; Prayson and Wang, 1998). These observations make a particularly strong case for a role of mitochondria in the clinical symptoms of psychosis, since the primary disease-causing events in mitochondrial disorders are mutations in mitochondrial DNA or mRNAs coding for mitochondrial proteins.

In most studies of psychiatric populations, the majority of patients are on medication. Thus, medication effects on mitochondrial genes cannot be entirely excluded. In studies of medication effects, mood stabilizers provided partial protection for mitochondria under stress, whereas first-generation antipsychotic drugs had adverse effects on mitochondrial respiration (Bachmann et al., 2009; King et al., 2001; Maurer and Moller, 1997; Maurer et al., 2009; Pereira et al., 1992; Prince et al., 1997; Sagara, 1998; Struewing et al., 2007; Valvassori et al.; Washizuka et al., 2009). These findings suggest that mitochondrial pathology could be introduced by antipsychotic drugs, but not by mood stabilizers.

As with other studies, significant overlaps between BPD and SZ were seen. For example, Iwamoto et al. observed a tendency for the same expression changes in the two diseases and commented that the overall pattern of differential gene expression was very similar in BD and SZ (Iwamoto et al., 2005), In a study of hippocampal gene expression patterns, the downregulation of mitochondrial genes was restricted to subjects with BPD and was not observed in subjects with SZ (Konradi et al., 2004). These differences could be brain-area specific, or the result of false-negative or false-positive findings in SZ.

Abnormal patterns of oligodendrocyte markers in BPD

In the central nervous system, oligodendrocytes insulate long-range axons with myelin sheaths to form white matter tracts. Expression levels of oligodendrocyte-specific mRNAs were found to be downregulated in BPD and related disorders. For example, gene expression profiling of the PFC (Brodmann area 9) in 15 BPD samples from the Stanley Foundation showed a reduction of oligodendrocyte- and myelin-specific genes. These expression changes had a high degree of overlap with SZ (table 1), (Hakak et al., 2001; Tkachev et al., 2003). Gene expression profiling in the temporal cortex (Brodmann area 21) of patients with MDD, collected by the Stanley Foundation, also showed downregulation of oligodendrocyte-specific genes (Aston et al., 2005). Not surprisingly, the authors of the latter publication concluded that MDD may share common oligodendroglial abnormalities with SZ and BPD. Myelin abnormalities in BPD, SZ and MDD have been supported by histological observations as well as by imaging studies such as diffusion tensor imaging (for a review see Herring and Konradi, 2011), lending further credibility to the gene expression profiling observations.

Although oligodendrocyte abnormalities are not likely key to the pathology of all instances of BPD, they could be causal in some cases. For example, an increased incidence of psychiatric symptoms associated with BPD, MDD and anxiety disorders has been documented in multiple sclerosis (MS), an inflammatory disease of the central nervous system whose primary pathology is demyelination (Chwastiak and Ehde, 2007; Hogancamp et al., 1997). Although depression and anxiety might be normal responses to the diagnosis of a severe neurodegenerative disorder, this argument cannot explain the increased incidence of BPD in MS. Moreover, neuregulin (NRG), a gene involved in oligodendrocyte development and myelination of the central nervous system, is one of the genetic loci for BPD and SZ, (Taveggia et al., 2008). NRG belongs to a family of epidermal growth factor-like ligands that interact with ErbB receptor tyrosine kinases. A loss of ErbB signaling in transgenic mice leads to changes in oligodendrocyte number and morphology, increased levels of dopamine receptors and transporters, and behavioral alterations consistent with neuropsychiatric disorders (Roy et al., 2007). Although most genetic NRG studies were carried out in SZ (Stefansson et al., 2003; Stefansson et al., 2002), NRG1 haplotypes were also associated with psychotic BPD (Goes et al., 2009).

Because training and experience influence white matter structure, it is possible that myelin deficiencies may result from, or cause some, cases of psychiatric disorders. On the positive side, training of working memory leads to white matter changes and facilitates connectivity in the corpus callosum, the white matter tract that connects both cerebral hemispheres and other brain areas (Takeuchi et al., 2010). On the negative side, sensory deprivation in mice prevents the developmental increase in myelin mRNA transcripts (Lyckman et al., 2008). These observations suggest that hypoactive brain structures have decreased myelination. Although we do not know whether the reduction in oligodendrocyte markers observed in psychiatric illnesses is a cause or symptom of disease, it seems a reasonable assumption that disruptions of connectivity due to myelin deficiencies could contribute to the clinical symptoms in BPD and related disorders.

Gene expression patterns indicative of inflammatory processes

Transcriptome analysis in the PFC (Brodmann area 9) and the orbitofrontal cortex (Brodmann area 11) of BPD subjects showed an upregulation of immune response genes (table 1), (Ryan et al., 2006). Independent studies confirmed these findings in the frontal cortex (Brodmann area not mentioned Rao et al., 2010) and in Brodmann area 46 (PFC), where a similar signature was seen in SZ patients (Arion et al., 2007; Shao and Vawter, 2008). Gene expression analysis in the PFC (Brodmann area 10) of MDD yielded the same pattern (Shelton et al., 2010). Such gene expression changes could be due to duration of illness, as changes in inflammatory genes were found in SZ patients only with long-term illness (Narayan et al., 2008). Monocytes of BPD patients and offspring of BPD parents showed increased inflammatory gene expression (table 2), (Padmos et al., 2008), Comparisons of the presence of this gene expression signature between monozygotic and dizygotic twins revealed the signature to be primarily the result of environmental factors (Padmos et al., 2009).

The transcriptome studies emphasize the role of the immune system in BPD, SZ and MDD, a role that has long been hypothesized (Goldstein et al., 2009). The upregulation of genes involved in immune response indicates inflammatory processes in the brain and suggest that models of multiple sclerosis (MS) might provide some insights into the pathophysiology of psychiatric disorders. Although the pathogenesis, disease course, and clinical symptoms of MS differ from those of BPD, animal models and strain-specific vulnerabilities in mice (Lynch et al., 2010), might be useful to examine how the genetic environment influences inflammatory processes in the brain (see below).

Limitations of transcriptome studies of brain samples

Assuming that the technology itself is flawless - which no technology is - a number of limitations in transcriptome studies need to be considered when interpreting data. While these factors should be kept in mind, they should not detract from the valuable information that can be obtained from transcriptome studies.

False-positive and false-negative errors

Transcriptome studies, by virtue of the large number of genes analyzed, have a high number of false-positive and false-negative findings (Konradi, 2005). Although algorithms have been developed to reduce the probability of false-discoveries (for example Benjamini and Hochberg, 1995), the false-negative errors go mostly unnoticed. The strength of transcriptome studies is their ability to examine groups of genes of similar function or location, providing information on pathological processes. The likelihood that an entire group of functionally related genes reaches significance by chance is much lower than the likelihood for any individual gene. The inverse argument applies for false-negative findings, where it seems unlikely that an entire group of functionally related genes will be invisible in the analysis.

Limited experimental control over human brain samples

RNA is not particularly stable and is rapidly degraded by RNases, which are ubiquitous in all tissues. Factors that cannot be experimentally controlled in human brain samples, such as pre-mortem events, cause of death, and post-mortem events affect RNA integrity (see discussion on pH, above). Due to the limited availability of human brain samples, overall sample sizes are small enough to be sensitive to these unavoidable variabilities. Studies might find false-positive results, though the chances of false-negative results are likely equally as high. Transcriptome studies are a good example of an instance where the absence of evidence cannot be interpreted as evidence of absence.

Brain-specific genes will be over-represented in the group of regulated genes

Each tissue type has a specific profile of expressed genes. The list of regulated genes in the brain will naturally be biased toward neuronal and glial markers expressed in the brain area examined. Although excitement might be high when a favorite pathway or neurotransmitter system emerges in the analysis, proper statistical approaches that correct for tissue bias are crucial. It is therefore necessary to determine if regulated genes are over-representative of any particular brain function. Fisher’s exact test and variations thereof must be used to examine if the percentage of genes regulated in a specific category is statistically different from the percentage of genes regulated in all categories (for example, see Hosack et al., 2003).

Potential reasons for diverse findings in expression patterns

The relatively small sample sizes of post-mortem studies leads to inherent statistical errors, with false-positive and false-negative results. Still, the considerable overlap in the findings raises confidence in the value of these studies. A number of factors that might contribute to the variability between studies will be briefly discussed here.

First, it is likely that different brain areas have different disease signatures. Thus, it cannot be expected that all brain areas yield the same results. Second, differences in dissection methods could play a role. For example, abnormalities in oligodendrocyte genes might predominate in white matter and thus depend on the relative proportion of grey and white matter in a sample. Mitochondria predominate in synapses (neuropil), whereas glycolysis predominates in astroglia (Nehlig and Coles, 2007), further supporting the relevance of balancing grey and white matter. Thirdly, gender and race differences influence the data. Race differences are well known in mitochondria, where different haplogroups define genetic populations (Torroni and Wallace, 1994). Gender differences in brain structure, function and biology are also amply described (Cosgrove et al., 2007) and extend to gene expression profiles in the brain (Vawter et al., 2004). Fourthly, some gene expression profiles are likely the result of exposure to medication. Varying prescription practices, as well as geographically distinct dietary habits could distort the gene expression patterns. Finally, given the overlap between BPD, SZ and MDD, disease symptoms may be interpreted differently and nosology might vary (Blacker and Tsuang, 1992).

The common thread between myelin, mitochondria and inflammation

Multiple sclerosis research has long focused on pathological processes that link myelin, inflammation and mitochondrial function (figure 2), (Kalman et al., 2007). Experimental MS models can potentially provide insight into the pathophysiological mechanisms of BPD. For example, the copper chelator cuprizone (CPZ) induces demyelination in mice and serves as a murine model for MS (Herring and Konradi, 2011). We have shown that CPZ reduces the expression of oligodendrocyte genes in the rat PFC, and that this reduction is accompanied by deficits in PFC-mediated cognitive behaviors (Gregg et al., 2009). By carefully selecting the age of exposure – adolescence - and the concentration of CPZ, the PFC was specifically targeted without any effect on other brain areas and no decline in motor function. This study showed that MS-inducing drugs can affect higher cognitive function in the absence of motor deficits, presumably modulated by genetic background, force of impact, i.e. dose of CPZ, and periods of heightened vulnerability of different brain areas.

Figure 2.

Figure 2

Hypothesis connecting mitochondria, myelin and immune function. Multiple sclerosis models have shown that mitochondria and the immune system play a role in demyelination, and that the genetic environment modulates the severity of oligodendrocyte stress and demyelination. Thus, under certain conditions mRNA levels of oligodendrocyte genes can be downregulated without causing overt cell death and MS symptoms. The downregulation of oligodendrocyte genes might be an indication of cell stress in BPD. Transcriptome studies have shown that genes involved in myelination, inflammation and energy metabolism are dysregulated in BPD. We hypothesize that these systems interact to reduce myelination and thereby hypopolarize the neuronal membrane potential, making it more susceptible to depolarization. Neuronal function and network activities get compromised, leading to symptoms associated with BPD and other psychiatric conditions.

CPZ selectively injures oligodendrocytes (Matsushima and Morell, 2001). Inflammatory processes and mitochondrial stress are required to mediate oligodendrocyte toxicity by CPZ. CPZ causes the formation of megamitochondria in the liver, which are formed in the presence of free radicals and can lead to apoptosis (Wakabayashi, 2002). Transgenic mice lacking neuronal nitric oxide synthase, an inhibitor of the electron transfer chain and inducing agent of oxidative stress, are relatively resistant to CPZ-induced demyelination (Linares et al., 2006).

In primary oligodendrocyte cultures, inflammatory cytokines are required for CPZ-mediated cell death (Pasquini et al., 2007), demonstrating the involvement of the immune system in cell death. Transgenic mice lacking the type 2 CXC chemokine receptor (CXCR2), expressed by neutrophil cells of the immune system, are resistant to CPZ-induced demyelination, although their initial response to CPZ is similar to the response of wild-type mice (Liu et al., 2010). Whereas wild-type mice progress from an initial downregulation of oligodendrocyte-specific genes to apoptosis and widespread demyelination, CXCR2 knockout mice recover from the initial insult with most of their oligodendrocytes spared. These observations demonstrate that the energy metabolism and the immune system control the mechanism and magnitude of demyelination.

Pathological connections between myelin loss, mitochondria and inflammation have been observed in other MS models as well. For example, the experimental autoimmune encephalomyelitis (EAE) model has been used for a long time to model MS, and it is well known that mitochondria and inflammation play a significant role in this model (Gold et al., 2006). Mice lacking cyclophilin D, a key regulator of the mitochondrial permeability transition pore, partially recovered in the EAE model, whereas wild-type mice progressed in the disease (Forte et al., 2007). Activation of the mitochondrial permeability transition pore seemed a critical factor in the disease progression. Similarly, suppressing the activation of microglia repressed the development of EAE, confirming a role for the immune system (Heppner et al., 2005).

MS models show that the genetic environment determines the extent of myelin stress and the range of pathological consequences (deLuca et al., 2010; Piehl and Olsson, 2009). Thus, the combination of polymorphisms in genes of immune system proteins, and/or genes of proteins involved in bioenergetic function, could determine which brain areas are affected, and to what extent (figure 2).

Concluding remarks

Despite many confounding factors, transcriptome profiling has produced results that have been verified in multiple studies. Similar molecular signatures have been observed in BPD, SZ and MDD, generating novel hypotheses on the pathophysiology of BPD and related disorders. The three factors we focused on, mitochondria, myelination and inflammation, are systemic in the brain and not circumscribed like the neurotransmitter systems of previous hypotheses. We propose that variant environmental influences together with genetic predispositions determine the degree to which different brain areas and functions are affected in individual patients, thus leading to variations in psychiatric symptoms and diagnoses. Introducing these hypotheses does not negate previous hypotheses, but can interdigitate with them. Mitochondrial vulnerabilities and myelination defects could be the disease cause in some cases, and a consequence of the disease processes in others. Similarly, inflammation could be secondary to demyelination and/or apoptotic processes in cells under mitochondrial stress, or it could be the primary cause of demyelination triggered by viral infections (Hart et al., 1999; Karlsson et al., 2001). Although we do not propose that BPD, SZ and MDD are variants of MS, the large knowledge base of MS research might provide some interesting and novel insights into disease mechanisms and treatment approaches in these disorders. Transcriptome studies have played an essential role in the development of these hypotheses, and further research of gene expression profiles in BPD will be helpful to deepen and expand our knowledge.

Acknowledgement

The work was supported by MH084131 and MH67999. The content is solely the responsibility of the authors and does not necessarily represent the official views of the funding institutes or the National Institutes of Health.

Abbreviations

BPD

bipolar disorder

SZ

schizophrenia

MDD

major depressive disorder

OXPHOS

oxidative phosphorylation

ETTs

electron transfer mRNA transcripts

TCA cycle

tricarboxylic acid cycle

Gene Symbol/Gene Name

ATP5C1

ATP synthase gamma chain

ATP5G2

ATP synthase, c (subunit 9) isoform 2

ATP5G3

ATP synthase, c (subunit 9) isoform 3

ATP5H

ATP synthase, subunit d

ATP5J

ATP synthase subunit f6

ATP5J2

ATP synthase, subunit f, isoform 2

ATP5L

ATP synthase, g

ATP5O

ATP synthase, subunit O (OSCP)

ATP5S

ATP synthase, subunit s (factor B)

ATP6V1A

ATPase, lysosomal 70kDa, V1 subunit A, isoform 1

ATP6V1D

ATPase, lysosomal 34kDa, V1 subunit D

AZU1

Azurocidin 1 (cationic antimicrobial protein 37)

BCL2A1

BCL2-related protein A1

C8A

Complement component 8, alpha polypeptide

CD7

CD7 antigen (p41)

CD72

CD72 antigen

CLDN11

Claudin 11

CLPX

ClpX caseinolytic peptidase X

COQ7

coenzyme Q7 homolog, ubiquinone

COX11

COX11

COX15

COX15

COX4I1

Cytochrome c oxidase IV-isoform 1

COX5A

Cytochrome c oxidase polypeptide Va

COX6C

Cytochrome c oxidase polypeptide Vic

COX7A2

Cytochrome c oxidase subunit VIIa polypeptide 2 (liver)

COX7A2L

Cytochrome c oxidase VIIa 2 like

COX7B

Cytochrome c oxidase subunit VIIb

COX7C

Cytochrome c oxidase VIIc

CRIP1

Cysteine-rich protein 1 (intestinal)

DIABLO

Diablo homolog

DNM1L

Dynamin 1-like

EMP1

Epithelial membrane protein 1

ERBB2

v-erb-b2 erythroblastic leukemia viral oncogene homolog 2

ERBB3

Neuregulin receptor tyrosine kinase

FH

Fumarate hydratase

GALC

Galactosylceramidase

GBX2

Gastrulation brain homeo box 2

GPX4

Glutathione peroxidase 4

HLA-E

Major histocompatibility complex, class I, E

HMGCS2

3-hydroxy-3-methylglutaryl-CoA synthase 2

IL17RB

Interleukin 17 receptor B

IL1B

Interleukin 1 beta

IL1R2

Interleukin 1 receptor, type II

IL2RA

Interleukin 2 receptor, alpha

IL4

Interleukin 4

IL4R

Interleukin 4 receptor

IL5

Interleukin 5 (colony-stimulating factor, eosinophil)

IL6

Interleukin 6

IL8RB

Interleukin 8 receptor, beta

JARID2

Jumonji, AT rich interactive domain 2

KIR2DS1

Killer cell immunoglobulin-like receptor, two domains, short cytoplasmic tail, 1

KNG1

Kininogen 1

LGALS3

Lectin, galactoside-binding, soluble, 3

MAG

Myelin associated glycoprotein

MBP

Myelin basic protein

MCCC1

Methylcrotonoyl-CoA carboxylase 1 (alpha)

MCEE

Methylmalonyl CoA epimerase

MMAA

Methylmalonic aciduria (cobalamin deficiency) cblA typ

MOBP

Myelin-associated oligodendrocyte basic protein

MOG

Myelin oligodendrocyte glycoprotein

MPST

Mercaptopyruvate sulfurtransferase

MPZL1

Myelin protein zero-like 1

MRPL3

Mitochondrial ribosomal protein L3

MRPS12

Mitochondrial ribosomal protein S12

MS4A2

Fc fragment of IgE, high affinity I, receptor for; beta polypeptide

MSRB2

Methionine sulfoxide reductase B2

NDUFA1

NADH dehydrogenase 1, alpha, 1

NDUFA5

NADH dehydrogenase 1, alpha, 5

NDUFA6

NADH dehydrogenase 1, alpha, 6

NDUFAB1

NADH dehydrogenase 1, alpha/beta subcomplex, 1

NDUFB1

NADH dehydrogenase 1, beta, 1

NDUFS7

NADH dehydrogenase Fe-S protein 7

NDUFS8

NADH dehydrogenase Fe-S protein 8

NDUFV3

NADH dehydrogenase 1, flavoprotein 3

NFATC1

Nuclear factor of activated T-cells, cytoplasmic, calcineurin-dependent 1

NMU

Neuromedin U

OLIG1

Oligodendrocyte lineage transcription factor 1

OLIG2

Oligodendrocyte lineage transcription factor 2

ORM1/ORM2

Orosomucoid 1, orosomucoid 2

OXCT1

3-oxoacid CoA transferase

PBX4

Pre-B-cell leukemia homeobox 4

PDE4B

Phosphodiesterase 4B, cAMP specific

PFC

Properdin P factor, complement

PLP1

Lipophilin, primary constituent of myelin, proteolipid protein

PPARA

Peroxisome proliferator activated receptor alpha

PRLR

Prolactin receptor

PRSS25

HtrA serine peptidase 2

PTGS2

Prostaglandin-endoperoxide synthase 2

PTX3

Pentraxin 3, long

PVR

PDGF- and VEGF-receptor related

SLC25A16

Solute carrier family 25, member 16 (mitochondrial)

SLC25A4

Solute carrier family 25, member 4 (mitochondrial)

SOCS5

Suppressor of cytokine signaling 5

SOX10

Transcription factor, regulates myelin-related genes, myelination deficiency

SSBP1

Single-stranded DNA binding protein

TF

Transferrin, transports iron, involved in myelin maintenance/formation

TNF

Tumor necrosis factor

TNFAIP3

Tumor necrosis factor, alpha-induced protein 3

TNFRSF17

Tumor necrosis factor receptor superfamily, member 17

TOMM40

Translocase of outer mitochondrial membrane 40

TXNIP

Thioredoxin interacting protein

UGP2

UDP-glucose pyrophosphorylase 2

UQCRB

Ubiquinol-cytochrome c reductase binding protein

UQCRC2

Ubiquinol-cytochrome c reductase complex core protein 2

UQCRH

Ubiquinol-cytochrome c reductase hinge protein

UQCRQ

Ubiquinol-cytochrome c reductase, complex III subunit VII

VDAC1

Voltage-dependent anion channel 1; porin

VDAC1P1

VDAC1 pseudogene, porin protein, isoform 1

VDAC2

Voltage-dependent anion channel 2; porin

WWP1

WW domain containing E3 ubiquitin protein ligase 1

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Arion D, Unger T, Lewis DA, Levitt P, Mirnics K. Molecular evidence for increased expression of genes related to immune and chaperone function in the prefrontal cortex in schizophrenia. Biol Psychiatry. 2007;62:711–21. doi: 10.1016/j.biopsych.2006.12.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Aston C, Jiang L, Sokolov BP. Transcriptional profiling reveals evidence for signaling and oligodendroglial abnormalities in the temporal cortex from patients with major depressive disorder. Mol Psychiatry. 2005;10:309–22. doi: 10.1038/sj.mp.4001565. [DOI] [PubMed] [Google Scholar]
  3. Bachmann RF, et al. Common effects of lithium and valproate on mitochondrial functions: protection against methamphetamine-induced mitochondrial damage. Int J Neuropsychopharmacol. 2009;12:805–22. doi: 10.1017/S1461145708009802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Benjamini Y, Hochberg Y. Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing. Journal of the Royal Statistical Society. Series B (Methodological) 1995;57:289–300. [Google Scholar]
  5. Berrettini W. Evidence for shared susceptibility in bipolar disorder and schizophrenia. Am J Med Genet C Semin Med Genet. 2003;123C:59–64. doi: 10.1002/ajmg.c.20014. [DOI] [PubMed] [Google Scholar]
  6. Blacker D, Tsuang MT. Contested boundaries of bipolar disorder and the limits of categorical diagnosis in psychiatry. Am J Psychiatry. 1992;149:1473–83. doi: 10.1176/ajp.149.11.1473. [DOI] [PubMed] [Google Scholar]
  7. Cade JF. Lithium salts in the treatment of psychotic excitement. Med J Aust. 1949;2:349–52. doi: 10.1080/j.1440-1614.1999.06241.x. [DOI] [PubMed] [Google Scholar]
  8. Calabresi P, Pisani A, Mercuri NB, Bernardi G. On the mechanisms underlying hypoxia-induced membrane depolarization in striatal neurons. Brain. 1995;118(Pt 4):1027–38. doi: 10.1093/brain/118.4.1027. [DOI] [PubMed] [Google Scholar]
  9. Campos Y, et al. Mitochondrial myopathy, cardiomyopathy and psychiatric illness in a Spanish family harbouring the mtDNA 3303C > T mutation. J Inherit Metab Dis. 2001;24:685–7. doi: 10.1023/a:1012719211505. [DOI] [PubMed] [Google Scholar]
  10. Chwastiak LA, Ehde DM. Psychiatric issues in multiple sclerosis. Psychiatr Clin North Am. 2007;30:803–17. doi: 10.1016/j.psc.2007.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Clay HB, Sillivan S, Konradi C. Mitochondrial dysfunction and pathology in bipolar disorder and schizophrenia. Int J Dev Neurosci. 2010 doi: 10.1016/j.ijdevneu.2010.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cosgrove KP, Mazure CM, Staley JK. Evolving knowledge of sex differences in brain structure, function, and chemistry. Biol Psychiatry. 2007;62:847–55. doi: 10.1016/j.biopsych.2007.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cousins DA, Butts K, Young AH. The role of dopamine in bipolar disorder. Bipolar Disord. 2009;11:787–806. doi: 10.1111/j.1399-5618.2009.00760.x. [DOI] [PubMed] [Google Scholar]
  14. Daban C, Vieta E, Mackin P, Young AH. Hypothalamic-pituitary-adrenal axis and bipolar disorder. Psychiatr Clin North Am. 2005;28:469–80. doi: 10.1016/j.psc.2005.01.005. [DOI] [PubMed] [Google Scholar]
  15. Dager SR, et al. Brain metabolic alterations in medication-free patients with bipolar disorder. Arch Gen Psychiatry. 2004;61:450–8. doi: 10.1001/archpsyc.61.5.450. [DOI] [PubMed] [Google Scholar]
  16. deLuca LE, et al. Substrain differences reveal novel disease-modifying gene candidates that alter the clinical course of a rodent model of multiple sclerosis. J Immunol. 2010;184:3174–85. doi: 10.4049/jimmunol.0902881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Dennis G, Jr., et al. DAVID: Database for Annotation, Visualization, and Integrated Discovery. Genome Biol. 2003;4:P3. [PubMed] [Google Scholar]
  18. Fattal O, Budur K, Vaughan AJ, Franco K. Review of the literature on major mental disorders in adult patients with mitochondrial diseases. Psychosomatics. 2006;47:1–7. doi: 10.1176/appi.psy.47.1.1. [DOI] [PubMed] [Google Scholar]
  19. Forte M, et al. Cyclophilin D inactivation protects axons in experimental autoimmune encephalomyelitis, an animal model of multiple sclerosis. Proc Natl Acad Sci U S A. 2007;104:7558–63. doi: 10.1073/pnas.0702228104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Gibson G. Microarray analysis: genome-scale hypothesis scanning. PLoS Biol. 2003;1:E15. doi: 10.1371/journal.pbio.0000015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Goes FS, et al. Family-based association study of Neuregulin 1 with psychotic bipolar disorder. Am J Med Genet B Neuropsychiatr Genet. 2009;150B:693–702. doi: 10.1002/ajmg.b.30895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Gold R, Linington C, Lassmann H. Understanding pathogenesis and therapy of multiple sclerosis via animal models: 70 years of merits and culprits in experimental autoimmune encephalomyelitis research. Brain. 2006;129:1953–71. doi: 10.1093/brain/awl075. [DOI] [PubMed] [Google Scholar]
  23. Goldstein BI, Kemp DE, Soczynska JK, McIntyre RS. Inflammation and the phenomenology, pathophysiology, comorbidity, and treatment of bipolar disorder: a systematic review of the literature. J Clin Psychiatry. 2009;70:1078–90. doi: 10.4088/JCP.08r04505. [DOI] [PubMed] [Google Scholar]
  24. Green EK, et al. The bipolar disorder risk allele at CACNA1C also confers risk of recurrent major depression and of schizophrenia. Mol Psychiatry. 2009 doi: 10.1038/mp.2009.49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Gregg JR, Herring NR, Naydenov AV, Hanlin RP, Konradi C. Downregulation of oligodendrocyte transcripts is associated with impaired prefrontal cortex function in rats. Schizophr Res. 2009;113:277–87. doi: 10.1016/j.schres.2009.05.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Grover S, et al. Mania as a first presentation in mitochondrial myopathy. Psychiatry Clin Neurosci. 2006;60:774–5. doi: 10.1111/j.1440-1819.2006.01599.x. [DOI] [PubMed] [Google Scholar]
  27. Hakak Y, et al. Genome-wide expression analysis reveals dysregulation of myelination-related genes in chronic schizophrenia. Proc Natl Acad Sci U S A. 2001;98:4746–51. doi: 10.1073/pnas.081071198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Hart DJ, et al. Antiretroviral antibodies: implications for schizophrenia, schizophrenia spectrum disorders, and bipolar disorder. Biol Psychiatry. 1999;45:704–14. doi: 10.1016/s0006-3223(98)00229-7. [DOI] [PubMed] [Google Scholar]
  29. Heppner FL, et al. Experimental autoimmune encephalomyelitis repressed by microglial paralysis. Nat Med. 2005;11:146–52. doi: 10.1038/nm1177. [DOI] [PubMed] [Google Scholar]
  30. Herring NR, Konradi C. Myelin, copper, and the cuprizone model of schizophrenia. Front Biosci. 2011;S3:23–40. doi: 10.2741/s129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Hogancamp WE, Rodriguez M, Weinshenker BG. The epidemiology of multiple sclerosis. Mayo Clin Proc. 1997;72:871–8. doi: 10.4065/72.9.871. [DOI] [PubMed] [Google Scholar]
  32. Horvath S, Janka Z, Mirnics K. Analyzing Schizophrenia by DNA Microarrays. Biol Psychiatry. 2010 doi: 10.1016/j.biopsych.2010.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Hosack DA, Dennis G, Jr., Sherman BT, Lane HC, Lempicki RA. Identifying biological themes within lists of genes with EASE. Genome Biol. 2003;4:R70. doi: 10.1186/gb-2003-4-10-r70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Huang DW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009;4:44–57. doi: 10.1038/nprot.2008.211. [DOI] [PubMed] [Google Scholar]
  35. Huang J, et al. Cross-Disorder Genomewide Analysis of Schizophrenia, Bipolar Disorder, and Depression. Am J Psychiatry. 2010 doi: 10.1176/appi.ajp.2010.09091335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Ikonomov OC, Manji HK. Molecular mechanisms underlying mood stabilization in manic-depressive illness: the phenotype challenge. Am J Psychiatry. 1999;156:1506–14. doi: 10.1176/ajp.156.10.1506. [DOI] [PubMed] [Google Scholar]
  37. Iwamoto K, Bundo M, Kato T. Altered expression of mitochondria-related genes in postmortem brains of patients with bipolar disorder or schizophrenia, as revealed by large-scale DNA microarray analysis. Hum Mol Genet. 2005;14:241–53. doi: 10.1093/hmg/ddi022. [DOI] [PubMed] [Google Scholar]
  38. Kalman B, Laitinen K, Komoly S. The involvement of mitochondria in the pathogenesis of multiple sclerosis. J Neuroimmunol. 2007;188:1–12. doi: 10.1016/j.jneuroim.2007.03.020. [DOI] [PubMed] [Google Scholar]
  39. Karlsson H, et al. Retroviral RNA identified in the cerebrospinal fluids and brains of individuals with schizophrenia. Proc Natl Acad Sci U S A. 2001;98:4634–9. doi: 10.1073/pnas.061021998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kato T, et al. Decreased brain intracellular pH measured by 31P-MRS in bipolar disorder: a confirmation in drug-free patients and correlation with white matter hyperintensity. Eur Arch Psychiatry Clin Neurosci. 1998;248:301–6. doi: 10.1007/s004060050054. [DOI] [PubMed] [Google Scholar]
  41. Kato T, Takahashi S, Shioiri T, Inubushi T. Alterations in brain phosphorous metabolism in bipolar disorder detected by in vivo 31P and 7Li magnetic resonance spectroscopy. J Affect Disord. 1993;27:53–9. doi: 10.1016/0165-0327(93)90097-4. [DOI] [PubMed] [Google Scholar]
  42. King TD, Bijur GN, Jope RS. Caspase-3 activation induced by inhibition of mitochondrial complex I is facilitated by glycogen synthase kinase-3beta and attenuated by lithium. Brain Res. 2001;919:106–14. doi: 10.1016/s0006-8993(01)03005-0. [DOI] [PubMed] [Google Scholar]
  43. Klein PS, Melton DA. A molecular mechanism for the effect of lithium on development. Proc Natl Acad Sci U S A. 1996;93:8455–9. doi: 10.1073/pnas.93.16.8455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Knight HM, et al. A cytogenetic abnormality and rare coding variants identify ABCA13 as a candidate gene in schizophrenia, bipolar disorder, and depression. Am J Hum Genet. 2009;85:833–46. doi: 10.1016/j.ajhg.2009.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Konradi C. Gene expression microarray studies in polygenic psychiatric disorders: applications and data analysis. Brain Res Brain Res Rev. 2005;50:142–55. doi: 10.1016/j.brainresrev.2005.05.004. [DOI] [PubMed] [Google Scholar]
  46. Konradi C, et al. Molecular evidence for mitochondrial dysfunction in bipolar disorder. Arch Gen Psychiatry. 2004;61:300–8. doi: 10.1001/archpsyc.61.3.300. [DOI] [PubMed] [Google Scholar]
  47. Krystal JH, et al. Glutamate and GABA systems as targets for novel antidepressant and mood-stabilizing treatments. Mol Psychiatry. 2002;7(Suppl 1):S71–80. doi: 10.1038/sj.mp.4001021. [DOI] [PubMed] [Google Scholar]
  48. Linares D, et al. Neuronal nitric oxide synthase plays a key role in CNS demyelination. J Neurosci. 2006;26:12672–81. doi: 10.1523/JNEUROSCI.0294-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Liu L, et al. CXCR2-positive neutrophils are essential for cuprizone-induced demyelination: relevance to multiple sclerosis. Nat Neurosci. 2010;13:319–26. doi: 10.1038/nn.2491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Lopez-Munoz F, et al. History of the discovery and clinical introduction of chlorpromazine. Ann Clin Psychiatry. 2005;17:113–35. doi: 10.1080/10401230591002002. [DOI] [PubMed] [Google Scholar]
  51. Lyckman AW, et al. Gene expression patterns in visual cortex during the critical period: synaptic stabilization and reversal by visual deprivation. Proc Natl Acad Sci U S A. 2008;105:9409–14. doi: 10.1073/pnas.0710172105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lynch RM, et al. Identifying genetic loci and spleen gene coexpression networks underlying immunophenotypes in BXD recombinant inbred mice. Physiol Genomics. 2010 doi: 10.1152/physiolgenomics.00020.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Mancuso M, et al. Autosomal dominant psychiatric disorders and mitochondrial DNA multiple deletions: report of a family. J Affect Disord. 2008;106:173–7. doi: 10.1016/j.jad.2007.05.016. [DOI] [PubMed] [Google Scholar]
  54. Manji HK, Lenox RH. Signaling: cellular insights into the pathophysiology of bipolar disorder. Biol Psychiatry. 2000;48:518–30. doi: 10.1016/s0006-3223(00)00929-x. [DOI] [PubMed] [Google Scholar]
  55. Matsushima GK, Morell P. The neurotoxicant, cuprizone, as a model to study demyelination and remyelination in the central nervous system. Brain Pathol. 2001;11:107–16. doi: 10.1111/j.1750-3639.2001.tb00385.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Maurer I, Moller HJ. Inhibition of complex I by neuroleptics in normal human brain cortex parallels the extrapyramidal toxicity of neuroleptics. Mol Cell Biochem. 1997;174:255–9. [PubMed] [Google Scholar]
  57. Maurer IC, Schippel P, Volz HP. Lithium-induced enhancement of mitochondrial oxidative phosphorylation in human brain tissue. Bipolar Disord. 2009;11:515–22. doi: 10.1111/j.1399-5618.2009.00729.x. [DOI] [PubMed] [Google Scholar]
  58. Mirnics K, Levitt P, Lewis DA. Critical appraisal of DNA microarrays in psychiatric genomics. Biol Psychiatry. 2006;60:163–76. doi: 10.1016/j.biopsych.2006.02.003. [DOI] [PubMed] [Google Scholar]
  59. Moller HJ. Bipolar disorder and schizophrenia: distinct illnesses or a continuum? J Clin Psychiatry. 2003;64(Suppl 6):23–7. discussion 28. [PubMed] [Google Scholar]
  60. Narayan S, et al. Molecular profiles of schizophrenia in the CNS at different stages of illness. Brain Res. 2008;1239:235–48. doi: 10.1016/j.brainres.2008.08.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Naydenov AV, MacDonald ML, Ongur D, Konradi C. Differences in lymphocyte electron transport gene expression levels between subjects with bipolar disorder and normal controls in response to glucose deprivation stress. Arch Gen Psychiatry. 2007;64:555–64. doi: 10.1001/archpsyc.64.5.555. [DOI] [PubMed] [Google Scholar]
  62. Nehlig A, Coles JA. Cellular pathways of energy metabolism in the brain: is glucose used by neurons or astrocytes? Glia. 2007;55:1238–50. doi: 10.1002/glia.20376. [DOI] [PubMed] [Google Scholar]
  63. Nieber K. Hypoxia and neuronal function under in vitro conditions. Pharmacol Ther. 1999;82:71–86. doi: 10.1016/s0163-7258(98)00061-8. [DOI] [PubMed] [Google Scholar]
  64. Owens MJ. Selectivity of antidepressants: from the monoamine hypothesis of depression to the SSRI revolution and beyond. J Clin Psychiatry. 2004;65(Suppl 4):5–10. [PubMed] [Google Scholar]
  65. Padmos RC, et al. A discriminating messenger RNA signature for bipolar disorder formed by an aberrant expression of inflammatory genes in monocytes. Arch Gen Psychiatry. 2008;65:395–407. doi: 10.1001/archpsyc.65.4.395. [DOI] [PubMed] [Google Scholar]
  66. Padmos RC, et al. Genetic and environmental influences on pro-inflammatory monocytes in bipolar disorder: a twin study. Arch Gen Psychiatry. 2009;66:957–65. doi: 10.1001/archgenpsychiatry.2009.116. [DOI] [PubMed] [Google Scholar]
  67. Pasquini LA, et al. The neurotoxic effect of cuprizone on oligodendrocytes depends on the presence of pro-inflammatory cytokines secreted by microglia. Neurochem Res. 2007;32:279–92. doi: 10.1007/s11064-006-9165-0. [DOI] [PubMed] [Google Scholar]
  68. Pennington K, et al. Prominent synaptic and metabolic abnormalities revealed by proteomic analysis of the dorsolateral prefrontal cortex in schizophrenia and bipolar disorder. Mol Psychiatry. 2008;13:1102–17. doi: 10.1038/sj.mp.4002098. [DOI] [PubMed] [Google Scholar]
  69. Pereira RS, Bertocchi AP, Vercesi AE. Protective effect of trifluoperazine on the mitochondrial damage induced by Ca2+ plus prooxidants. Biochem Pharmacol. 1992;44:1795–801. doi: 10.1016/0006-2952(92)90074-s. [DOI] [PubMed] [Google Scholar]
  70. Petronis A. Epigenetics as a unifying principle in the aetiology of complex traits and diseases. Nature. 2010;465:721–7. doi: 10.1038/nature09230. [DOI] [PubMed] [Google Scholar]
  71. Phiel CJ, et al. Histone deacetylase is a direct target of valproic acid, a potent anticonvulsant, mood stabilizer, and teratogen. J Biol Chem. 2001;276:36734–41. doi: 10.1074/jbc.M101287200. [DOI] [PubMed] [Google Scholar]
  72. Piehl F, Olsson T. Inflammation and susceptibility to neurodegeneration: the use of unbiased genetics to decipher critical regulatory pathways. Neuroscience. 2009;158:1143–50. doi: 10.1016/j.neuroscience.2008.08.031. [DOI] [PubMed] [Google Scholar]
  73. Post RM. Non-lithium treatment for bipolar disorder. J Clin Psychiatry. 1990;51(Suppl):9–16. discussion 17-9. [PubMed] [Google Scholar]
  74. Prayson RA, Wang N. Mitochondrial myopathy, encephalopathy, lactic acidosis, and strokelike episodes (MELAS) syndrome: an autopsy report. Arch Pathol Lab Med. 1998;122:978–81. [PubMed] [Google Scholar]
  75. Prince JA, Yassin MS, Oreland L. Neuroleptic-induced mitochondrial enzyme alterations in the rat brain. J Pharmacol Exp Ther. 1997;280:261–7. [PubMed] [Google Scholar]
  76. Rao JS, Harry GJ, Rapoport SI, Kim HW. Increased excitotoxicity and neuroinflammatory markers in postmortem frontal cortex from bipolar disorder patients. Mol Psychiatry. 2010;15:384–92. doi: 10.1038/mp.2009.47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Roy K, et al. Loss of erbB signaling in oligodendrocytes alters myelin and dopaminergic function, a potential mechanism for neuropsychiatric disorders. Proc Natl Acad Sci U S A. 2007;104:8131–6. doi: 10.1073/pnas.0702157104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Ryan MM, et al. Gene expression analysis of bipolar disorder reveals downregulation of the ubiquitin cycle and alterations in synaptic genes. Mol Psychiatry. 2006;11:965–78. doi: 10.1038/sj.mp.4001875. [DOI] [PubMed] [Google Scholar]
  79. Sagara Y. Induction of reactive oxygen species in neurons by haloperidol. J Neurochem. 1998;71:1002–12. doi: 10.1046/j.1471-4159.1998.71031002.x. [DOI] [PubMed] [Google Scholar]
  80. Salomonis N, et al. GenMAPP 2: new features and resources for pathway analysis. BMC Bioinformatics. 2007;8:217. doi: 10.1186/1471-2105-8-217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Santos MS, Moreno AJ, Carvalho AP. Relationships between ATP depletion, membrane potential, and the release of neurotransmitters in rat nerve terminals. An in vitro study under conditions that mimic anoxia, hypoglycemia, and ischemia. Stroke. 1996;27:941–50. doi: 10.1161/01.str.27.5.941. [DOI] [PubMed] [Google Scholar]
  82. Schildkraut JJ. Biogenic amines and affective disorders. Annu Rev Med. 1974;25:333–48. doi: 10.1146/annurev.me.25.020174.002001. [DOI] [PubMed] [Google Scholar]
  83. Shao L, Vawter MP. Shared gene expression alterations in schizophrenia and bipolar disorder. Biol Psychiatry. 2008;64:89–97. doi: 10.1016/j.biopsych.2007.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Shelton RC, et al. Altered expression of genes involved in inflammation and apoptosis in frontal cortex in major depression. Mol Psychiatry. 2010 doi: 10.1038/mp.2010.52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Stefansson H, et al. Association of neuregulin 1 with schizophrenia confirmed in a Scottish population. Am J Hum Genet. 2003;72:83–7. doi: 10.1086/345442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Stefansson H, et al. Neuregulin 1 and susceptibility to schizophrenia. Am J Hum Genet. 2002;71:877–92. doi: 10.1086/342734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Struewing IT, Barnett CD, Tang T, Mao CD. Lithium increases PGC-1alpha expression and mitochondrial biogenesis in primary bovine aortic endothelial cells. FEBS J. 2007;274:2749–65. doi: 10.1111/j.1742-4658.2007.05809.x. [DOI] [PubMed] [Google Scholar]
  88. Subramanian A, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci U S A. 2005;102:15545–50. doi: 10.1073/pnas.0506580102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Sun X, Wang JF, Tseng M, Young LT. Downregulation in components of the mitochondrial electron transport chain in the postmortem frontal cortex of subjects with bipolar disorder. J Psychiatry Neurosci. 2006;31:189–96. [PMC free article] [PubMed] [Google Scholar]
  90. Takeuchi H, et al. White matter structures associated with creativity: evidence from diffusion tensor imaging. Neuroimage. 2010;51:11–8. doi: 10.1016/j.neuroimage.2010.02.035. [DOI] [PubMed] [Google Scholar]
  91. Taveggia C, et al. Type III neuregulin-1 promotes oligodendrocyte myelination. Glia. 2008;56:284–93. doi: 10.1002/glia.20612. [DOI] [PubMed] [Google Scholar]
  92. Tkachev D, et al. Oligodendrocyte dysfunction in schizophrenia and bipolar disorder. Lancet. 2003;362:798–805. doi: 10.1016/S0140-6736(03)14289-4. [DOI] [PubMed] [Google Scholar]
  93. Torroni A, Wallace DC. Mitochondrial DNA variation in human populations and implications for detection of mitochondrial DNA mutations of pathological significance. J Bioenerg Biomembr. 1994;26:261–71. doi: 10.1007/BF00763098. [DOI] [PubMed] [Google Scholar]
  94. Valvassori SS, et al. Effects of mood stabilizers on mitochondrial respiratory chain activity in brain of rats treated with d-amphetamine. J Psychiatr Res. 2010 doi: 10.1016/j.jpsychires.2010.02.009. [DOI] [PubMed] [Google Scholar]
  95. Vawter MP, et al. Gender-specific gene expression in post-mortem human brain: localization to sex chromosomes. Neuropsychopharmacology. 2004;29:373–84. doi: 10.1038/sj.npp.1300337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Vawter MP, et al. Mitochondrial-related gene expression changes are sensitive to agonal-pH state: implications for brain disorders. Mol Psychiatry. 2006;11:615, 663–79. doi: 10.1038/sj.mp.4001830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Wakabayashi T. Megamitochondria formation - physiology and pathology. J Cell Mol Med. 2002;6:497–538. doi: 10.1111/j.1582-4934.2002.tb00452.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet. 2009;10:57–63. doi: 10.1038/nrg2484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Washizuka S, Iwamoto K, Kakiuchi C, Bundo M, Kato T. Expression of mitochondrial complex I subunit gene NDUFV2 in the lymphoblastoid cells derived from patients with bipolar disorder and schizophrenia. Neurosci Res. 2009;63:199–204. doi: 10.1016/j.neures.2008.12.004. [DOI] [PubMed] [Google Scholar]
  100. Washizuka S, et al. Expression of mitochondria-related genes in lymphoblastoid cells from patients with bipolar disorder. Bipolar Disord. 2005;7:146–52. doi: 10.1111/j.1399-5618.2005.00184.x. [DOI] [PubMed] [Google Scholar]
  101. Watson S, Gallagher P, Ritchie JC, Ferrier IN, Young AH. Hypothalamic-pituitary-adrenal axis function in patients with bipolar disorder. Br J Psychiatry. 2004;184:496–502. doi: 10.1192/bjp.184.6.496. [DOI] [PubMed] [Google Scholar]
  102. Zhang B, Kirov S, Snoddy J. WebGestalt: an integrated system for exploring gene sets in various biological contexts. Nucleic Acids Res. 2005;33:W741–8. doi: 10.1093/nar/gki475. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES