Metabolic pathways in the periphery and brain: Contribution to mental disorders?

Metabolic pathways in the periphery and brain: Contribution to mental disorders?

Accepted Manuscript Title: Metabolic pathways in the periphery and brain: contribution to mental disorders? Author: Andrzej Nagalski Kamil Kozinski Ma...

562KB Sizes 0 Downloads 37 Views

Accepted Manuscript Title: Metabolic pathways in the periphery and brain: contribution to mental disorders? Author: Andrzej Nagalski Kamil Kozinski Marta B. Wisniewska PII: DOI: Reference:

S1357-2725(16)30274-6 http://dx.doi.org/doi:10.1016/j.biocel.2016.09.012 BC 4984

To appear in:

The International Journal of Biochemistry & Cell Biology

Received date: Revised date: Accepted date:

16-5-2016 14-9-2016 15-9-2016

Please cite this article as: Nagalski, Andrzej., Kozinski, Kamil., & Wisniewska, Marta B., Metabolic pathways in the periphery and brain: contribution to mental disorders?.International Journal of Biochemistry and Cell Biology http://dx.doi.org/10.1016/j.biocel.2016.09.012 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 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.

Metabolic pathways in the periphery and brain: contribution to mental disorders? Andrzej Nagalski,1, * Kamil Kozinski,1, * Marta B Wisniewska1, *, † 1

Laboratory of Molecular Neurobiology, Centre of New Technologies, University of Warsaw, 02097 Warsaw, Poland * Equal contribution † Corresponding author: [email protected] Graphical abstract

List of 5 keywords: mental disorders, diabetes, genetic risk factors, insulin, inflammation Abbreviations list: ACCORD-MIND - Action to Control Cardiovascular Risk in Diabetes—Memory in Diabetes; BBB blood-brain barrier; BD - bipolar disorder; BDNF - brain-derived neurotrophic factor; DISC1 disrupted in schizophrenia 1; GLUT - glucose transporter; GSK3 - glycogen synthase kinase 3; GWAS - genome-wide association study; HPA - hypothalamic-pituitary-adrenal; IL - interleukin; INSR - insulin receptor; MAPK - mitogen-activated protein kinase; MD - major depression; MRI magnetic resonance imaging; NHGRI-EBI - National Human Genome Research Institute-European 1

Bioinformatics Institute; NPY - neuropeptide Y; SCZ - schizophrenia; SMI - serious mental illness; T2D - type 2 diabetes; SNP - single-nucleotide polymorphism; TCF7L2 - transcription factor 7-like 2; TNFα - tumor necrosis factor α

2

Abstract The association between mental disorders and diabetes has a long history. Recent large-scale, well-controlled epidemiological studies confirmed a link between diabetes and psychiatric illnesses. The scope of this review is to summarize our current understanding of this relationship from a molecular perspective. We first discuss the potential contribution of diabetes-associated metabolic impairments to the etiology of mental conditions. Then, we focus on possible shared molecular risk factors and mechanisms. Simple comorbidity, shared susceptibility loci, and common pathophysiological processes in diabetes and mental illnesses have changed our traditional way of thinking about mental illness. We conclude that schizophrenia and affective disorders are not limited to an imbalance in dopaminergic and serotoninergic neurotransmission in the brain. They are also systemic disorders that can be considered, to some extent, as metabolic disorders.

3

1. INTRODUCTION The high prevalence between schizophrenia (SCZ) and diabetes was first observed by Sir Henry Maudsley in 1879. In the 1920s and 1930s, a series of preliminary studies of small groups of patients revealed higher glucose levels, diabetes-like glucose tolerance curves, and frequent insulin resistance in patients with “dementia precox,” which SCZ was referred to at the time (Kohen, 2004). Also at that time, insulin-shock therapy was introduced by Manfred Sakel to treat dementia precox by inducing coma and convulsions (Sakel, 1994). Soon afterward, clinicians noticed that some patients required more insulin to achieve seizures (Meduna et al., 1942), further corroborating the link between diabetes and SCZ. In recent decades, many potential biological risk factors and mechanisms of diabetes and serious mental illnesses (SMIs) have been proposed. This was accompanied by an expansion of the identification of genetic risk factors based on genome-wide association studies (GWASs) and other large cohort analyses. Combining genetic and biological risk factors between diabetes and SMIs led to tentative evidence that both diseases might share a common pathophysiological nexus. This review examines and discusses the potential factors that are responsible for the observed comorbidity, including the effects of impaired glucose metabolism on brain function, genetic risk factors, and potential pathophysiological mechanisms.

1.1. Serious mental illnesses: schizophrenia, bipolar disorder, and major depression Schizophrenia, bipolar disorder (BD), and major depression (MD) are three psychiatric disorders that are often described under the term SMIs. These disorders are separate disease entities with multifactorial and poorly understood etiology and heterogeneous symptoms (Insel, 2010; Kato, 2008; Kavanagh et al., 2015; Wong and Licinio, 2001). They are diagnosed according to established criteria, but the current biological and medical view is that they lie on a continuum with overlapping phenotypes. Major depression is characterized by low mood, anhedonia, fatigue, psychomotor retardation, ideas of guilt and unworthiness, and, in cases of psychotic depression, hallucinations. All of these symptoms occur in BD and SCZ. In BD, the periods of depression alternate with periods of mania that manifest as heightened energy and mood. Schizophrenia is the most severe disorder of these three. Patients with SCZ suffer from negative and positive symptoms and cognitive impairments. Negative symptoms include anhedonia, 4

flattened affect, and asociality. Positive symptoms include delusions, hallucinations, disorganized thinking and speech, agitation, and catatonia. Lying between these illnesses is schizoaffective disorder, which is a combination of positive symptoms and mood alterations. Serious mental illnesses are generally accepted to be caused by impairments in monoaminergic systems in the brain, and multiple genetic, biological, and social risk factors contribute to the development of these conditions (Green et al., 2015; Kato, 2008; Kavanagh et al., 2015). However, the mechanisms that link these risk factors with etiology and the actual manifestations of SMIs need to be discovered.

1.2. Type 2 diabetes Diabetes is the most common metabolic disorder, occurring in one out of every 11 adults. It is characterized by chronically elevated glucose levels in the blood as a result of deficits in insulin secretion or insulin action (Tripathy and Chavez, 2010). In healthy people, high blood glucose levels after a meal induce the secretion of insulin by β-cells in the pancreas (Del Prato et al., 2002). Glucose is then absorbed from the blood by insulin-sensitive cells (primarily skeletal muscle cells, adipocytes, and liver cells) to be stored as glycogen or fat. This mechanism is disturbed in people with diabetes, ultimately leading to hyperglycemia, which is harmful to many organs and tissues (Tuomi et al., 2014). There are two main types of this disease. Type 2 diabetes (T2D) is the most common form of diabetes, which results from a reduction of the sensitivity of target organs to the actions of insulin (Tuomi et al., 2014). Such insulin resistance is followed by a failure of pancreatic β-cells to produce and secrete sufficient insulin. Similar to SMIs, T2D is a complex polygenic disorder that is triggered by both genetic and environmental risk factors, the pathophysiology of which is still not fully understood.

1.3. Comorbidity of diabetes and serious mental illnesses Looking at the association between diabetes and SMIs, meta-analyses have shown that people who suffer from any of these three SMIs develop T2D at least two-times more often than the general population (Anderson et al., 2001; Annamalai and Tek, 2015; Asghar et al., 2007; Knol et al., 2006; Mezuk et al., 2008; Roy and Lloyd, 2012; Vancampfort et al., 2013). The prevalence of T2D in individuals with SMIs may be even higher. An estimated 70% of cases of T2D in people 5

with SMIs are undiagnosed, contrasting with 25-30% in the general population (Taylor et al., 2005; Voruganti et al., 2007). Although the prevalence of T2D in SMIs is generally accepted, antipsychotics, moodstabilizing medications, and an unhealthy lifestyle are often blamed for this comorbidity, rather than a direct link between these disorders. Numerous studies have addressed the problem of obesity, metabolic syndrome, and diabetes as side effects of psychoactive medications (Holt and Peveler, 2009; Rojo et al., 2015). Differentiating between the effects of antipsychotic medications and the medication-independent association between T2D and SMIs is difficult because virtually all psychiatric patients receive pharmaceutical treatment. Two pieces of evidence, however, corroborate the latter possibility (although they do not necessarily exclude the former). First, the bidirectional relationship between depression and T2D has been consistently reported (Chen et al., 2013; Pan et al., 2012; Rotella and Mannucci, 2013) and was recently supported by a significant genetic correlation (Kan et al., 2016). In SCZ (or BD), it is not possible to observe a bidirectional relationship because these diseases affect mostly young people who are unlikely to have established T2D. Nevertheless, several studies reported lower glucose tolerance in firstepisode and drug-naive SCZ patients compared with matched controls (Enez Darcin et al., 2015; Fernandez-Egea et al., 2009; Garcia-Rizo et al., 2016; Kirkpatrick et al., 2009; Ryan et al., 2003). For example, in a recent study of a sample of 84 SCZ and 98 matched healthy subjects, the average glucose level in a 2-h glucose tolerance test was approximately 30% higher in SCZ patients (GarciaRizo et al., 2016). Second, T2D is more prevalent not only in SCZ patients but also in their families (Foley et al., 2016; van Welie et al., 2013; Yang et al., 2012). To summarize, the association between SMIs and T2D might result from an impairment of glucose metabolism or shared genetic risk factors and pathophysiological mechanisms. This review considers both of these possibilities.

2. ROLE OF GLUCOSE IN PHYSIOLOGICAL AND PATHOLOGICAL BRAIN FUNCTION Numerous metabolic disorders, such as homocysteine metabolism disorders, urea cycle disorders, lipid storage disorders, and leukodystrophies, are rare but important causes of psychiatric disorders in adolescents and adults (Demily and Sedel, 2014). Therefore, changes in metabolism can have serious consequences on the central nervous system. These effects can range from psychosis, depression, and mania in cases of severe disruptions to periodic or subtle behavioral disturbances in cases of mild disruptions. Glucose metabolism is closely integrated 6

with brain physiology, but unclear is whether dysregulated glucose metabolism can exert these kinds of effects.

2.1. Glucose energy metabolism in the brain The brain demands high amounts of energy. Although the brain represents only 2% of the mass of the human body, approximately 20% of the oxygen and 25% of the glucose that are consumed in the body are dedicated to brain function (Belanger et al., 2011). Two main processes contribute to the high energy demands of the brain: (i) maintenance and restoration of ion gradients that are dissipated by signaling processes, such as postsynaptic and action potentials, and (ii) the synthesis, uptake, and recycling of neurotransmitters (Alle et al., 2009; Harris et al., 2012; Mergenthaler et al., 2013). The overall energy budget of the brain is a result of intense coordination between the main cell types in the brain (i.e., neurons, astrocytes, and oligodendrocytes) and epithelial cells of cerebral blood vessels, which form the blood-brain barrier (BBB). The obligatory fuel for the brain is glucose. The brain requires a constant supply of glucose because it can store only a small amount of glycogen in astrocytes (Brown et al., 2005; Sickmann et al., 2009; Wender et al., 2000). Glucose is transported across the BBB through the saturable and insulin-independent glucose transporter 1 (GLUT1), which is expressed on vascular endothelial cells (Bondy et al., 1992; Kobayashi et al., 1996). The uptake of glucose by glial cells is also facilitated by GLUT1, whereas neurons use another insulin-independent transporter, GLUT3, which has higher affinity for glucose and higher transport capacity (Bondy et al., 1992). During maturation in childhood or upon prolonged starvation, brain cells can utilize ketone bodies that are produced in the liver (Lutas and Yellen, 2013). Different cell types in the brain have distinct metabolic profiles, which have been studied particularly for neurons and astrocytes (Belanger et al., 2011; Magistretti and Allaman, 2015) and are beginning to be elucidated for oligodendrocytes (Saab et al., 2013). Neurons sustain a high rate of oxidative metabolism compared with glial cells, which are characterized by a high rate of glycolysis (Harris et al., 2012). A large portion of glucose that enters the glycolytic pathway in astrocytes is released as lactate in the extracellular space and further used by neurons as an energy source. Recent evidence shows that lactate is also released by oligodendrocytes to feed 7

myelinated axons because myelin itself might limit the entry of glucose into the axonal compartment (Fünfschilling et al., 2012; Lee et al., 2012).

2.2. Effect of hyperglycemia and hypoglycemia on brain glucose metabolism The influx of glucose into the brain through the BBB by GLUT1 depends on the blood glucose concentration, which is strictly regulated by endocrine responses and varies between 3.9 and 6.1 mmol/L under physiological conditions (Falkowska et al., 2015). Hyper- and hypoglycemic clamp studies of conscious human subjects found that rising blood glucose levels were followed by a parallel increase in brain glucose (Abi-Saab et al., 2002; Heikkila et al., 2010; van de Ven et al., 2012). Notably, however, the level of glucose in extracellular fluid in the brain was always a few times lower than in blood and lagged approximately 30 min behind changes in plasma (Abi-Saab et al., 2002). The molar content of lactate levels in extracellular fluid in the brain was higher than in plasma and exceeded glucose levels under physiological conditions, reflecting the local generation of lactate from glucose by astrocytes (Belanger et al., 2011). Such a lag time, the local availability of lactate, and vascular barriers might play a protective role in limiting brain injury during severe hypo- or hyperglycemia. When blood glucose levels were acutely elevated to 11.5 mmol/L, the extracellular level in the brain was ~2.6 mmol/L (Abi-Saab et al., 2002). This indicates that cells in the brain might not be exposed to very high glucose levels during hyperglycemia. This contrasts with peripheral nerves, in which hyperglycemia is associated with marked increases in intracellular glucose (Stewart et al., 1966), and persistent hyperglycemia can lead to the development of diabetic neuropathy (Tomlinson and Gardiner, 2008). In studies of hypoglycemia, blood and brain levels dropped to 50% of baseline, while plasma lactate levels increased (AbiSaab et al., 2002). Many studies on diabetic rats reported a decrease in the transport of glucose into the brain during chronic hyperglycemia (Hou et al., 2007; Puchowicz et al., 2004), which is a possible adaptation to high glucose levels. If this also occurs in humans, then this could explain why diabetic patients develop transient symptoms of cerebral hypoglycemia after the rapid normalization of blood glucose levels (Bathla et al., 2014). The basis of this adaptation is unknown. A decrease in glucose transporters on epithelial cells in the BBB could also be an explanation, but no consensus has yet been reached in the literature (Hou et al., 2007; Shah and Parent, 2003). 8

Several studies of hypoglycemia in animals have reported an increase in GLUT expression in both the BBB and neurons (Kumagai et al., 1995; Simpson et al., 1999; Uehara et al., 1997). However, whether similar effects occur in humans is controversial (Boyle et al., 1994; Pelligrino et al., 1990; Segel et al., 2001). Further potential adaptation to lower glucose levels can involve the utilization of alternative energy substrates (e.g., ketone bodies or lactate) by increasing the level of monocarboxylate transporters (Herzog et al., 2013; Leino et al., 2001). In summary, the brain apparently attempts to adjust to recurrent and persistent hyper- or hypoglycemia in the long term.

2.3. Effects of acute hyperglycemia and hypoglycemia on cognition and mental states During ongoing hyperglycemia, diabetic patients often report acute cognitive and mental dysfunction, but only a few studies have confirmed such an association. In a study of 20 T2D patients (60 years old), impairments in learning and memory were observed during induced acute hyperglycemia (Sommerfield et al., 2004). In another study that simultaneously assessed fasting glucose levels and cognitive performance in diabetes patients approximately 50 times over 1 month, a positive correlation was found between acute hyperglycemia and mild cognitive impairment (Cox et al., 2005). The possible link between acute hyperglycemia and mental symptoms has been scarcely documented. Only a few studies have addressed this issue, and unequivocal conclusions have not been made. In one study, T2D patients (n = 20) experienced a lack of energy, sadness, and anxiety during acute hyperglycemia (glucose level = 16.5 mmol/L; (Sommerfield et al., 2004). In another small study (n = 15), patients reported an increase in well-being and less anger during mild hyperglycemia (10.5 mmol/L; (Pais et al., 2007). Hypoglycemia occurs in approximately 10-30% of T2D patients who are treated with insulin (McCrimmon and Sherwin, 2010). Hypoglycemia is moderate when blood glucose levels fall to 3.5-2.5 mmol/L and severe when blood glucose levels fall below 2.5 mmol/L (Languren et al., 2013). When blood glucose levels fall, compensatory neuroendocrine responses are initiated. If these defenses fail to abort the hypoglycemic episode, then a more intense sympathoadrenal response can cause so-called neuroglycopenic symptoms, including confusion, dizziness, difficulty speaking, blurred vision, irritability, fainting, seizures, or coma (Cryer, 2007; de Galan et al., 2006). 9

This is unsurprising when considering that neuronal function is energy demanding, and neurons depend on a regular supply of glucose. Severe hypoglycemia can lead to coma, characterized by unconsciousness and the cessation of electrical brain activity (isoelectric period), which might induce permanent neuronal damage, especially in the cortex, hippocampus, putamen, and caudate nucleus (Kalimo and Olsson, 1980; Ma et al., 2009; Shaefer et al., 2016). Nevertheless, brain damage is very rare, and recovery from hypoglycemia is usually complete after plasma glucose concentrations increase (1996).

2.4. Long-term effects of hyper- and hypoglycemia on cognition and brain integrity The effects of recurrent hyperglycemia on cognition are more pronounced than in the case of acute elevations of blood sugar. Long-term glycemic control (assessed by the glycohemoglobin test) in T2D patients is inversely related to performance on tasks that assess learning, reasoning, and memory. For example, a trial of the Action to Control Cardiovascular Risk in Diabetes– Memory in Diabetes (ACCORD-MIND) that included approximately 3,000 T2D patients found an association between cognitive decline and poor glycemic control (Cukierman-Yaffe et al., 2009). Importantly, however, the differences in cognitive function were mild to moderate. This cognitive decline was likely associated with hyperglycemia-related complications, such as brain atrophy associated with the presence of microvascular complications. Hyperglycemia is dangerous for cells because it causes oxidative stress and subsequent tissue deterioration (Tomlinson and Gardiner, 2008). One reason for these sequelae is that excess glucose is converted to sorbitol in the polyol pathway by utilizing NADPH as a reducer. Consequently, NADPH is depleted, similar to the key antioxidant glutathione, which also requires NADPH to be synthesized. This leads to the accumulation of reactive oxygen species, mitochondrial damage, and cell death. In parallel, sorbitol spontaneously interacts with proteins to produce advanced glycation end products, excess levels of which are harmful to cells in several ways, including cross-linking molecules and triggering inflammation (Ott et al., 2014). An oversupply of glucose is also metabolized in the hexosamine pathway, leading to the generation of N-acetylglucosamine. The subsequent glycosylation of proteins results in the dysregulation of signal transduction and transcription. Finally, very high extracellular glucose (> 30 mmol/L) increases osmolality, which pulls water out of cells and leads to dehydration. Because endothelial vascular cells are particularly exposed to hyperglycemia, diabetes-associated recurrent 10

hyperglycemia progressively compromises the vascular endothelium (Vinik and Flemmer, 2002). Failures in the micro- and macro-vasculature that are observed in the brain in T2D (Li et al., 2015) can lead to neurodegeneration in the brain (Tennant and Brown, 2013). This is probably the reason why different forms of dementia (e.g., vascular dementia, Alzheimer’s disease, and Parkinson’s disease) are more frequent in diabetic patients (Cheng et al., 2012; Cukierman et al., 2005; Gudala et al., 2013; Smolina et al., 2015; Wang et al., 2014). Human autopsy studies of patients who died from hypoglycemia showed multifocal and diffuse necrosis in the brain (Kodl and Seaquist, 2008). Unclear is whether recurrent hypoglycemia can cause any long-term deterioration of brain structures and permanent dysfunction. This issue has been addressed in only a few studies. The long-term consequences of recurrent hypoglycemia were recently assessed in the ACCORD-MIND magnetic resonance imaging (MRI) trial. This study included 500 T2D patients who were followed-up for 3 years (Zhang et al., 2014). Patients who experienced severe hypoglycemia during this follow-up period did not show more brain atrophy or white matter impairments than the other patients.

3. COMMON BIOLOGICAL RISK FACTORS UNDERLYING THE COMORBIDITY BETWEEN SERIOUS MENTAL ILLNESSES AND TYPE 2 DIABETES Glucose is essential to brain function, and many compensatory effects prevent or correct changes in physiological plasma glucose concentrations. The effects of hypo- or hyperglycemia on human cognition, discussed in the previous section, are quite well documented but indicate a lack of correlations, aside from those that are associated with vascular complications after chronic hyperglycemia. Therefore, a more plausible explanation for the comorbidity between T2D and SMIs are shared risk factors that independently contribute to these disorders.

3.1. Genetic risk factors Family history is an important risk factor for developing SMIs and diabetes, prompting scientists to search for genetic factors that are associated with these disorders. The possibility that T2D and SMIs may have a common genetic etiology was reinforced by GWASs. According to a widely accepted model, hundreds of common alleles confer a risk for diseases with a hereditary component, but each of these alleles alone has little effect (odds ratio > 1.5; (Ku et al., 2010). The 11

last decade has seen an explosion of GWASs, in which thousands of single-nucleotide polymorphisms (SNPs) are analyzed at the same time. The number of disease-associated genes is quickly growing. Considering that SCZ and T2D share some genetic vulnerability, an analysis of candidate genes for these disorders may shed light on possible common pathophysiology. The intersection of 268 SCZ susceptibility genes with 338 T2D susceptibility genes that were listed in the Genetic Association Database (http://geneticassociationdb.nih.gov) revealed a total of 37 common genes (Lin and Shuldiner, 2010) (Fig. 1). A similar procedure was applied to 200 SCZ genes that were listed in the Genetic Association Database and Catalog of Published Genome-Wide Association Studies and 196 T2D genes listed in the above database plus Type 2 Diabetes Association Database (Liu et al., 2013). In this study, 14 genes were found to intersect. Two functional categories could be identified within these groups: inflammation-associated genes (IL1B, IL6, IL10, HLA-A, HLA-DQA1, HLA-DQB1, HLADRB1, HSPA1B, TNF, CTLA4, APOE, and PTGS1) and genes that are involved in protecting cells from oxidative stress (PON1, SOD2, MTHFR, UCP2, and GSTM1). A network analysis based on interactions between protein products of all SCZ and T2D susceptibility genes and their interacting partners identified several hub proteins that are involved in calcium signaling, insulin signaling, adipokine signaling, and AKT signaling (Liu et al., 2013). Very recently, a genetic overlap between MD and T2D was identified based on data from DIAbetes Genetics Replication And Meta-analysis (DIAGRAM) and Psychiatric Genomics Consortium meta-GWAS statistics (Ji et al., 2016). A functional analysis of the shared 496 SNPs revealed enrichment within pathways that are involved in immune responses, cell signaling (mitogen-activated protein kinase [MAPK], Wnt signaling), and lipid metabolism. To our knowledge, no analysis has been conducted concerning BD and T2D shared SNPs or genes. Currently, the National Human Genome Research Institute–European Bioinformatics Institute (NHGRI-EBI) catalog of published GWASs lists 402 SCZ susceptibility genes, 231 BD susceptibility genes, 257 depression susceptibility genes, 319 genes associated with SMIs (with no distinction between them), and 890 genes associated with T2D (calculated by the authors; https://www.ebi.ac.uk/gwas; accessed February 24, 2016). For the purpose of this review, we performed an analysis of the genes that are associated with mental disorders and T2D from the aforementioned NHGRI-EBI (Fig. 1). We found 79 candidate genes that are shared by T2D and any 12

of the SMIs. In this group, 26, 19, and 18 genes were specifically associated with T2D and SCZ, T2D and BD, and T2D and MD, respectively. Functional analysis of the group of 79 genes (DAVID database, https://david.ncifcrf.gov) revealed several clusters of common candidate risk genes. One cluster included genes that are involved in regulating multiple processes by intracellular signal transduction, such as genes that encode receptor regulators and ligand-binding proteins (ERRγ, NRG3, CDH13, FAF1, and APOB), components of signaling cascades (PRKCQ), and transcription factors and their regulators (NFKB1, TCF7L2, HNF1, ZMIZ1, and PPARGC1A). Some of these proteins have well-established roles in inflammation (NFKB1 and HLA-DQA1) or are involved in regulating the energy metabolism or action of insulin (HNF1, TCF7L2, CDH13, APOB, and SORBS1). Some genes that were identified in the intersection encode proteins that might have different functions in the nervous system and in peripheral organs (e.g., synapsin 2, transcription factor 7-like 2 [TCF7L2], and disrupted in schizophrenia 1 [DISC1]). Synapsin 2 is a synaptic protein that is involved in regulating neurotransmitter release from vesicles (Sudhof, 2004) and also expressed in β-cells, but its role in these cells is still unclear (Wendt et al., 2012). Below we present more details of the two other examples: the transcription factor TCF7L2 and multifunctional protein DISC1.

3.1.1. TCF7L2 The TCF7L2 gene encodes a transcription factor and effector of Wnt signaling. This signaling pathway is implicated in the pathophysiology and treatment of depression (Bordonaro, 2009; Niciu et al., 2013; Voleti and Duman, 2012) and was enriched among shared SNPs associated with both T2D and MD mentioned previously (Ji et al., 2016). TCF7L2 is the strongest genetic risk factor for T2D (Cauchi et al., 2007; Grant et al., 2006; Shen et al., 2015; Sladek et al., 2007; Welters and Kulkarni, 2008). Evidence from two independent cohorts showed an association between TCF7L2 and BD in obese patients (Cuellar-Barboza et al., 2016; Winham et al., 2014). Variants of TCF7L2 were linked with SCZ, but the statistical power was low in these studies (Alkelai et al., 2012; BenDavid et al., 2010; Hansen et al., 2011). Much research has sought to understand the mechanisms that underlie the metabolic function of TCF7L2 in organs that are involved in the pathogenesis of T2D, including the pancreas and liver. Accumulating evidence suggests that the activity of TCF7L2 is critical for the proliferation and survival of β-cells, modulation of insulin secretion (Liu and 13

Habener, 2010), and synthesis and processing of proinsulin and insulin (Zhou et al., 2014). The importance of TCF7L2 was also demonstrated in genetically modified mice. The disruption of Tcf7l2 in β-cells induced the development of glucose intolerance, β-cell dysfunction, a reduction of hepatic glucose production during fasting, and an improvement in glucose homeostasis when maintained on a high-fat diet (Boj et al., 2012; da Silva Xavier et al., 2012; Mitchell et al., 2015). Advances in our understanding of the role of TCF7L2 in organs that are involved in maintaining metabolic homeostasis have not been accompanied by similar analyses in the brain, where TCF7L2 is expressed at high levels in the medial habenula, thalamus, and inferior colliculus (i.e., areas that are implicated in sensory perception, attention, awareness, and decision-making; (Nagalski et al., 2013). Neuroimaging studies showed that these brain structures were affected in SCZ patients (Haijma et al., 2013; Kang et al., 2008; Shepard et al., 2006). The precise role of TCF7L2 in the brain has been enigmatic. Research from our group revealed the enrichment of TCF7L2 binding sites in regulatory regions of genes that are associated with neuronal excitability, such as genes that encode voltage-gated ion channels and neurotransmitter receptors, and transcription factors that are involved in the development of the thalamus (Nagalski et al., 2015; Wisniewska et al., 2010; Wisniewska et al., 2012). This suggests that TCF7L2 is an important regulator of neuronal excitability in the thalamus. The possible involvement of this (potential) function of TCF7L2 in pathophysiological mechanisms in SMIs needs further investigations.

3.1.2. DISC1 Another example of a SCZ-associated protein that could play separate roles in the brain and in the regulation of metabolism in the periphery is DISC1 (Soares et al., 2011). Although large cohort studies did not find an association between DISC1 and SCZ (Mathieson et al., 2012), mutations within this gene were found in members of large Scottish and American families who were affected by different SMIs (Farrell et al., 2015; Sachs et al., 2005). DISC1 acts as a scaffold protein. One of its interacting partners is glycogen synthase kinase 3/ (GSK3α/β), a mediator of the aforementioned Wnt pathway (Mao et al., 2009). Studies in genetically modified mouse models revealed multiple roles for DISC1 in the brain in neural precursor proliferation, neuronal migration, axonal guidance, dendritic growth, synaptic plasticity, and mitochondrial function and trafficking (Thomson et al., 2013). 14

Unexpectedly, a recent study showed that DISC1 is involved in regulating β-cell physiology in the pancreas (Jurczyk et al., 2016). In this study, the loss of DISC1 function in transgenic mice resulted in a decrease in β-cell proliferation, an increase in β-cell apoptosis, lower insulin secretion, and glucose intolerance. These effects were mediated by the inappropriate activation of GSK3, which was previously shown to play an important role in pancreatic β-cell function and survival (Mussmann et al., 2007).

3.2. Potential shared pathophysiology of serious mental illnesses and type 2 diabetes Accumulating evidence, including the aforementioned functional analyses of susceptibility genes, suggests that shared molecular and cellular mechanisms contribute to the development of SMIs and T2D. The following section focuses on insulin signaling, inflammation, mitochondrial dysfunction, oxidative stress, and neuroendocrine systems. Cross talk among these systems may be important in understanding the comorbidity between T2D and SMIs.

3.2.1. Insulin signaling in the brain Insulin was long considered a peripheral hormone that regulates energy metabolism in the liver, muscles, and fat. However, insulin receptors (INSRs) are widely distributed in the brain and expressed on both neurons and glial cells (Kleinridders et al., 2014; Werner and LeRoith, 2014). Furthermore, peripheral insulin crosses the BBB (Banks, 2004). There is a consensus that insulin acts in the brain, particularly in the hypothalamus, to regulate carbohydrate and lipid metabolism in the periphery and feeding behavior (Heni et al., 2015; Sanchez-Lasheras et al., 2010) (Fig. 2). For example, mice with neuron-specific deletion of the Insr gene (NIRKO mice) exhibited an imbalance in energy metabolism in the periphery and insulin resistance (Bruning et al., 2000), and insulin administration into the third ventricle in the brain in rodents suppressed glucose production in the body (Obici et al., 2002). Although the uptake of glucose by brain cells is generally insensitive to insulin, insulin likely regulates energy metabolism in a scattered population of neurons that express the insulinsensitive transporter GLUT4. A rodent model showed that GLUT4 colocalizes with INSRs in brain regions associated with motor control, the hippocampus, and the hypothalamus (El Messari et 15

al., 1998; Ren et al., 2014; Vannucci et al., 1998), forming a functional insulin-sensitive signaling pathway of glucose uptake. The role of this pathway in neurons is enigmatic, but it appears to regulate metabolism in these cells. For example, the local delivery of insulin in the rat hippocampus enhanced local glycolytic metabolism (McNay et al., 2010). Interestingly, this effect was attenuated in animals in which T2D was induced by a high-fat diet (McNay et al., 2010), suggesting that insulin resistance can also develop in the brain. A human study that used positron emission tomography and functional MRI found that insulin enhanced glucose uptake in regions with a high density of INSRs (Bingham et al., 2002; Kullmann et al., 2015). In neurons that express INSRs, similar to INSR-positive cells in the periphery, the activation of INSRs induces phosphoinositide 3-kinase (PI3K)/AKT signaling via INSR substrates. This cascade targets multiple downstream pathways (e.g., mTOR, ERK/MAPK, GSK3α/β or FOXO pathways), which in the case of neurons are implicated in the regulation of cell survival and synaptic plasticity (Adams and Sweatt, 2002; Kim et al., 2012; Ren et al., 2013; Salcedo-Tello et al., 2011; Stoica et al., 2011). Therefore, insulin in the brain can exert general neuroprotective and trophic effects and improve learning and memory. Such an assertion was corroborated by animal studies (Fanne et al., 2011; Grillo et al., 2015; Sanderson et al., 2009) and human studies with intranasally administered insulin (Ott et al., 2012). Insulin was also proposed to target dopaminergic systems that regulate reward-motivated behavior and play a crucial role in SMIs. In a mouse model, insulin had a significant excitatory effect in a major subpopulation of dopaminergic neurons in the basal ganglia, and this effect was abolished by knockout of the Insr gene (Konner et al., 2011). Furthermore, NIRKO mice exhibited depressive-like behavior (Haider et al., 2013; Kleinridders et al., 2015; Sharma et al., 2010). Unknown is the way in which insulin mechanistically contributes to the regulation of dopaminergic systems. Nevertheless, insulin treatment in streptozotocin-induced diabetic mice showed that insulin decreased monoamine oxidase (i.e., the enzyme that is responsible for the degradation of dopamine) activity in the brain (Gupta et al., 2014). Converging evidence indicates that insulin, apart from regulating peripheral glucose metabolism, affects a wide range of normal brain functions, such as memory formation and reward circuit, and these functions can be altered in T2D, potentially contributing to cognitive or mental impairments. 16

3.2.2. Inflammation Many inflammation-associated genes have been identified as genetic risk factors for the development of SMIs and T2D. Obesity, insulin resistance, and T2D are closely associated with inflammation (Hotamisligil, 2006; Pradhan et al., 2001). An imbalance in the cytokine/chemokine network is also involved in SMIs (Landek-Salgado et al., 2016). Tumor necrosis factor α (TNFα), interleukin 1β (IL-1β), and IL-6 levels were shown to increase in murine models of obesity and T2D (McArdle et al., 2013). TNFα is overexpressed in adipose and muscle tissue in obese humans, and exogenous TNFα administration leads to insulin resistance (Hotamisligil et al., 1995; Kern et al., 1995; Saghizadeh et al., 1996). The secretion of TNFα, IL-1β, and IL-6 draws immune cells to adipose tissue and initiates the inflammatory response. The activation of TNFα signaling pathways impairs insulin signaling at the level of INSR substrates through inhibitory phosphorylation (Lorenzo et al., 2008). The disruption of adipose tissue function causes further defects in hepatic and skeletal muscle glucose homeostasis, resulting in systemic insulin resistance and leading to the development of T2D (McArdle et al., 2013). It is generally accepted that maternal infection during pregnancy and multiple infections in childhood increase the risk of SCZ (Brown, 2011), and the reason for it can be an effect of inflammation on brain development. Nevertheless, inflammatory cytokines are elevated also in adult SMI patients. A recent meta-analysis found that serum levels of IL-1β, IL-6, TNFα, and serum IL-2 receptor were increased in medication-naive patients with first-episode psychosis and in BD patients (Bauer et al., 2014; Upthegrove et al., 2014). A meta-analysis of MD confirmed higher mean levels of IL-6 and C-reactive protein but not TNFα or IL-1β (Haapakoski et al., 2015). Evidence indicates that proinflammatory cytokines have a negative effect on brain-derived neurotrophic factor (BDNF) expression levels in patients with affective disorders (Grassi-Oliveira et al., 2008; Hayley et al., 2005; Kauer-Sant'Anna et al., 2007; Mondelli et al., 2011). Polymorphisms within the BDNF gene were associated with SCZ (Bonaccorso et al., 2015; Li et al., 2012; Li et al., 2013; Watanabe et al., 2013), mood disorders (Bonaccorso et al., 2015), and obesity (Beckers et al., 2008; Ernst et al., 2012; Gray et al., 2006; Hotta et al., 2009), although large cohort studies failed to confirm this association. Moreover, mice with deficient BDNF expression exhibited behavioral abnormalities, such as aggression and hyperactivity (Coppola and Tessarollo, 17

2004), and also obesity (Sha et al., 2007). Therefore, inflammation-associated alterations in BDNF expression could be another mechanism of the development of SMIs and diabetes. A GWAS that included nearly 40,000 SCZ and 120,000 control subjects corroborated a strong association between SCZ and genes that are implicated in the immune response, particularly the major histocompatibility complex (MHC) locus (Schizophrenia Working Group of the Psychiatric Genomics, 2014). A study that used a mouse model found that variations in the complement component 4 (C4) gene within this locus affected C4 expression in the brain, and C4 in synapses was critically involved in synapse elimination by microglia during postnatal development (Schafer et al., 2012; Sekar et al., 2016). This mechanism may link immune response-associated alterations with synaptic pruning in SCZ (Boksa, 2012). There is strong genetic and pathophysiological evidence of a role for the inflammatory response and MHC complex in the development of SMIs and TD2. These associations can contribute to the comorbidity between these disorders.

3.2.3. Mitochondria The implication of mitochondrial dysfunction in the development of insulin resistance is still debated (Montgomery and Turner, 2015). Although reductions of mitochondrial content, alterations of the expression of subsets of mitochondrial genes, and decreases in the expression of mitochondrial oxidative proteins have been well described in the pathogenesis of T2D, various independent studies have failed to show such a correlation, as reviewed elsewhere (Kaufman et al., 2015; Martin and McGee, 2014; Montgomery and Turner, 2015). Several lines of evidence suggest that mitochondrial dysfunction is an important component in the pathology of SMIs, especially BD and SCZ (Manji et al., 2012). Postmortem examinations of brain tissue ultrastructure and metabolomic studies have provided evidence of mitochondrial dysfunction in the brain in SCZ and BD patients (Clay et al., 2011). Perturbations in mitochondrial morphology and a decrease in the number of mitochondria were observed in different brain regions in qualitative electron microscopy studies of SCZ brains (Kolomeets and Uranova, 2010; Kung and Roberts, 1999; Roberts et al., 2015; Uranova et al., 2011; Vikhreva et al., 2016). Decreases in the activity of the mitochondrial respiratory chain complexes were also reported in 18

different brain areas in SCZ and BD patients (Andreazza et al., 2010; Maurer et al., 2001). Largescale transcriptomic and proteomic studies further supported the mitochondrial hypothesis of SMIs by showing decreases in the gene/protein expression of Krebs cycle and/or respiratory chain enzymes in SCZ and BD brains (Altar et al., 2005; Gottschalk et al., 2015; Iwamoto et al., 2005; Konradi et al., 2004; Middleton et al., 2002; Pennington et al., 2008; Prabakaran et al., 2004), mostly in the prefrontal cortex. However, these results were only partially consistent with each other. Although some of the studies reported decreases in both SCZ and BD samples, others reported a decrease only in SCZ samples or only in BD samples. Moreover, one study reported that the expression of the aforementioned genes actually increased in BD samples (Gottschalk et al., 2015). Another question is whether mitochondrial dysfunction is a cause or consequence of T2D and SMIs. The former is corroborated by numerous data that showed that mitochondrial genes are impacted by copy number variations, polymorphisms, and mutational events in SCZ, BD, and MD (Bi et al., 2016; Purcell et al., 2014; Rollins et al., 2009; Sequeira et al., 2012; Sequeira et al., 2015). Variations in mitochondrial DNA and mitochondria -associated genes were also reported to be associated with the risk of developing diabetes (Cho et al., 2007; Kwak and Park, 2016). Another line of evidence comes from case studies that reported the prevalence of various psychiatric symptoms in patients with recognized mitochondrial diseases (Fattal et al., 2006; Inczedy-Farkas et al., 2012). However, animal studies, in which the function of mitochondria was compromised, did not confirm the causative involvement of mitochondria in the development of insulin resistance (Montgomery and Turner, 2015). With the cumulative evidence of alterations in mitochondrial function in both SMIs and T2D, there is limited evidence that mitochondrial dysfunction is an etiological factor in these disorders.

3.2.4. Oxidative stress Inflammation and mitochondrial dysfunction are interconnected with oxidative stress, which occurs when antioxidant defense mechanisms fail to counterbalance reactive oxygen species. Recent meta-analyses revealed that markers of oxidative stress increase in SCZ (Emiliani et al., 2014; Flatow et al., 2013), BD (Brown et al., 2014), and MD (Black et al., 2015). Imbalances in oxidative and antioxidant defense systems have also been associated with T2D (Maiese, 2015). 19

An interesting link between genetic associations and oxidative stress in these disorders was provided by methylenetetrahydrofolate reductase (MTHFR), an enzyme that contributes to homocysteine metabolism. MTHFR gene was associated both with T2D and SCZ. Elevated levels of homocysteine may lead to an increase in the generation of superoxide, which is further amplified by homocysteine-dependent alterations in the function of cellular antioxidant enzymes (Weiss et al., 2003). Many other genes that are implicated in antioxidant defense systems have been linked to T2D (Banerjee and Vats, 2014), SCZ (Chowdari et al., 2011), and BD (Fullerton et al., 2010). The correlation between oxidative stress, T2D, and SMIs appears to be well established.

3.2.5. Neuroendocrine system Hypothalamic-pituitary-adrenal (HPA) axis is a major neuroendocrine system that control reaction to stress, and regulates, among others, processes of energy storage and expenditures, immune functions, mood and emotions. It is not surprising that dysfunctions of HPA axis is suggested as a pathological factor in SMIs and T2D (Chan et al., 2003; Keller et al., 2006; Lamers et al., 2013; Pariante and Lightman, 2008; Steen et al., 2014) The elevated systemic cortisol metabolism is present in both SCZ (Steen et al., 2014; Steen et al., 2011) and BP (Cervantes et al., 2001; Daban et al., 2005; Watson et al., 2004). The hyperactivity of the HPA axis and hypercotisolemia is one of the most consistent biological findings in MD (Pariante and Lightman, 2008; Stetler and Miller, 2011). Also, there is growing evidence that in humans prenatal stress influences HPA regulation and is associated with higher prevalence of mental illnesses including SCZ and depression (Kim et al., 2015). Insulin resistant and obese individuals exhibit elevated cortisol that is produced by the adrenal cortex (Chiodini et al., 2007). Hypercortisolaemia may increase adiposity; therefore contribute to metabolic syndrome, insulin resistance, and T2D (Adam and Epel, 2007; Gibson, 2012). This effect is mediated by an influence of cortisol on the secretion of leptin (secreted by adipose cells), insulin and brain neuropeptide Y (NPY), which are both abovementioned risk factors for T2D and SMIs. Insulin and NPY, which both act on hypothalamic neurons, are involved in the regulation of appetite. They also influence neurons in the dopaminergic reward circuit (described for insulin hereinbefore). It was recently proposed that another neuroendocrine system, the dopamine-prolactin pathway contribute to the association between T2D and SCZ (Gragnoli et al., 2016). Dopamine 20

that is released to the pituitary gland acts as a prolactin-inhibitory factor. The levels of both dopamine and prolactin appear to be dysregulated in SCZ (Garcia-Rizo et al., 2012). Prolactin is essential for pancreatic development and insulin secretion capability (Auffret et al., 2013; Freemark et al., 2002; Huang et al., 2009; Terra et al., 2011). Dysregulation of the dopamineprolactin pathway may impair glucose homeostasis and thus contribute to the higher incidence of T2D in SCZ patients. However, this hypothesis needs further investigation because dopamine and prolactin levels were found to both increase and decrease in SCZ patients (Albayrak et al., 2014; Garcia-Rizo et al., 2012; Howes and Kapur, 2009; Jose et al., 2015). Additionally, NPY is of interest to the hypothesized correlation of the dopamine–prolactin pathway with the SMI comorbidity, as NPY appears to amplify the inhibitory action of dopamine on prolactin secretion (Wang et al., 1996). Accumulating evidence suggesting a link between dysregulation of HPA axis and dopamineprolactin pathways and the risk of developing SMIs and T2D. NPY might be an important player in this association.

4.

CONCLUSIONS The comorbidity between T2D and SMIs is unquestionable. Less certain is whether biological

links between these disorders per se are contributing factors. The long-term consequences of hyperglycemia, which is the principal diabetic pathology, appear to contribute to cognitive decline and neurodegeneration, but is rather a result of vascular complications. Instead, accumulating evidence suggests that the pathogenesis of diabetes and SMIs includes shared molecular and cellular mechanisms. Candidate risk genes for mental and metabolic illnesses that have been revealed by recent large-scale GWASs, together with neuroimaging and animal studies, can help identify these mechanisms. The concept of systemic disorders in SMIs has been proposed in the literature on BD, SCZ, and MD (Berk et al., 2013; Garcia-Rizo et al., 2015; Kirkpatrick and Miller, 2013), mostly based on abnormal neuroendocrine status. Systemic inflammation and oxidative stress appear to be other possible and well-documented mediators of SMIs. There is also an emerging role for insulin in the pathophysiology of SMIs. We may conclude that SMIs could be at least partially considered metabolic and systemic disorders. We believe that the development of future therapeutic approaches for the treatment of psychiatric illnesses will rely on identifying 21

their different molecular and cellular endophenotypes, such as those that are associated with T2D, to develop novel personalized treatments that are more safe and effective than those that are currently used. ACKNOWLEDGEMENTS This work was supported by Polish National Science Centre grant 2015/19/B/NZ4/03571.

BIBLIOGRAPHY 1996. Effects of intensive diabetes therapy on neuropsychological function in adults in the Diabetes Control and Complications Trial. Ann Intern Med 124(4), 379-388. Abi-Saab, W.M., Maggs, D.G., Jones, T., Jacob, R., Srihari, V., Thompson, J., Kerr, D., Leone, P., Krystal, J.H., Spencer, D.D., During, M.J., Sherwin, R.S., 2002. Striking differences in glucose and lactate levels between brain extracellular fluid and plasma in conscious human subjects: effects of hyperglycemia and hypoglycemia. Journal of cerebral blood flow and metabolism : official journal of the International Society of Cerebral Blood Flow and Metabolism 22(3), 271-279. Adam, T.C., Epel, E.S., 2007. Stress, eating and the reward system. Physiol Behav 91(4), 449-458. Adams, J.P., Sweatt, J.D., 2002. Molecular psychology: roles for the ERK MAP kinase cascade in memory. Annu Rev Pharmacol Toxicol 42, 135-163. Albayrak, Y., Beyazyuz, M., Beyazyuz, E., Kuloglu, M., 2014. Increased serum prolactin levels in drug-naive first-episode male patients with schizophrenia. Nord J Psychiatry 68(5), 341-346. Alkelai, A., Greenbaum, L., Lupoli, S., Kohn, Y., Sarner-Kanyas, K., Ben-Asher, E., Lancet, D., Macciardi, F., Lerer, B., 2012. Association of the type 2 diabetes mellitus susceptibility gene, TCF7L2, with schizophrenia in an Arab-Israeli family sample. PloS one 7(1), e29228. Alle, H., Roth, A., Geiger, J.R., 2009. Energy-efficient action potentials in hippocampal mossy fibers. Science 325(5946), 1405-1408. Altar, C.A., Jurata, L.W., Charles, V., Lemire, A., Liu, P., Bukhman, Y., Young, T.A., Bullard, J., Yokoe, H., Webster, M.J., Knable, M.B., Brockman, J.A., 2005. Deficient hippocampal neuron expression of proteasome, ubiquitin, and mitochondrial genes in multiple schizophrenia cohorts. Biol Psychiatry 58(2), 85-96. Anderson, R.J., Freedland, K.E., Clouse, R.E., Lustman, P.J., 2001. The prevalence of comorbid depression in adults with diabetes: a meta-analysis. Diabetes care 24(6), 1069-1078. Andreazza, A.C., Shao, L., Wang, J.F., Young, L.T., 2010. Mitochondrial complex I activity and oxidative damage to mitochondrial proteins in the prefrontal cortex of patients with bipolar disorder. Archives of general psychiatry 67(4), 360-368. Annamalai, A., Tek, C., 2015. An overview of diabetes management in schizophrenia patients: office based strategies for primary care practitioners and endocrinologists. International journal of endocrinology 2015, 969182. Asghar, S., Hussain, A., Ali, S.M., Khan, A.K., Magnusson, A., 2007. Prevalence of depression and diabetes: a population-based study from rural Bangladesh. Diabetic medicine : a journal of the British Diabetic Association 24(8), 872-877. 22

Auffret, J., Freemark, M., Carre, N., Mathieu, Y., Tourrel-Cuzin, C., Lombes, M., Movassat, J., Binart, N., 2013. Defective prolactin signaling impairs pancreatic beta-cell development during the perinatal period. Am J Physiol Endocrinol Metab 305(10), E1309-1318. Banerjee, M., Vats, P., 2014. Reactive metabolites and antioxidant gene polymorphisms in type 2 diabetes mellitus. Indian J Hum Genet 20(1), 10-19. Banks, W.A., 2004. The source of cerebral insulin. Eur J Pharmacol 490(1-3), 5-12. Bathla, G., Policeni, B., Agarwal, A., 2014. Neuroimaging in patients with abnormal blood glucose levels. AJNR Am J Neuroradiol 35(5), 833-840. Bauer, I.E., Pascoe, M.C., Wollenhaupt-Aguiar, B., Kapczinski, F., Soares, J.C., 2014. Inflammatory mediators of cognitive impairment in bipolar disorder. J Psychiatr Res 56, 18-27. Beckers, S., Peeters, A., Zegers, D., Mertens, I., Van Gaal, L., Van Hul, W., 2008. Association of the BDNF Val66Met variation with obesity in women. Mol Genet Metab 95(1-2), 110-112. Belanger, M., Allaman, I., Magistretti, P.J., 2011. Brain energy metabolism: focus on astrocyteneuron metabolic cooperation. Cell Metab 14(6), 724-738. Ben-David, E., Shifman, S., International Schizophrenia, C., 2010. Further investigation of the association between rs7341475 and rs17746501 and schizophrenia. American journal of medical genetics. Part B, Neuropsychiatric genetics : the official publication of the International Society of Psychiatric Genetics 153B(6), 1244-1247. Berk, M., Williams, L.J., Jacka, F.N., O'Neil, A., Pasco, J.A., Moylan, S., Allen, N.B., Stuart, A.L., Hayley, A.C., Byrne, M.L., Maes, M., 2013. So depression is an inflammatory disease, but where does the inflammation come from? BMC Med 11, 200. Bi, R., Tang, J., Zhang, W., Li, X., Chen, S.Y., Yu, D., Chen, X., Yao, Y.G., 2016. Mitochondrial genome variations and functional characterization in Han Chinese families with schizophrenia. Schizophr Res 171(1-3), 200-206. Bingham, E.M., Hopkins, D., Smith, D., Pernet, A., Hallett, W., Reed, L., Marsden, P.K., Amiel, S.A., 2002. The role of insulin in human brain glucose metabolism: an 18fluoro-deoxyglucose positron emission tomography study. Diabetes 51(12), 3384-3390. Black, C.N., Bot, M., Scheffer, P.G., Cuijpers, P., Penninx, B.W., 2015. Is depression associated with increased oxidative stress? A systematic review and meta-analysis. Psychoneuroendocrinology 51, 164-175. Boj, S.F., van Es, J.H., Huch, M., Li, V.S., Jose, A., Hatzis, P., Mokry, M., Haegebarth, A., van den Born, M., Chambon, P., Voshol, P., Dor, Y., Cuppen, E., Fillat, C., Clevers, H., 2012. Diabetes risk gene and Wnt effector Tcf7l2/TCF4 controls hepatic response to perinatal and adult metabolic demand. Cell 151(7), 1595-1607. Boksa, P., 2012. Abnormal synaptic pruning in schizophrenia: Urban myth or reality? J Psychiatry Neurosci 37(2), 75-77. Bonaccorso, S., Sodhi, M., Li, J., Bobo, W.V., Chen, Y., Tumuklu, M., Theleritis, C., Jayathilake, K., Meltzer, H.Y., 2015. The brain-derived neurotrophic factor (BDNF) Val66Met polymorphism is associated with increased body mass index and insulin resistance measures in bipolar disorder and schizophrenia. Bipolar Disord 17(5), 528-535. Bondy, C.A., Lee, W.H., Zhou, J., 1992. Ontogeny and cellular distribution of brain glucose transporter gene expression. Mol Cell Neurosci 3(4), 305-314. Bordonaro, M., 2009. Role of Wnt signaling in the development of type 2 diabetes. Vitam Horm 80, 563-581.

23

Boyle, P.J., Nagy, R.J., O'Connor, A.M., Kempers, S.F., Yeo, R.A., Qualls, C., 1994. Adaptation in brain glucose uptake following recurrent hypoglycemia. Proc Natl Acad Sci U S A 91(20), 93529356. Brown, A.M., Sickmann, H.M., Fosgerau, K., Lund, T.M., Schousboe, A., Waagepetersen, H.S., Ransom, B.R., 2005. Astrocyte glycogen metabolism is required for neural activity during aglycemia or intense stimulation in mouse white matter. J Neurosci Res 79(1-2), 74-80. Brown, A.S., 2011. The environment and susceptibility to schizophrenia. Prog Neurobiol 93(1), 2358. Brown, N.C., Andreazza, A.C., Young, L.T., 2014. An updated meta-analysis of oxidative stress markers in bipolar disorder. Psychiatry Res 218(1-2), 61-68. Bruning, J.C., Gautam, D., Burks, D.J., Gillette, J., Schubert, M., Orban, P.C., Klein, R., Krone, W., Muller-Wieland, D., Kahn, C.R., 2000. Role of brain insulin receptor in control of body weight and reproduction. Science 289(5487), 2122-2125. Cauchi, S., El Achhab, Y., Choquet, H., Dina, C., Krempler, F., Weitgasser, R., Nejjari, C., Patsch, W., Chikri, M., Meyre, D., Froguel, P., 2007. TCF7L2 is reproducibly associated with type 2 diabetes in various ethnic groups: a global meta-analysis. Journal of molecular medicine 85(7), 777-782. Cervantes, P., Gelber, S., Kin, F.N., Nair, V.N., Schwartz, G., 2001. Circadian secretion of cortisol in bipolar disorder. J Psychiatry Neurosci 26(5), 411-416. Chan, O., Inouye, K., Riddell, M.C., Vranic, M., Matthews, S.G., 2003. Diabetes and the hypothalamo-pituitary-adrenal (HPA) axis. Minerva Endocrinol 28(2), 87-102. Chen, P.C., Chan, Y.T., Chen, H.F., Ko, M.C., Li, C.Y., 2013. Population-based cohort analyses of the bidirectional relationship between type 2 diabetes and depression. Diabetes care 36(2), 376-382. Cheng, G., Huang, C., Deng, H., Wang, H., 2012. Diabetes as a risk factor for dementia and mild cognitive impairment: a meta-analysis of longitudinal studies. Internal medicine journal 42(5), 484-491. Chiodini, I., Adda, G., Scillitani, A., Coletti, F., Morelli, V., Di Lembo, S., Epaminonda, P., Masserini, B., Beck-Peccoz, P., Orsi, E., Ambrosi, B., Arosio, M., 2007. Cortisol secretion in patients with type 2 diabetes: relationship with chronic complications. Diabetes Care 30(1), 83-88. Cho, Y.M., Park, K.S., Lee, H.K., 2007. Genetic factors related to mitochondrial function and risk of diabetes mellitus. Diabetes Res Clin Pract 77 Suppl 1, S172-177. Chowdari, K.V., Bamne, M.N., Nimgaonkar, V.L., 2011. Genetic association studies of antioxidant pathway genes and schizophrenia. Antioxid Redox Signal 15(7), 2037-2045. Clay, H.B., Sillivan, S., Konradi, C., 2011. Mitochondrial dysfunction and pathology in bipolar disorder and schizophrenia. Int J Dev Neurosci 29(3), 311-324. Coppola, V., Tessarollo, L., 2004. Control of hyperphagia prevents obesity in BDNF heterozygous mice. Neuroreport 15(17), 2665-2668. Cox, D.J., Kovatchev, B.P., Gonder-Frederick, L.A., Summers, K.H., McCall, A., Grimm, K.J., Clarke, W.L., 2005. Relationships between hyperglycemia and cognitive performance among adults with type 1 and type 2 diabetes. Diabetes care 28(1), 71-77. Cryer, P.E., 2007. Hypoglycemia, functional brain failure, and brain death. J Clin Invest 117(4), 868-870. Cuellar-Barboza, A.B., Winham, S.J., McElroy, S.L., Geske, J.R., Jenkins, G.D., Colby, C.L., Prieto, M.L., Ryu, E., Cunningham, J.M., Frye, M.A., Biernacka, J.M., 2016. Accumulating evidence for a role of TCF7L2 variants in bipolar disorder with elevated body mass index. Bipolar Disord 18(2), 124-135. 24

Cukierman-Yaffe, T., Gerstein, H.C., Williamson, J.D., Lazar, R.M., Lovato, L., Miller, M.E., Coker, L.H., Murray, A., Sullivan, M.D., Marcovina, S.M., Launer, L.J., Action to Control Cardiovascular Risk in Diabetes-Memory in Diabetes, I., 2009. Relationship between baseline glycemic control and cognitive function in individuals with type 2 diabetes and other cardiovascular risk factors: the action to control cardiovascular risk in diabetes-memory in diabetes (ACCORD-MIND) trial. Diabetes care 32(2), 221-226. Cukierman, T., Gerstein, H.C., Williamson, J.D., 2005. Cognitive decline and dementia in diabetes-systematic overview of prospective observational studies. Diabetologia 48(12), 2460-2469. da Silva Xavier, G., Mondragon, A., Sun, G., Chen, L., McGinty, J.A., French, P.M., Rutter, G.A., 2012. Abnormal glucose tolerance and insulin secretion in pancreas-specific Tcf7l2-null mice. Diabetologia 55(10), 2667-2676. Daban, C., Vieta, E., Mackin, P., Young, A.H., 2005. Hypothalamic-pituitary-adrenal axis and bipolar disorder. Psychiatr Clin North Am 28(2), 469-480. de Galan, B.E., Schouwenberg, B.J., Tack, C.J., Smits, P., 2006. Pathophysiology and management of recurrent hypoglycaemia and hypoglycaemia unawareness in diabetes. Neth J Med 64(8), 269279. Del Prato, S., Marchetti, P., Bonadonna, R.C., 2002. Phasic insulin release and metabolic regulation in type 2 diabetes. Diabetes 51 Suppl 1, S109-116. Demily, C., Sedel, F., 2014. Psychiatric manifestations of treatable hereditary metabolic disorders in adults. Ann Gen Psychiatry 13, 27. El Messari, S., Leloup, C., Quignon, M., Brisorgueil, M.J., Penicaud, L., Arluison, M., 1998. Immunocytochemical localization of the insulin-responsive glucose transporter 4 (Glut4) in the rat central nervous system. J Comp Neurol 399(4), 492-512. Emiliani, F.E., Sedlak, T.W., Sawa, A., 2014. Oxidative stress and schizophrenia: recent breakthroughs from an old story. Curr Opin Psychiatry 27(3), 185-190. Enez Darcin, A., Yalcin Cavus, S., Dilbaz, N., Kaya, H., Dogan, E., 2015. Metabolic syndrome in drugnaive and drug-free patients with schizophrenia and in their siblings. Schizophrenia research 166(1-3), 201-206. Ernst, C., Marshall, C.R., Shen, Y., Metcalfe, K., Rosenfeld, J., Hodge, J.C., Torres, A., Blumenthal, I., Chiang, C., Pillalamarri, V., Crapper, L., Diallo, A.B., Ruderfer, D., Pereira, S., Sklar, P., Purcell, S., Wildin, R.S., Spencer, A.C., Quade, B.F., Harris, D.J., Lemyre, E., Wu, B.L., Stavropoulos, D.J., Geraghty, M.T., Shaffer, L.G., Morton, C.C., Scherer, S.W., Gusella, J.F., Talkowski, M.E., 2012. Highly penetrant alterations of a critical region including BDNF in human psychopathology and obesity. Arch Gen Psychiatry 69(12), 1238-1246. Falkowska, A., Gutowska, I., Goschorska, M., Nowacki, P., Chlubek, D., Baranowska-Bosiacka, I., 2015. Energy Metabolism of the Brain, Including the Cooperation between Astrocytes and Neurons, Especially in the Context of Glycogen Metabolism. International journal of molecular sciences 16(11), 25959-25981. Fanne, R.A., Nassar, T., Heyman, S.N., Hijazi, N., Higazi, A.A., 2011. Insulin and glucagon share the same mechanism of neuroprotection in diabetic rats: role of glutamate. Am J Physiol Regul Integr Comp Physiol 301(3), R668-673. Farrell, M.S., Werge, T., Sklar, P., Owen, M.J., Ophoff, R.A., O'Donovan, M.C., Corvin, A., Cichon, S., Sullivan, P.F., 2015. Evaluating historical candidate genes for schizophrenia. Mol Psychiatry 20(5), 555-562. Fattal, O., Budur, K., Vaughan, A.J., Franco, K., 2006. Review of the literature on major mental disorders in adult patients with mitochondrial diseases. Psychosomatics 47(1), 1-7. 25

Fernandez-Egea, E., Bernardo, M., Donner, T., Conget, I., Parellada, E., Justicia, A., Esmatjes, E., Garcia-Rizo, C., Kirkpatrick, B., 2009. Metabolic profile of antipsychotic-naive individuals with nonaffective psychosis. The British journal of psychiatry : the journal of mental science 194(5), 434438. Flatow, J., Buckley, P., Miller, B.J., 2013. Meta-analysis of oxidative stress in schizophrenia. Biol Psychiatry 74(6), 400-409. Foley, D.L., Mackinnon, A., Morgan, V.A., Watts, G.F., Castle, D.J., Waterreus, A., Galletly, C.A., 2016. Common familial risk factors for schizophrenia and diabetes mellitus. Aust N Z J Psychiatry 50(5), 488-494. Freemark, M., Avril, I., Fleenor, D., Driscoll, P., Petro, A., Opara, E., Kendall, W., Oden, J., Bridges, S., Binart, N., Breant, B., Kelly, P.A., 2002. Targeted deletion of the PRL receptor: effects on islet development, insulin production, and glucose tolerance. Endocrinology 143(4), 1378-1385. Fullerton, J.M., Tiwari, Y., Agahi, G., Heath, A., Berk, M., Mitchell, P.B., Schofield, P.R., 2010. Assessing oxidative pathway genes as risk factors for bipolar disorder. Bipolar Disord 12(5), 550556. Fünfschilling, U., Supplie, L.M., Mahad, D., Boretius, S., Saab, A.S., Edgar, J., Brinkmann, B.G., Kassmann, C.M., Tzvetanova, I.D., Möbius, W., Diaz, F., Meijer, D., Suter, U., Hamprecht, B., Sereda, M.W., Moraes, C.T., Frahm, J., Goebbels, S., Nave, K.A., 2012. Glycolytic oligodendrocytes maintain myelin and long-term axonal integrity. Nature 485(7399), 517-521. Garcia-Rizo, C., Fernandez-Egea, E., Bernardo, M., Kirkpatrick, B., 2015. The thrifty psychiatric phenotype. Acta Psychiatr Scand 131(1), 18-20. Garcia-Rizo, C., Fernandez-Egea, E., Oliveira, C., Justicia, A., Parellada, E., Bernardo, M., Kirkpatrick, B., 2012. Prolactin concentrations in newly diagnosed, antipsychotic-naive patients with nonaffective psychosis. Schizophr Res 134(1), 16-19. Garcia-Rizo, C., Kirkpatrick, B., Fernandez-Egea, E., Oliveira, C., Bernardo, M., 2016. Abnormal glycemic homeostasis at the onset of serious mental illnesses: A common pathway. Psychoneuroendocrinology 67, 70-75. Gibson, E.L., 2012. The psychobiology of comfort eating: implications for neuropharmacological interventions. Behav Pharmacol 23(5-6), 442-460. Gottschalk, M.G., Wesseling, H., Guest, P.C., Bahn, S., 2015. Proteomic enrichment analysis of psychotic and affective disorders reveals common signatures in presynaptic glutamatergic signaling and energy metabolism. Int J Neuropsychopharmacol 18(2). Gragnoli, C., Reeves, G.M., Reazer, J., Postolache, T.T., 2016. Dopamine-prolactin pathway potentially contributes to the schizophrenia and type 2 diabetes comorbidity. Transl Psychiatry 6, e785. Grant, S.F., Thorleifsson, G., Reynisdottir, I., Benediktsson, R., Manolescu, A., Sainz, J., Helgason, A., Stefansson, H., Emilsson, V., Helgadottir, A., Styrkarsdottir, U., Magnusson, K.P., Walters, G.B., Palsdottir, E., Jonsdottir, T., Gudmundsdottir, T., Gylfason, A., Saemundsdottir, J., Wilensky, R.L., Reilly, M.P., Rader, D.J., Bagger, Y., Christiansen, C., Gudnason, V., Sigurdsson, G., Thorsteinsdottir, U., Gulcher, J.R., Kong, A., Stefansson, K., 2006. Variant of transcription factor 7like 2 (TCF7L2) gene confers risk of type 2 diabetes. Nature genetics 38(3), 320-323. Grassi-Oliveira, R., Stein, L.M., Lopes, R.P., Teixeira, A.L., Bauer, M.E., 2008. Low plasma brainderived neurotrophic factor and childhood physical neglect are associated with verbal memory impairment in major depression--a preliminary report. Biol Psychiatry 64(4), 281-285. Gray, J., Yeo, G.S., Cox, J.J., Morton, J., Adlam, A.L., Keogh, J.M., Yanovski, J.A., El Gharbawy, A., Han, J.C., Tung, Y.C., Hodges, J.R., Raymond, F.L., O'Rahilly, S., Farooqi, I.S., 2006. Hyperphagia, 26

severe obesity, impaired cognitive function, and hyperactivity associated with functional loss of one copy of the brain-derived neurotrophic factor (BDNF) gene. Diabetes 55(12), 3366-3371. Green, M.F., Horan, W.P., Lee, J., 2015. Social cognition in schizophrenia. Nat Rev Neurosci 16(10), 620-631. Grillo, C.A., Piroli, G.G., Lawrence, R.C., Wrighten, S.A., Green, A.J., Wilson, S.P., Sakai, R.R., Kelly, S.J., Wilson, M.A., Mott, D.D., Reagan, L.P., 2015. Hippocampal Insulin Resistance Impairs Spatial Learning and Synaptic Plasticity. Diabetes 64(11), 3927-3936. Gudala, K., Bansal, D., Schifano, F., Bhansali, A., 2013. Diabetes mellitus and risk of dementia: A meta-analysis of prospective observational studies. Journal of diabetes investigation 4(6), 640650. Gupta, D., Kurhe, Y., Radhakrishnan, M., 2014. Antidepressant effects of insulin in streptozotocin induced diabetic mice: Modulation of brain serotonin system. Physiol Behav 129, 73-78. Haapakoski, R., Mathieu, J., Ebmeier, K.P., Alenius, H., Kivimaki, M., 2015. Cumulative metaanalysis of interleukins 6 and 1beta, tumour necrosis factor alpha and C-reactive protein in patients with major depressive disorder. Brain Behav Immun 49, 206-215. Haider, S., Ahmed, S., Tabassum, S., Memon, Z., Ikram, M., Haleem, D.J., 2013. Streptozotocininduced insulin deficiency leads to development of behavioral deficits in rats. Acta Neurol Belg 113(1), 35-41. Haijma, S.V., Van Haren, N., Cahn, W., Koolschijn, P.C., Hulshoff Pol, H.E., Kahn, R.S., 2013. Brain volumes in schizophrenia: a meta-analysis in over 18 000 subjects. Schizophr Bull 39(5), 11291138. Hansen, T., Ingason, A., Djurovic, S., Melle, I., Fenger, M., Gustafsson, O., Jakobsen, K.D., Rasmussen, H.B., Tosato, S., Rietschel, M., Frank, J., Owen, M., Bonetto, C., Suvisaari, J., Thygesen, J.H., Petursson, H., Lonnqvist, J., Sigurdsson, E., Giegling, I., Craddock, N., O'Donovan, M.C., Ruggeri, M., Cichon, S., Ophoff, R.A., Pietilainen, O., Peltonen, L., Nothen, M.M., Rujescu, D., St Clair, D., Collier, D.A., Andreassen, O.A., Werge, T., 2011. At-risk variant in TCF7L2 for type II diabetes increases risk of schizophrenia. Biological psychiatry 70(1), 59-63. Harris, J.J., Jolivet, R., Attwell, D., 2012. Synaptic energy use and supply. Neuron 75(5), 762-777. Hayley, S., Poulter, M.O., Merali, Z., Anisman, H., 2005. The pathogenesis of clinical depression: stressor- and cytokine-induced alterations of neuroplasticity. Neuroscience 135(3), 659-678. Heikkila, O., Lundbom, N., Timonen, M., Groop, P.H., Heikkinen, S., Makimattila, S., 2010. Evidence for abnormal glucose uptake or metabolism in thalamus during acute hyperglycaemia in type 1 diabetes--a 1H MRS study. Metabolic brain disease 25(2), 227-234. Heni, M., Kullmann, S., Preissl, H., Fritsche, A., Haring, H.U., 2015. Impaired insulin action in the human brain: causes and metabolic consequences. Nat Rev Endocrinol 11(12), 701-711. Herzog, R.I., Jiang, L., Herman, P., Zhao, C., Sanganahalli, B.G., Mason, G.F., Hyder, F., Rothman, D.L., Sherwin, R.S., Behar, K.L., 2013. Lactate preserves neuronal metabolism and function following antecedent recurrent hypoglycemia. J Clin Invest 123(5), 1988-1998. Holt, R.I., Peveler, R.C., 2009. Obesity, serious mental illness and antipsychotic drugs. Diabetes, obesity & metabolism 11(7), 665-679. Hotamisligil, G.S., 2006. Inflammation and metabolic disorders. Nature 444(7121), 860-867. Hotamisligil, G.S., Arner, P., Caro, J.F., Atkinson, R.L., Spiegelman, B.M., 1995. Increased adipose tissue expression of tumor necrosis factor-alpha in human obesity and insulin resistance. J Clin Invest 95(5), 2409-2415. Hotta, K., Nakamura, M., Nakamura, T., Matsuo, T., Nakata, Y., Kamohara, S., Miyatake, N., Kotani, K., Komatsu, R., Itoh, N., Mineo, I., Wada, J., Masuzaki, H., Yoneda, M., Nakajima, A., Funahashi, 27

T., Miyazaki, S., Tokunaga, K., Kawamoto, M., Ueno, T., Hamaguchi, K., Tanaka, K., Yamada, K., Hanafusa, T., Oikawa, S., Yoshimatsu, H., Nakao, K., Sakata, T., Matsuzawa, Y., Kamatani, N., Nakamura, Y., 2009. Association between obesity and polymorphisms in SEC16B, TMEM18, GNPDA2, BDNF, FAIM2 and MC4R in a Japanese population. J Hum Genet 54(12), 727-731. Hou, W.K., Xian, Y.X., Zhang, L., Lai, H., Hou, X.G., Xu, Y.X., Yu, T., Xu, F.Y., Song, J., Fu, C.L., Zhang, W.W., Chen, L., 2007. Influence of blood glucose on the expression of glucose trans-porter proteins 1 and 3 in the brain of diabetic rats. Chinese medical journal 120(19), 1704-1709. Howes, O.D., Kapur, S., 2009. The dopamine hypothesis of schizophrenia: version III--the final common pathway. Schizophr Bull 35(3), 549-562. Huang, C., Snider, F., Cross, J.C., 2009. Prolactin receptor is required for normal glucose homeostasis and modulation of beta-cell mass during pregnancy. Endocrinology 150(4), 16181626. Inczedy-Farkas, G., Remenyi, V., Gal, A., Varga, Z., Balla, P., Udvardy-Meszaros, A., Bereznai, B., Molnar, M.J., 2012. Psychiatric symptoms of patients with primary mitochondrial DNA disorders. Behav Brain Funct 8, 9. Insel, T.R., 2010. Rethinking schizophrenia. Nature 468(7321), 187-193. Iwamoto, K., Bundo, M., Kato, T., 2005. Altered expression of mitochondria-related genes in postmortem brains of patients with bipolar disorder or schizophrenia, as revealed by large-scale DNA microarray analysis. Human molecular genetics 14(2), 241-253. Ji, H.F., Zhuang, Q.S., Shen, L., 2016. Genetic overlap between type 2 diabetes and major depressive disorder identified by bioinformatics analysis. Oncotarget 7(14), 17410-17414. Jose, J., Nandeesha, H., Kattimani, S., Meiyappan, K., Sarkar, S., Sivasankar, D., 2015. Association between prolactin and thyroid hormones with severity of psychopathology and suicide risk in drug free male schizophrenia. Clin Chim Acta 444, 78-80. Jurczyk, A., Nowosielska, A., Przewozniak, N., Aryee, K.E., DiIorio, P., Blodgett, D., Yang, C., Campbell-Thompson, M., Atkinson, M., Shultz, L., Rittenhouse, A., Harlan, D., Greiner, D., Bortell, R., 2016. Beyond the brain: disrupted in schizophrenia 1 regulates pancreatic beta-cell function via glycogen synthase kinase-3beta. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 30(2), 983-993. Kalimo, H., Olsson, Y., 1980. Effects of severe hypoglycemia on the human brain. Neuropathological case reports. Acta Neurol Scand 62(6), 345-356. Kan, C., Pedersen, N.L., Christensen, K., Bornstein, S.R., Licinio, J., MacCabe, J.H., Ismail, K., Rijsdijk, F., 2016. Genetic overlap between type 2 diabetes and depression in Swedish and Danish twin registries. Mol Psychiatry 21(7), 903-909. Kang, D.H., Kwon, K.W., Gu, B.M., Choi, J.S., Jang, J.H., Kwon, J.S., 2008. Structural abnormalities of the right inferior colliculus in schizophrenia. Psychiatry Res 164(2), 160-165. Kato, T., 2008. Molecular neurobiology of bipolar disorder: a disease of 'mood-stabilizing neurons'? Trends Neurosci 31(10), 495-503. Kauer-Sant'Anna, M., Tramontina, J., Andreazza, A.C., Cereser, K., da Costa, S., Santin, A., Yatham, L.N., Kapczinski, F., 2007. Traumatic life events in bipolar disorder: impact on BDNF levels and psychopathology. Bipolar Disord 9 Suppl 1, 128-135. Kaufman, B.A., Li, C., Soleimanpour, S.A., 2015. Mitochondrial regulation of beta-cell function: maintaining the momentum for insulin release. Molecular aspects of medicine 42, 91-104. Kavanagh, D.H., Tansey, K.E., O'Donovan, M.C., Owen, M.J., 2015. Schizophrenia genetics: emerging themes for a complex disorder. Mol Psychiatry 20(1), 72-76. 28

Keller, J., Flores, B., Gomez, R.G., Solvason, H.B., Kenna, H., Williams, G.H., Schatzberg, A.F., 2006. Cortisol circadian rhythm alterations in psychotic major depression. Biol Psychiatry 60(3), 275281. Kern, P.A., Saghizadeh, M., Ong, J.M., Bosch, R.J., Deem, R., Simsolo, R.B., 1995. The expression of tumor necrosis factor in human adipose tissue. Regulation by obesity, weight loss, and relationship to lipoprotein lipase. J Clin Invest 95(5), 2111-2119. Kim, D.R., Bale, T.L., Epperson, C.N., 2015. Prenatal programming of mental illness: current understanding of relationship and mechanisms. Curr Psychiatry Rep 17(2), 5. Kim, K.W., Donato, J., Jr., Berglund, E.D., Choi, Y.H., Kohno, D., Elias, C.F., Depinho, R.A., Elmquist, J.K., 2012. FOXO1 in the ventromedial hypothalamus regulates energy balance. J Clin Invest 122(7), 2578-2589. Kirkpatrick, B., Fernandez-Egea, E., Garcia-Rizo, C., Bernardo, M., 2009. Differences in glucose tolerance between deficit and nondeficit schizophrenia. Schizophrenia research 107(2-3), 122127. Kirkpatrick, B., Miller, B.J., 2013. Inflammation and schizophrenia. Schizophr Bull 39(6), 11741179. Kleinridders, A., Cai, W., Cappellucci, L., Ghazarian, A., Collins, W.R., Vienberg, S.G., Pothos, E.N., Kahn, C.R., 2015. Insulin resistance in brain alters dopamine turnover and causes behavioral disorders. Proc Natl Acad Sci U S A 112(11), 3463-3468. Kleinridders, A., Ferris, H.A., Cai, W., Kahn, C.R., 2014. Insulin action in brain regulates systemic metabolism and brain function. Diabetes 63(7), 2232-2243. Knol, M.J., Twisk, J.W., Beekman, A.T., Heine, R.J., Snoek, F.J., Pouwer, F., 2006. Depression as a risk factor for the onset of type 2 diabetes mellitus. A meta-analysis. Diabetologia 49(5), 837-845. Kobayashi, M., Nikami, H., Morimatsu, M., Saito, M., 1996. Expression and localization of insulinregulatable glucose transporter (GLUT4) in rat brain. Neurosci Lett 213(2), 103-106. Kodl, C.T., Seaquist, E.R., 2008. Cognitive dysfunction and diabetes mellitus. Endocr Rev 29(4), 494-511. Kohen, D., 2004. Diabetes mellitus and schizophrenia: historical perspective. Br J Psychiatry Suppl 47, S64-66. Kolomeets, N.S., Uranova, N., 2010. Ultrastructural abnormalities of astrocytes in the hippocampus in schizophrenia and duration of illness: a postortem morphometric study. World J Biol Psychiatry 11(2 Pt 2), 282-292. Konner, A.C., Hess, S., Tovar, S., Mesaros, A., Sanchez-Lasheras, C., Evers, N., Verhagen, L.A., Bronneke, H.S., Kleinridders, A., Hampel, B., Kloppenburg, P., Bruning, J.C., 2011. Role for insulin signaling in catecholaminergic neurons in control of energy homeostasis. Cell Metab 13(6), 720728. Konradi, C., Eaton, M., MacDonald, M.L., Walsh, J., Benes, F.M., Heckers, S., 2004. Molecular evidence for mitochondrial dysfunction in bipolar disorder. Archives of general psychiatry 61(3), 300-308. Ku, C.S., Loy, E.Y., Pawitan, Y., Chia, K.S., 2010. The pursuit of genome-wide association studies: where are we now? Journal of human genetics 55(4), 195-206. Kullmann, S., Heni, M., Fritsche, A., Preissl, H., 2015. Insulin action in the human brain: evidence from neuroimaging studies. J Neuroendocrinol 27(6), 419-423. Kumagai, A.K., Kang, Y.S., Boado, R.J., Pardridge, W.M., 1995. Upregulation of blood-brain barrier GLUT1 glucose transporter protein and mRNA in experimental chronic hypoglycemia. Diabetes 44(12), 1399-1404. 29

Kung, L., Roberts, R.C., 1999. Mitochondrial pathology in human schizophrenic striatum: a postmortem ultrastructural study. Synapse 31(1), 67-75. Kwak, S.H., Park, K.S., 2016. Role of mitochondrial DNA variation in the pathogenesis of diabetes mellitus. Front Biosci (Landmark Ed) 21, 1151-1167. Lamers, F., Vogelzangs, N., Merikangas, K.R., de Jonge, P., Beekman, A.T., Penninx, B.W., 2013. Evidence for a differential role of HPA-axis function, inflammation and metabolic syndrome in melancholic versus atypical depression. Mol Psychiatry 18(6), 692-699. Landek-Salgado, M.A., Faust, T.E., Sawa, A., 2016. Molecular substrates of schizophrenia: homeostatic signaling to connectivity. Mol Psychiatry 21(1), 10-28. Languren, G., Montiel, T., Julio-Amilpas, A., Massieu, L., 2013. Neuronal damage and cognitive impairment associated with hypoglycemia: An integrated view. Neurochem Int 63(4), 331-343. Lee, Y., Morrison, B.M., Li, Y., Lengacher, S., Farah, M.H., Hoffman, P.N., Liu, Y., Tsingalia, A., Jin, L., Zhang, P.W., Pellerin, L., Magistretti, P.J., Rothstein, J.D., 2012. Oligodendroglia metabolically support axons and contribute to neurodegeneration. Nature 487(7408), 443-448. Leino, R.L., Gerhart, D.Z., Duelli, R., Enerson, B.E., Drewes, L.R., 2001. Diet-induced ketosis increases monocarboxylate transporter (MCT1) levels in rat brain. Neurochem Int 38(6), 519-527. Li, M., Shi, C.J., Shi, Y.Y., Luo, X.J., Zheng, X.B., Li, Z.Q., Liu, J.J., Chong, S.A., Lee, J., Wang, Y., Liu, X.Y., Yin, L.D., Pu, X.F., Diao, H.B., Xu, Q., Su, B., 2012. ZNF804A and schizophrenia susceptibility in Asian populations. Am J Med Genet B Neuropsychiatr Genet 159B(7), 794-802. Li, W., Maloney, R.E., Aw, T.Y., 2015. High glucose, glucose fluctuation and carbonyl stress enhance brain microvascular endothelial barrier dysfunction: Implications for diabetic cerebral microvasculature. Redox biology 5, 80-90. Li, W., Zhou, N., Yu, Q., Li, X., Yu, Y., Sun, S., Kou, C., Chen da, C., Xiu, M.H., Kosten, T.R., Zhang, X.Y., 2013. Association of BDNF gene polymorphisms with schizophrenia and clinical symptoms in a Chinese population. Am J Med Genet B Neuropsychiatr Genet 162B(6), 538-545. Lin, P.I., Shuldiner, A.R., 2010. Rethinking the genetic basis for comorbidity of schizophrenia and type 2 diabetes. Schizophrenia research 123(2-3), 234-243. Liu, Y., Li, Z., Zhang, M., Deng, Y., Yi, Z., Shi, T., 2013. Exploring the pathogenetic association between schizophrenia and type 2 diabetes mellitus diseases based on pathway analysis. BMC medical genomics 6 Suppl 1, S17. Liu, Z., Habener, J.F., 2010. Wnt signaling in pancreatic islets. Advances in experimental medicine and biology 654, 391-419. Lorenzo, M., Fernandez-Veledo, S., Vila-Bedmar, R., Garcia-Guerra, L., De Alvaro, C., NietoVazquez, I., 2008. Insulin resistance induced by tumor necrosis factor-alpha in myocytes and brown adipocytes. J Anim Sci 86(14 Suppl), E94-104. Lutas, A., Yellen, G., 2013. The ketogenic diet: metabolic influences on brain excitability and epilepsy. Trends Neurosci 36(1), 32-40. Ma, J.H., Kim, Y.J., Yoo, W.J., Ihn, Y.K., Kim, J.Y., Song, H.H., Kim, B.S., 2009. MR imaging of hypoglycemic encephalopathy: lesion distribution and prognosis prediction by diffusion-weighted imaging. Neuroradiology 51(10), 641-649. Magistretti, P.J., Allaman, I., 2015. A cellular perspective on brain energy metabolism and functional imaging. Neuron 86(4), 883-901. Maiese, K., 2015. New Insights for Oxidative Stress and Diabetes Mellitus. Oxid Med Cell Longev 2015, 875961. Manji, H., Kato, T., Di Prospero, N.A., Ness, S., Beal, M.F., Krams, M., Chen, G., 2012. Impaired mitochondrial function in psychiatric disorders. Nat Rev Neurosci 13(5), 293-307. 30

Mao, Y., Ge, X., Frank, C.L., Madison, J.M., Koehler, A.N., Doud, M.K., Tassa, C., Berry, E.M., Soda, T., Singh, K.K., Biechele, T., Petryshen, T.L., Moon, R.T., Haggarty, S.J., Tsai, L.H., 2009. Disrupted in schizophrenia 1 regulates neuronal progenitor proliferation via modulation of GSK3beta/betacatenin signaling. Cell 136(6), 1017-1031. Martin, S.D., McGee, S.L., 2014. The role of mitochondria in the aetiology of insulin resistance and type 2 diabetes. Biochimica et biophysica acta 1840(4), 1303-1312. Mathieson, I., Munafo, M.R., Flint, J., 2012. Meta-analysis indicates that common variants at the DISC1 locus are not associated with schizophrenia. Mol Psychiatry 17(6), 634-641. Maurer, I., Zierz, S., Moller, H., 2001. Evidence for a mitochondrial oxidative phosphorylation defect in brains from patients with schizophrenia. Schizophr Res 48(1), 125-136. McArdle, M.A., Finucane, O.M., Connaughton, R.M., McMorrow, A.M., Roche, H.M., 2013. Mechanisms of obesity-induced inflammation and insulin resistance: insights into the emerging role of nutritional strategies. Front Endocrinol (Lausanne) 4, 52. McCrimmon, R.J., Sherwin, R.S., 2010. Hypoglycemia in type 1 diabetes. Diabetes 59(10), 23332339. McNay, E.C., Ong, C.T., McCrimmon, R.J., Cresswell, J., Bogan, J.S., Sherwin, R.S., 2010. Hippocampal memory processes are modulated by insulin and high-fat-induced insulin resistance. Neurobiol Learn Mem 93(4), 546-553. Meduna, L.J., Gery, F.J., Urse, V.G., 1942. Biochemical Disturbances in Mental disorders. I. Antiinsulin effect of blood in case of schizophrenia. Arch Neurol Psych 47(1), 38-52. Mergenthaler, P., Lindauer, U., Dienel, G.A., Meisel, A., 2013. Sugar for the brain: the role of glucose in physiological and pathological brain function. Trends Neurosci 36(10), 587-597. Mezuk, B., Eaton, W.W., Albrecht, S., Golden, S.H., 2008. Depression and type 2 diabetes over the lifespan: a meta-analysis. Diabetes care 31(12), 2383-2390. Middleton, F.A., Mirnics, K., Pierri, J.N., Lewis, D.A., Levitt, P., 2002. Gene expression profiling reveals alterations of specific metabolic pathways in schizophrenia. J Neurosci 22(7), 2718-2729. Mitchell, R.K., Mondragon, A., Chen, L., McGinty, J.A., French, P.M., Ferrer, J., Thorens, B., Hodson, D.J., Rutter, G.A., Da Silva Xavier, G., 2015. Selective disruption of Tcf7l2 in the pancreatic beta cell impairs secretory function and lowers beta cell mass. Human molecular genetics 24(5), 13901399. Mondelli, V., Cattaneo, A., Belvederi Murri, M., Di Forti, M., Handley, R., Hepgul, N., Miorelli, A., Navari, S., Papadopoulos, A.S., Aitchison, K.J., Morgan, C., Murray, R.M., Dazzan, P., Pariante, C.M., 2011. Stress and inflammation reduce brain-derived neurotrophic factor expression in firstepisode psychosis: a pathway to smaller hippocampal volume. J Clin Psychiatry 72(12), 16771684. Montgomery, M.K., Turner, N., 2015. Mitochondrial dysfunction and insulin resistance: an update. Endocr Connect 4(1), R1-R15. Mussmann, R., Geese, M., Harder, F., Kegel, S., Andag, U., Lomow, A., Burk, U., Onichtchouk, D., Dohrmann, C., Austen, M., 2007. Inhibition of GSK3 promotes replication and survival of pancreatic beta cells. J Biol Chem 282(16), 12030-12037. Nagalski, A., Irimia, M., Szewczyk, L., Ferran, J.L., Misztal, K., Kuznicki, J., Wisniewska, M.B., 2013. Postnatal isoform switch and protein localization of LEF1 and TCF7L2 transcription factors in cortical, thalamic, and mesencephalic regions of the adult mouse brain. Brain Struct Funct 218(6), 1531-1549.

31

Nagalski, A., Puelles, L., Dabrowski, M., Wegierski, T., Kuznicki, J., Wisniewska, M.B., 2015. Molecular anatomy of the thalamic complex and the underlying transcription factors. Brain structure & function. Niciu, M.J., Ionescu, D.F., Mathews, D.C., Richards, E.M., Zarate, C.A., 2013. Second messenger/signal transduction pathways in major mood disorders: moving from membrane to mechanism of action, part I: major depressive disorder. CNS Spectr 18(5), 231-241. Obici, S., Zhang, B.B., Karkanias, G., Rossetti, L., 2002. Hypothalamic insulin signaling is required for inhibition of glucose production. Nat Med 8(12), 1376-1382. Ott, C., Jacobs, K., Haucke, E., Navarrete Santos, A., Grune, T., Simm, A., 2014. Role of advanced glycation end products in cellular signaling. Redox Biol 2, 411-429. Ott, V., Benedict, C., Schultes, B., Born, J., Hallschmid, M., 2012. Intranasal administration of insulin to the brain impacts cognitive function and peripheral metabolism. Diabetes Obes Metab 14(3), 214-221. Pais, I., Hallschmid, M., Jauch-Chara, K., Schmid, S.M., Oltmanns, K.M., Peters, A., Born, J., Schultes, B., 2007. Mood and cognitive functions during acute euglycaemia and mild hyperglycaemia in type 2 diabetic patients. Experimental and clinical endocrinology & diabetes : official journal, German Society of Endocrinology [and] German Diabetes Association 115(1), 4246. Pan, A., Keum, N., Okereke, O.I., Sun, Q., Kivimaki, M., Rubin, R.R., Hu, F.B., 2012. Bidirectional association between depression and metabolic syndrome: a systematic review and meta-analysis of epidemiological studies. Diabetes care 35(5), 1171-1180. Pariante, C.M., Lightman, S.L., 2008. The HPA axis in major depression: classical theories and new developments. Trends Neurosci 31(9), 464-468. Pelligrino, D.A., Segil, L.J., Albrecht, R.F., 1990. Brain glucose utilization and transport and cortical function in chronic vs. acute hypoglycemia. Am J Physiol 259(5 Pt 1), E729-735. Pennington, K., Beasley, C.L., Dicker, P., Fagan, A., English, J., Pariante, C.M., Wait, R., Dunn, M.J., Cotter, D.R., 2008. Prominent synaptic and metabolic abnormalities revealed by proteomic analysis of the dorsolateral prefrontal cortex in schizophrenia and bipolar disorder. Mol Psychiatry 13(12), 1102-1117. Prabakaran, S., Swatton, J.E., Ryan, M.M., Huffaker, S.J., Huang, J.T., Griffin, J.L., Wayland, M., Freeman, T., Dudbridge, F., Lilley, K.S., Karp, N.A., Hester, S., Tkachev, D., Mimmack, M.L., Yolken, R.H., Webster, M.J., Torrey, E.F., Bahn, S., 2004. Mitochondrial dysfunction in schizophrenia: evidence for compromised brain metabolism and oxidative stress. Mol Psychiatry 9(7), 684-697, 643. Pradhan, A.D., Manson, J.E., Rifai, N., Buring, J.E., Ridker, P.M., 2001. C-reactive protein, interleukin 6, and risk of developing type 2 diabetes mellitus. JAMA 286(3), 327-334. Puchowicz, M.A., Xu, K., Magness, D., Miller, C., Lust, W.D., Kern, T.S., LaManna, J.C., 2004. Comparison of glucose influx and blood flow in retina and brain of diabetic rats. Journal of cerebral blood flow and metabolism : official journal of the International Society of Cerebral Blood Flow and Metabolism 24(4), 449-457. Purcell, S.M., Moran, J.L., Fromer, M., Ruderfer, D., Solovieff, N., Roussos, P., O'Dushlaine, C., Chambert, K., Bergen, S.E., Kahler, A., Duncan, L., Stahl, E., Genovese, G., Fernandez, E., Collins, M.O., Komiyama, N.H., Choudhary, J.S., Magnusson, P.K., Banks, E., Shakir, K., Garimella, K., Fennell, T., DePristo, M., Grant, S.G., Haggarty, S.J., Gabriel, S., Scolnick, E.M., Lander, E.S., Hultman, C.M., Sullivan, P.F., McCarroll, S.A., Sklar, P., 2014. A polygenic burden of rare disruptive mutations in schizophrenia. Nature 506(7487), 185-190. 32

Ren, H., Plum-Morschel, L., Gutierrez-Juarez, R., Lu, T.Y., Kim-Muller, J.Y., Heinrich, G., Wardlaw, S.L., Silver, R., Accili, D., 2013. Blunted refeeding response and increased locomotor activity in mice lacking FoxO1 in synapsin-Cre-expressing neurons. Diabetes 62(10), 3373-3383. Ren, H., Yan, S., Zhang, B., Lu, T.Y., Arancio, O., Accili, D., 2014. Glut4 expression defines an insulinsensitive hypothalamic neuronal population. Mol Metab 3(4), 452-459. Roberts, R.C., Barksdale, K.A., Roche, J.K., Lahti, A.C., 2015. Decreased synaptic and mitochondrial density in the postmortem anterior cingulate cortex in schizophrenia. Schizophr Res 168(1-2), 543-553. Rojo, L.E., Gaspar, P.A., Silva, H., Risco, L., Arena, P., Cubillos-Robles, K., Jara, B., 2015. Metabolic syndrome and obesity among users of second generation antipsychotics: A global challenge for modern psychopharmacology. Pharmacol Res 101, 74-85. Rollins, B., Martin, M.V., Sequeira, P.A., Moon, E.A., Morgan, L.Z., Watson, S.J., Schatzberg, A., Akil, H., Myers, R.M., Jones, E.G., Wallace, D.C., Bunney, W.E., Vawter, M.P., 2009. Mitochondrial variants in schizophrenia, bipolar disorder, and major depressive disorder. PloS one 4(3), e4913. Rotella, F., Mannucci, E., 2013. Depression as a risk factor for diabetes: a meta-analysis of longitudinal studies. The Journal of clinical psychiatry 74(1), 31-37. Roy, T., Lloyd, C.E., 2012. Epidemiology of depression and diabetes: a systematic review. Journal of affective disorders 142 Suppl, S8-21. Ryan, M.C., Collins, P., Thakore, J.H., 2003. Impaired fasting glucose tolerance in first-episode, drug-naive patients with schizophrenia. The American journal of psychiatry 160(2), 284-289. Saab, A.S., Tzvetanova, I.D., Nave, K.A., 2013. The role of myelin and oligodendrocytes in axonal energy metabolism. Curr Opin Neurobiol 23(6), 1065-1072. Sachs, N.A., Sawa, A., Holmes, S.E., Ross, C.A., DeLisi, L.E., Margolis, R.L., 2005. A frameshift mutation in Disrupted in Schizophrenia 1 in an American family with schizophrenia and schizoaffective disorder. Mol Psychiatry 10(8), 758-764. Saghizadeh, M., Ong, J.M., Garvey, W.T., Henry, R.R., Kern, P.A., 1996. The expression of TNF alpha by human muscle. Relationship to insulin resistance. J Clin Invest 97(4), 1111-1116. Sakel, M., 1994. The methodical use of hypoglycemia in the treatment of psychoses. 1937. Am J Psychiatry 151(6 Suppl), 240-247. Salcedo-Tello, P., Ortiz-Matamoros, A., Arias, C., 2011. GSK3 Function in the Brain during Development, Neuronal Plasticity, and Neurodegeneration. Int J Alzheimers Dis 2011, 189728. Sanchez-Lasheras, C., Konner, A.C., Bruning, J.C., 2010. Integrative neurobiology of energy homeostasis-neurocircuits, signals and mediators. Front Neuroendocrinol 31(1), 4-15. Sanderson, T.H., Kumar, R., Murariu-Dobrin, A.C., Page, A.B., Krause, G.S., Sullivan, J.M., 2009. Insulin activates the PI3K-Akt survival pathway in vulnerable neurons following global brain ischemia. Neurol Res 31(9), 947-958. Schafer, D.P., Lehrman, E.K., Kautzman, A.G., Koyama, R., Mardinly, A.R., Yamasaki, R., Ransohoff, R.M., Greenberg, M.E., Barres, B.A., Stevens, B., 2012. Microglia sculpt postnatal neural circuits in an activity and complement-dependent manner. Neuron 74(4), 691-705. Schizophrenia Working Group of the Psychiatric Genomics, C., 2014. Biological insights from 108 schizophrenia-associated genetic loci. Nature 511(7510), 421-427. Segel, S.A., Fanelli, C.G., Dence, C.S., Markham, J., Videen, T.O., Paramore, D.S., Powers, W.J., Cryer, P.E., 2001. Blood-to-brain glucose transport, cerebral glucose metabolism, and cerebral blood flow are not increased after hypoglycemia. Diabetes 50(8), 1911-1917. Sekar, A., Bialas, A.R., de Rivera, H., Davis, A., Hammond, T.R., Kamitaki, N., Tooley, K., Presumey, J., Baum, M., Van Doren, V., Genovese, G., Rose, S.A., Handsaker, R.E., Schizophrenia Working 33

Group of the Psychiatric Genomics, C., Daly, M.J., Carroll, M.C., Stevens, B., McCarroll, S.A., 2016. Schizophrenia risk from complex variation of complement component 4. Nature 530(7589), 177183. Sequeira, A., Martin, M.V., Rollins, B., Moon, E.A., Bunney, W.E., Macciardi, F., Lupoli, S., Smith, E.N., Kelsoe, J., Magnan, C.N., van Oven, M., Baldi, P., Wallace, D.C., Vawter, M.P., 2012. Mitochondrial mutations and polymorphisms in psychiatric disorders. Front Genet 3, 103. Sequeira, A., Rollins, B., Magnan, C., van Oven, M., Baldi, P., Myers, R.M., Barchas, J.D., Schatzberg, A.F., Watson, S.J., Akil, H., Bunney, W.E., Vawter, M.P., 2015. Mitochondrial mutations in subjects with psychiatric disorders. PLoS One 10(5), e0127280. Sha, H., Xu, J., Tang, J., Ding, J., Gong, J., Ge, X., Kong, D., Gao, X., 2007. Disruption of a novel regulatory locus results in decreased Bdnf expression, obesity, and type 2 diabetes in mice. Physiol Genomics 31(2), 252-263. Shaefer, C., Hinnen, D., Sadler, C., 2016. Hypoglycemia and diabetes: increased need for awareness. Curr Med Res Opin, 1-8. Shah, A.A., Parent, M.B., 2003. Septal infusions of glucose or pyruvate, but not fructose, produce avoidance deficits when co-infused with the GABA agonist muscimol. Neurobiology of learning and memory 79(3), 243-251. Sharma, A.N., Elased, K.M., Garrett, T.L., Lucot, J.B., 2010. Neurobehavioral deficits in db/db diabetic mice. Physiol Behav 101(3), 381-388. Shen, J., Fang, Y., Ge, W., 2015. Polymorphism in the transcription factor 7-like 2 (TCF7L2) gene is associated with impaired proinsulin conversion--A meta-analysis. Diabetes Res Clin Pract 109(1), 117-123. Shepard, P.D., Holcomb, H.H., Gold, J.M., 2006. Schizophrenia in translation: the presence of absence: habenular regulation of dopamine neurons and the encoding of negative outcomes. Schizophr Bull 32(3), 417-421. Sickmann, H.M., Walls, A.B., Schousboe, A., Bouman, S.D., Waagepetersen, H.S., 2009. Functional significance of brain glycogen in sustaining glutamatergic neurotransmission. J Neurochem 109 Suppl 1, 80-86. Simpson, I.A., Appel, N.M., Hokari, M., Oki, J., Holman, G.D., Maher, F., Koehler-Stec, E.M., Vannucci, S.J., Smith, Q.R., 1999. Blood-brain barrier glucose transporter: effects of hypo- and hyperglycemia revisited. J Neurochem 72(1), 238-247. Sladek, R., Rocheleau, G., Rung, J., Dina, C., Shen, L., Serre, D., Boutin, P., Vincent, D., Belisle, A., Hadjadj, S., Balkau, B., Heude, B., Charpentier, G., Hudson, T.J., Montpetit, A., Pshezhetsky, A.V., Prentki, M., Posner, B.I., Balding, D.J., Meyre, D., Polychronakos, C., Froguel, P., 2007. A genomewide association study identifies novel risk loci for type 2 diabetes. Nature 445(7130), 881-885. Smolina, K., Wotton, C.J., Goldacre, M.J., 2015. Risk of dementia in patients hospitalised with type 1 and type 2 diabetes in England, 1998-2011: a retrospective national record linkage cohort study. Diabetologia 58(5), 942-950. Soares, D.C., Carlyle, B.C., Bradshaw, N.J., Porteous, D.J., 2011. DISC1: Structure, Function, and Therapeutic Potential for Major Mental Illness. ACS Chem Neurosci 2(11), 609-632. Sommerfield, A.J., Deary, I.J., Frier, B.M., 2004. Acute hyperglycemia alters mood state and impairs cognitive performance in people with type 2 diabetes. Diabetes care 27(10), 2335-2340. Steen, N.E., Methlie, P., Lorentzen, S., Dieset, I., Aas, M., Nerhus, M., Haram, M., Agartz, I., Melle, I., Berg, J.P., Andreassen, O.A., 2014. Altered systemic cortisol metabolism in bipolar disorder and schizophrenia spectrum disorders. J Psychiatr Res 52, 57-62. 34

Steen, N.E., Methlie, P., Lorentzen, S., Hope, S., Barrett, E.A., Larsson, S., Mork, E., Almås, B., Løvås, K., Agartz, I., Melle, I., Berg, J.P., Andreassen, O.A., 2011. Increased systemic cortisol metabolism in patients with schizophrenia and bipolar disorder: a mechanism for increased stress vulnerability? J Clin Psychiatry 72(11), 1515-1521. Stetler, C., Miller, G.E., 2011. Depression and hypothalamic-pituitary-adrenal activation: a quantitative summary of four decades of research. Psychosom Med 73(2), 114-126. Stewart, M.A., Sherman, W.R., Anthony, S., 1966. Free sugars in alloxan diabetic rat nerve. Biochem Biophys Res Commun 22(5), 4-91. Stoica, L., Zhu, P.J., Huang, W., Zhou, H., Kozma, S.C., Costa-Mattioli, M., 2011. Selective pharmacogenetic inhibition of mammalian target of Rapamycin complex I (mTORC1) blocks longterm synaptic plasticity and memory storage. Proc Natl Acad Sci U S A 108(9), 3791-3796. Sudhof, T.C., 2004. The synaptic vesicle cycle. Annu Rev Neurosci 27, 509-547. Taylor, D., Young, C., Mohamed, R., Paton, C., Walwyn, R., 2005. Undiagnosed impaired fasting glucose and diabetes mellitus amongst inpatients receiving antipsychotic drugs. Journal of psychopharmacology 19(2), 182-186. Tennant, K.A., Brown, C.E., 2013. Diabetes augments in vivo microvascular blood flow dynamics after stroke. The Journal of neuroscience : the official journal of the Society for Neuroscience 33(49), 19194-19204. Terra, L.F., Garay-Malpartida, M.H., Wailemann, R.A., Sogayar, M.C., Labriola, L., 2011. Recombinant human prolactin promotes human beta cell survival via inhibition of extrinsic and intrinsic apoptosis pathways. Diabetologia 54(6), 1388-1397. Thomson, P.A., Malavasi, E.L., Grunewald, E., Soares, D.C., Borkowska, M., Millar, J.K., 2013. DISC1 genetics, biology and psychiatric illness. Front Biol (Beijing) 8(1), 1-31. Tomlinson, D.R., Gardiner, N.J., 2008. Glucose neurotoxicity. Nat Rev Neurosci 9(1), 36-45. Tripathy, D., Chavez, A.O., 2010. Defects in insulin secretion and action in the pathogenesis of type 2 diabetes mellitus. Curr Diab Rep 10(3), 184-191. Tuomi, T., Santoro, N., Caprio, S., Cai, M., Weng, J., Groop, L., 2014. The many faces of diabetes: a disease with increasing heterogeneity. Lancet 383(9922), 1084-1094. Uehara, Y., Nipper, V., McCall, A.L., 1997. Chronic insulin hypoglycemia induces GLUT-3 protein in rat brain neurons. Am J Physiol 272(4 Pt 1), E716-719. Upthegrove, R., Manzanares-Teson, N., Barnes, N.M., 2014. Cytokine function in medicationnaive first episode psychosis: a systematic review and meta-analysis. Schizophr Res 155(1-3), 101108. Uranova, N.A., Vikhreva, O.V., Rachmanova, V.I., Orlovskaya, D.D., 2011. Ultrastructural alterations of myelinated fibers and oligodendrocytes in the prefrontal cortex in schizophrenia: a postmortem morphometric study. Schizophr Res Treatment 2011, 325789. van de Ven, K.C., van der Graaf, M., Tack, C.J., Heerschap, A., de Galan, B.E., 2012. Steady-state brain glucose concentrations during hypoglycemia in healthy humans and patients with type 1 diabetes. Diabetes 61(8), 1974-1977. van Welie, H., Derks, E.M., Verweij, K.H., de Valk, H.W., Kahn, R.S., Cahn, W., 2013. The prevalence of diabetes mellitus is increased in relatives of patients with a non-affective psychotic disorder. Schizophr Res 143(2-3), 354-357. Vancampfort, D., Wampers, M., Mitchell, A.J., Correll, C.U., De Herdt, A., Probst, M., De Hert, M., 2013. A meta-analysis of cardio-metabolic abnormalities in drug naive, first-episode and multiepisode patients with schizophrenia versus general population controls. World psychiatry : official journal of the World Psychiatric Association 12(3), 240-250. 35

Vannucci, S.J., Koehler-Stec, E.M., Li, K., Reynolds, T.H., Clark, R., Simpson, I.A., 1998. GLUT4 glucose transporter expression in rodent brain: effect of diabetes. Brain Res 797(1), 1-11. Vikhreva, O.V., Rakhmanova, V.I., Orlovskaya, D.D., Uranova, N.A., 2016. Ultrastructural alterations of oligodendrocytes in prefrontal white matter in schizophrenia: A post-mortem morphometric study. Schizophr Res. Vinik, A., Flemmer, M., 2002. Diabetes and macrovascular disease. J Diabetes Complications 16(3), 235-245. Voleti, B., Duman, R.S., 2012. The roles of neurotrophic factor and Wnt signaling in depression. Clin Pharmacol Ther 91(2), 333-338. Voruganti, L.P., Punthakee, Z., Van Lieshout, R.J., MacCrimmon, D., Parker, G., Awad, A.G., Gerstein, H.C., 2007. Dysglycemia in a community sample of people treated for schizophrenia: the Diabetes in Schizophrenia in Central-South Ontario (DiSCO) study. Schizophrenia research 96(13), 215-222. Wang, F., Guo, X., Shen, X., Kream, R.M., Mantione, K.J., Stefano, G.B., 2014. Vascular dysfunction associated with type 2 diabetes and Alzheimer's disease: a potential etiological linkage. Medical science monitor basic research 20, 118-129. Wang, J., Ciofi, P., Crowley, W.R., 1996. Neuropeptide Y suppresses prolactin secretion from rat anterior pituitary cells: evidence for interactions with dopamine through inhibitory coupling to calcium entry. Endocrinology 137(2), 587-594. Watanabe, Y., Nunokawa, A., Someya, T., 2013. Association of the BDNF C270T polymorphism with schizophrenia: updated meta-analysis. Psychiatry Clin Neurosci 67(2), 123-125. Watson, S., Gallagher, P., Ritchie, J.C., Ferrier, I.N., Young, A.H., 2004. Hypothalamic-pituitaryadrenal axis function in patients with bipolar disorder. Br J Psychiatry 184, 496-502. Weiss, N., Heydrick, S.J., Postea, O., Keller, C., Keaney, J.F., Jr., Loscalzo, J., 2003. Influence of hyperhomocysteinemia on the cellular redox state--impact on homocysteine-induced endothelial dysfunction. Clin Chem Lab Med 41(11), 1455-1461. Welters, H.J., Kulkarni, R.N., 2008. Wnt signaling: relevance to beta-cell biology and diabetes. Trends in endocrinology and metabolism: TEM 19(10), 349-355. Wender, R., Brown, A.M., Fern, R., Swanson, R.A., Farrell, K., Ransom, B.R., 2000. Astrocytic glycogen influences axon function and survival during glucose deprivation in central white matter. J Neurosci 20(18), 6804-6810. Wendt, A., Speidel, D., Danielsson, A., Esguerra, J.L., Bogen, I.L., Walaas, S.I., Salehi, A., Eliasson, L., 2012. Synapsins I and II are not required for insulin secretion from mouse pancreatic beta-cells. Endocrinology 153(5), 2112-2119. Werner, H., LeRoith, D., 2014. Insulin and insulin-like growth factor receptors in the brain: physiological and pathological aspects. Eur Neuropsychopharmacol 24(12), 1947-1953. Winham, S.J., Cuellar-Barboza, A.B., Oliveros, A., McElroy, S.L., Crow, S., Colby, C., Choi, D.S., Chauhan, M., Frye, M., Biernacka, J.M., 2014. Genome-wide association study of bipolar disorder accounting for effect of body mass index identifies a new risk allele in TCF7L2. Mol Psychiatry 19(9), 1010-1016. Wisniewska, M.B., Misztal, K., Michowski, W., Szczot, M., Purta, E., Lesniak, W., Klejman, M.E., Dabrowski, M., Filipkowski, R.K., Nagalski, A., Mozrzymas, J.W., Kuznicki, J., 2010. LEF1/betacatenin complex regulates transcription of the Cav3.1 calcium channel gene (Cacna1g) in thalamic neurons of the adult brain. J Neurosci 30(14), 4957-4969.

36

Wisniewska, M.B., Nagalski, A., Dabrowski, M., Misztal, K., Kuznicki, J., 2012. Novel beta-catenin target genes identified in thalamic neurons encode modulators of neuronal excitability. BMC Genomics 13, 635. Wong, M.L., Licinio, J., 2001. Research and treatment approaches to depression. Nat Rev Neurosci 2(5), 343-351. Yang, J., Liu, X., Liu, X., Wang, L., Lv, H., Yu, J., Xun, Z., Yang, G., 2012. Abnormality of glycometabolism related factors in non-psychotic offspring of schizophrenic patients. Psychiatry Res 198(2), 183-186. Zhang, Z., Lovato, J., Battapady, H., Davatzikos, C., Gerstein, H.C., Ismail-Beigi, F., Launer, L.J., Murray, A., Punthakee, Z., Tirado, A.A., Williamson, J., Bryan, R.N., Miller, M.E., 2014. Effect of hypoglycemia on brain structure in people with type 2 diabetes: epidemiological analysis of the ACCORD-MIND MRI trial. Diabetes care 37(12), 3279-3285. Zhou, Y., Park, S.Y., Su, J., Bailey, K., Ottosson-Laakso, E., Shcherbina, L., Oskolkov, N., Zhang, E., Thevenin, T., Fadista, J., Bennet, H., Vikman, P., Wierup, N., Fex, M., Rung, J., Wollheim, C., Nobrega, M., Renstrom, E., Groop, L., Hansson, O., 2014. TCF7L2 is a master regulator of insulin production and processing. Human molecular genetics 23(24), 6419-6431.

FIGURE LEGENDS Fig. 1. Intersection of susceptibility genes for schizophrenia (SCZ), type 2 diabetes (T2D), and serious mental illnesses (SMIs) revealed in genome-wide association studies (GWASs). Numbers in circles indicate the number of genes associated with the risk of a particular disorder.

37

38

Fig. 2. Potential roles of insulin in the brain.

39