Entry - %125480 - MAJOR AFFECTIVE DISORDER 1; MAFD1 - OMIM
% 125480

MAJOR AFFECTIVE DISORDER 1; MAFD1


Alternative titles; symbols

MANIC-DEPRESSIVE PSYCHOSIS; MD1
BIPOLAR AFFECTIVE DISORDER; BPAD
MANIC-DEPRESSIVE PSYCHOSIS, AUTOSOMAL


Cytogenetic location: 18p     Genomic coordinates (GRCh38): 18:1-18,500,001


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
18p {Major affective disorder 1} 125480 AD 2
Clinical Synopsis
 

INHERITANCE
- Autosomal dominant
NEUROLOGIC
Behavioral Psychiatric Manifestations
- Depression with mania (bipolar 1)
- Depression with hypomania (bipolar 2)

TEXT

Bipolar affective disorder is a genetically heterogeneous complex trait. One susceptibility locus for bipolar disorder, MAFD1, has been mapped to chromosome 18p.

Other mapped loci include MAFD2 (309200) on chromosome Xq28, MAFD3 (609633) on chromosome 21q22, MAFD4 (611247) on chromosome 16p12, MAFD7 (612371) on chromosome 22q12, MAFD8 (612357) on chromosome 10q21, and MAFD9 (612372) on chromosome 12p13.

Epistatic interaction has been postulated for loci designated MAFD5 (611535) on chromosome 2q22-q24 and MAFD6 (611536) on chromosome 6q23-q24.


Clinical Features

Depressive disorders represent a prevalent (1 to 2%) and major illness characterized by episodes of dysphoria that are associated with somatic symptoms. It may have a manic-depressive (bipolar) or purely depressive (unipolar) course. The most characteristic features of bipolar affective disorder (BPAD) are episodes of mania (bipolar I, BP I) or hypomania (bipolar II, BP II) interspersed with periods of depression (Goodwin and Jamison, 1990). If untreated, manic-depressive illness is associated with a suicide rate of approximately 20%.


Biochemical Features

Wright et al. (1984) studied binding of radioiodine-labeled hydroxybenzylpindolol to beta-adrenoceptors in lymphoblastoid cell lines from members of 5 families affected by manic-depressive disorder. Binding was reduced to less than half of control values in cell lines from 4 of 6 patients with manic-depressive disorder and only 1 of 18 unaffected relatives or controls. All the cell lines with reduced binding came from 3 families; members of 2 remaining families showed normal binding. The findings were interpreted as indicating genetic heterogeneity in manic-depressive disorder and a role played by a beta-adrenoceptor defect in genetic susceptibility to the disorder in some cases.

O'Reilly et al. (1994) presented a 2-generation family in which 8 members met DSM-III-R criteria for major depression. Four of the affected individuals failed to respond to standard therapeutic doses of tricyclic and new generation antidepressants but subsequently responded to the monoamine oxidase inhibitor tranylcypromine. O'Reilly et al. (1994) suggested that families demonstrating preferential response to a particular psychotropic drug may be a more homogeneous group in which to perform linkage analysis. Grof et al. (1994) also suggested that a helpful criterion for selecting more homogeneous groups of patients for linkage analysis would be the presence of a response to specific treatment. To that end, they studied 121 probands with primary affective disorders and 903 first-degree relatives and spouses. Seventy-one probands responded to lithium treatment and 50 were nonresponders. Study of the first-degree relatives of lithium responders revealed that 3.8% had bipolar disorder, whereas none of the relatives of the nonresponders were so affected. Schizophrenia was more common in the families of nonresponders (2.4% vs 0.3%).

In studies of the effect of lithium on Xenopus morphogenesis, Klein and Melton (1996) determined that lithium acts through inhibition of glycogen synthase kinase 3-beta (GSK3B; 605004) and not through inhibition of inositol monophosphatase (602064). They suggested that their observations may provide insights into the pathogenesis and treatment of bipolar disorder.

Using positron emission tomographic (PET) images of cerebral blood flow and rate of glucose metabolism to measure brain activity, Drevets et al. (1997) localized an area of abnormally decreased activity in the prefrontal cortex ventral to the genu of the corpus callosum in both familial bipolar depression patients and familial unipolar depression patients. The decrement in activity was at least partly explained by a corresponding reduction in cortical volume, as demonstrated by magnetic resonance imaging (MRI). This region had previously been implicated in the mediation of emotional and autonomic responses to socially significant or provocative stimuli, and in the modulation of the neurotransmitter systems targeted by antidepressant drugs.

Williams et al. (2002) demonstrated that lithium, carbamazepine, and valproic acid, all drugs used to treat bipolar affective disorder, inhibit the collapse of sensory neuron growth cones and increase growth cone area. These effects do not depend on glycogen synthase kinase-3 (see 606784) or histone deacetylase (see 601241) inhibition. Inositol, however, reverses the effects of the drugs on growth cones, thus implicating inositol depletion in their action. Moreover, the development of Dictyostelium is sensitive to lithium and to valproic acid, but resistance to both is conferred by deletion of the gene that codes for prolyl oligopeptidase (600400), which also regulates inositol metabolism. Inhibitors of prolyl oligopeptidase reversed the effects of all 3 drugs on sensory neuron growth cone area and collapse. Williams et al. (2002) concluded that their results suggest a molecular basis for both bipolar affective disorder and its treatment.


Inheritance

The role of genetic factors in bipolar disorder is indicated by concordance in monozygotic and dizygotic twins, respectively, of 57% and 14%, and the correlation between adopted persons and their biologic relatives (Cadoret, 1978).

Brent and Mann (2005, 2006) noted that in studies of adopted children and twins, familial concordance for suicidal behavior is explained by both genetic and environmental factors. Suicidal behavior that begins before 25 years of age is highly familial, and having a greater number of affected family members is associated with an earlier age.

Craddock and Sklar (2013) noted that strong evidence exists for a polygenic contribution to risk of bipolar disorder (i.e., many risk alleles of small effect). They stated that most cases of bipolar disorder involve the interplay of several genes or more complex genetic mechanisms, together with the effects of nongenetic (environmental) risk factors and stochastic factors.

Anticipation

In 34 families unilineal for bipolar affective disorder, McInnis et al. (1993) compared age of onset and disease severity between 2 generations. They found that the second generation experienced onset 8.9 to 13.5 years earlier and illness 1.8 to 3.4 times more severe than did the first generation. They concluded that genetic anticipation was demonstrated and they suggested that genes with expanding trinucleotide repeats might be involved in the genetic etiology of the disorder.

Gelernter (1995) reviewed the genetics of BPAD and other behavioral disorders and discussed specifically the difficulties associated with possible anticipation in bipolar affective disorder as discussed by McInnis et al. (1993). McInnis (1996) traced the history of anticipation. He suggested that it has its roots in the social chaos of the French Revolution, as well as in the medical observations of the French 'alienist' (an early term for psychiatrist) (Morel, 1857) and in the theories of atavism advocated by the Italian psychiatrist of the late 19th century (Lombroso, 1911). As to the validity of anticipation in bipolar disorder and schizophrenia, McInnis (1996) quoted Hodge and Wickramaratne (1995) as concluding that 'in psychiatric disorders, bias of ascertainment is pervasive and that there is no simple way to circumvent it.'

Parent-of-Origin Effect

McMahon et al. (1995) tested for parent-of-origin effect on the transmission of bipolar affective disorder, which might reflect either imprinting or mitochondrial inheritance. They examined the frequency and risk of BPAD among relatives in a sample of 31 families ascertained through treated probands with BPAD and selected for the presence of affected phenotypes in only 1 parental lineage. They observed a higher than expected frequency of affected mothers (p less than 0.04), a 2.3- to 2.8-fold increased risk of illness for maternal relatives (P less than 0.006), and a 1.3- to 2.5-fold increased risk of illness for the offspring of affected mothers (p less than 0.017). In 7 large pedigrees, fathers repeatedly failed to transmit the affected phenotype to daughters or sons. Taken together, these findings were interpreted as indicating a maternal effect in the transmission in BPAD susceptibility and suggested that molecular studies of mtDNA and imprinted DNA are warranted in patients with BPAD.

A number of family studies had reported increased morbid risk to the mothers, relative to the fathers, of probands with bipolar affective disorder. An excess of mother-offspring pairs had also been reported. These observations suggested that bipolar affective disorder may be caused by mitochondrial DNA mutations. Kirk et al. (1999) sequenced the mitochondrial genome in 25 bipolar patients with family histories of psychiatric disorder that suggested matrilineal inheritance. No polymorphism identified more than once in this sequencing showed any significant association with bipolar affective disorder in association studies using 94 cases and 94 controls. To determine whether their sample of patients showed evidence of selection against the maternal lineage, Kirk et al. (1999) determined genetic distances between all possible pairwise comparisons within the bipolar and control groups, based on multilocus mitochondrial polymorphism haplotypes. These analyses revealed fewer closely related haplotypes in the bipolar group than in the matched control group, suggesting selection against maternal lineages in this disease. Such selection was considered compatible with recurrent mitochondrial mutations, which are associated with slightly decreased fitness.


Mapping

Linkage to Chromosome 18 (MAFD1)

In the course of a systematic genomic survey, Berrettini et al. (1994) examined 22 manic-depressive (bipolar) families for linkage to 11 chromosome 18 pericentromeric marker loci, under dominant and recessive models. Overall lod score analysis for the pedigrees was not significant under either model, but several families yielded scores consistent with linkage under dominant or recessive models. Affected sib pair analysis of these data yielded evidence for linkage (p less than 0.001) with D18S21. Affected pedigree member analysis also suggested linkage. The results were interpreted as suggesting a susceptibility gene in the pericentromeric region of chromosome 18, with a complex mode of inheritance.

Pursuant to the report by Berrettini et al. (1994) of a BPAD susceptibility locus on chromosome 18 and the report of a parent-of-origin effect by McMahon et al. (1995), Stine et al. (1995) undertook a linkage study in 28 nuclear families selected for apparent unilineal transmission of the BPAD phenotype. They used 31 polymorphic markers spanning chromosome 18. The study was distinguished by relatively small, densely affected families with apparent unilineal transmission and direct clinical evaluation of family members by psychiatrists. Evidence for linkage was tested with affected-sib-pair and lod score methods under 2 definitions of the affected phenotype. The affected-sib-pair analyses indicated excess allele sharing for markers on 18p within the region reported previously. The greatest sharing was at D18S37. In addition, excess sharing of the paternally, but not maternally, transmitted alleles was observed at 3 markers on 18q. The evidence for linkage to loci on both 18p and 18q was strongest in the 11 paternal pedigrees, i.e., those in which the father or one of the father's sibs was affected. The results were interpreted by Stine et al. (1995) as providing further support for linkage of BPAD to chromosome 18.

Prompted by the report by Berrettini et al. (1994) of a bipolar susceptibility locus in the region of the centromere of chromosome 18, Pauls et al. (1995) studied markers from this region in Old Order Amish families. Although linkage findings were replicated in 1 previously studied Amish pedigree containing 4 affected individuals, linkage to this region was excluded in the larger sample. Pauls et al. (1995) concluded that if a susceptibility locus for bipolar disorder is located in this region of chromosome 18, it is of minor significance among the Amish.

In a genetically isolated population of the central valley of Costa Rica (CVCR), Freimer et al. (1996) undertook a linkage study of severe bipolar disorder in 2 pedigrees. The 2.6 million residents of the CVCR are descended mainly from a small group of Spanish and Amerindian founders who lived in the 16th and 17th centuries; by the beginning of the 18th century, the CVCR had a single population that then grew rapidly, without subsequent immigration, for almost 200 years (Escamilla et al., 1996). Freimer et al. (1996) found strongest evidence for a specific locus on chromosome 18q22-q23 where 7 of 16 markers yielded peak lod scores over 1.0. This localization was supported by marker haplotypes shared by 23 of 26 affected individuals studied. As a continuation of that study, McInnes et al. (1996) performed a complete genome screen for genes predisposing to severe bipolar disorder. They considered as affected only individuals with bipolar mood disorder and screened the genome for linkage with 473 microsatellite markers. They used a model for linkage analysis that incorporated a high phenocopy rate and a conservative estimate of penetrance. They suggested on the basis of their results that 18q, 18p, and 11p deserve further study; in these regions suggestive lod scores were observed for 2 or more contiguous markers. Isolated lod scores that exceeded the threshold in 1 or both families studied also occurred on 10 other chromosomes. Additional linkage studies on 2 extended BP I pedigrees from the CVCR implicated a candidate region on 18p11.3 (Escamilla et al., 1999).

Knowles et al. (1998) could find no evidence for significant linkage between bipolar affective disorder and chromosome 18 pericentromeric markers in a large series of multiplex extended pedigrees. This was one of the largest samples reported to date: 1,013 genotyped individuals in 53 unilineal multiplex pedigrees. Ten highly polymorphic markers and a range of parametric and nonparametric analyses were used. Not only was there no evidence for linkage, but there was also no evidence for significant parent-of-origin effect.

McInnes et al. (2001) further investigated the 18p11.3 region by creating a physical map and developing 4 new microsatellite and 26 single-nucleotide polymorphism (SNP) markers for typing in the Costa Rican pedigree and population samples. The results of fine-scale association analyses in the population sample, as well as evaluation of haplotypes in 1 of the large pedigrees, suggested a candidate region containing 6 genes but also highlighted the complexities of linkage disequilibrium mapping of common disorders.

To clarify the issue of genetic linkage between bipolar affective disorder and 18q, McMahon et al. (2001) analyzed the relationship between clinical features and allele sharing. Relatives ascertained through a proband who had BP I disorder were interviewed by a psychiatrist, assigned an all-sources diagnosis, and genotyped with 32 markers on 18q21-q23. The authors found that paternal allele sharing on 18q21 was significantly associated with a diagnostic subtype, and was greatest in pairs where both sibs had BP II. Paternal allele sharing across 18q21-q23 was also significantly greater in families with at least 1 sib pair in which both had BP II. In these families, multipoint affected sib-pair linkage analysis produced a peak paternal lod score of 4.67 versus 1.53 in all families. Thus, affected sib pairs with BP II discriminated between families who showed evidence of linkage to 18q and families who did not. Families with a BP II sib pair produced an increased lod score and improved linkage resolution. These findings strengthened the evidence of genetic linkage between BPAD and 18q, and provided preliminary support for BP II as a genetically valid subtype of BPAD. Bipolar type II disorder is characterized by hypomania that is so brief or so slight as to cause no significant problems in functioning. BP I, on the other hand, is the diagnosis attached to anyone with a significantly problematic manic state (extreme symptoms of grandiosity, poor social judgment or functional impairment due to distractibility at work, etc.).

Linkage to Chromosome 2q22-q24

See MAFD5 (611535) for a discussion of linkage of susceptibility to bipolar disorder to chromosome 2q22-q24.

Linkage to Chromosome 3p

Etain et al. (2006) conducted a genomewide search with 384 microsatellite markers using nonparametric linkage (NPL) analysis in 87 sib pairs ascertained as part of the European Collaborative Study of Early Onset Bipolar Affective Disorder. Early-onset patients (age at onset of 21 years or below) were studied because age of onset may help to define homogeneous bipolar affective disorder subtypes. The 3p14 region showed the most significant linkage in the first phase of analysis with an NPL score of 3.51. Additional linkage analysis with increased marker density revealed an NPL score of 3.83 at chromosome 3p14.

Linkage to Chromosome 4

Blackwood et al. (1996) carried out a linkage study in Scotland in 12 bipolar families. In a single family, a genome search using 193 markers indicated linkage on 4p where D4S394 generated a 2-point lod score of 4.1 under a dominant model of inheritance. With 3-point analyses using neighboring markers, they obtained a maximum lod score of 4.8. (Eleven other bipolar families were typed using D4S394 and in all families combined there was evidence of linkage with heterogeneity with a maximum 2-point lod score of 4.1 (theta = 0.0; alpha = 0.35).)

Ginns et al. (1998) reported a different approach to linkage study of BPAD in the Old Order Amish. To determine whether there could be protective alleles that prevent or reduce the risk of developing BPAD, similar to what is observed in other genetic disorders, they used 'mental health wellness' (absence of any psychiatric disorder) as the phenotype in their genomewide linkage scan of several large multigeneration Old Order Amish pedigrees exhibiting a high incidence of BPAD. They found strong evidence for linkage of mental health wellness to a locus on 4p, designated MHW1 (603663), at D4S2949; maximum nonparametric linkage score = 4.05, p = 5.22 x 10(-4). They also found suggestive evidence for a locus on 4q, designated MHW2 (603664), at D4S397; maximum nonparametric linkage score = 3.29, p = 2.57 x 10(-3). Findings were consistent with the hypothesis that certain alleles can prevent or modify the clinical manifestations of BPAD and perhaps other related affective disorders.

Ekholm et al. (2003) performed a genomewide scan for susceptibility loci in bipolar disorder in 41 Finnish families with at least 2 affected sibs. They identified a distinct locus on 16p12 (see MAFD4, 611247) and observed 3 additional loci with a 2-point lod score greater than 2.0, at markers on 4q32, 12q23, and Xq25. After fine mapping these chromosomal regions and genotyping additional family members, 4q32 provided significant evidence of linkage for the 3-point analyses (maximum lod = 3.6 between D4S3049 and D4S1629).

Linkage to Chromosome 5

Garner et al. (2001) used an algorithm that permitted nonparametric linkage analysis of large, complex pedigrees with multiple inbreeding loops to reanalyze the genome-screen data from the Costa Rican kindred segregating severe bipolar disorder (Freimer et al., 1996; McInnes et al., 1996). The results were consistent with previous linkage findings on chromosome 18 and also suggested a novel locus on chromosome 5 that was not identified using traditional linkage analysis.

Hong et al. (2004) performed linkage analysis using 74 individuals from the Costa Rican pedigree and found evidence for a 3.2-Mb region between markers D5S1480 and D5S2090 on chromosome 5q31-q33. The authors suggested that conflicting haplotype data reflected incomplete penetrance, phenocopies, or locus/allelic heterogeneity. The authors noted that Freimer et al. (1996) described a conserved haplotype on 18q22-q23 in the same kindred. Hong et al. (2004) found that 12 of 20 affected individuals shared both haplotypes, suggesting that both loci are important in conferring disease risk.

Coon et al. (1993) carried out an extensive linkage analysis in 8 moderate-sized families with manic-depression. When autosomal dominant inheritance was assumed, 273 DNA markers gave lod scores less than -2.0 at theta = 0.0, 174 DNA loci produced lod scores less than -2.0 at theta = 0.05, and 4 DNA marker loci yielded lod scores greater than 1. Of the markers giving lod scores greater than 1, only D5S62 continued to show evidence for linkage when the affected-pedigree-member method was used. D5S62 maps to distal 5q, a region containing neurotransmitter receptor genes for dopamine (e.g., 126449), gamma-aminobutyric acid (e.g., 137160, 137164), glutamate (e.g., 138248), and norepinephrine (e.g., 109690, 104219, 104220).

Linkage to Chromosome 6p

Smeraldi et al. (1978) first suggested linkage between HLA on chromosome 6p21.3 and affective disorders on the basis of the finding that pairs of affected sibs shared HLA haplotypes more often than would be predicted by chance. Weitkamp et al. (1981) likewise found evidence of a susceptibility gene or genes linked to HLA. Neither group subdivided the depressive disorders into bipolar and unipolar subtypes. Stronger evidence of linkage might be found in 1 subtype, or it may turn out that both are linked to HLA, suggesting that they are different forms of the same illness. One of Weitkamp's study families was that reported earlier by Pardue (1975)--in fact, Pardue's own kindred (Wingerson, 1982). Weitkamp et al. (1981) found that HLA haplotype identity in pairs of affected sibs and in pairs of unaffected older sibs deviated markedly from expected (p less than 0.005). Perhaps surprisingly, no increase in HLA haplotype identity was found in sibships with more than 2 affected members. When parents had a difference in load of genes for susceptibility (as estimated by the occurrence of affective illness in themselves and their relatives), HLA haplotypes were randomly transmitted to unaffected or affected children from the affected, 'high-load' parent, but not randomly from the unaffected, 'low-load' parent (p less than 0.001), suggesting a recessive effect, i.e., greater chance of illness in homozygotes.

Stancer et al. (1988) published data apparently confirming the relationship between HLA and manic depression. When combined with their previous data, the total number of families analyzed was 117. As in the previous study, the increase in HLA haplotype sharing over random expectation was greater if 'high-load' sibships, i.e., sibships with 3 or more affected sibs, were omitted from the analysis. Weitkamp (1981, 1983) suggested that the extent of HLA haplotype sharing among affected sib pairs should decrease as the number of parental HLA haplotypes containing susceptibility genes increases from 1 to 4. Thus, he reasoned that there may actually be less HLA haplotype sharing among sibs when the parents have maximum genetic susceptibility ('high load') compared with families in which the genetic susceptibility that could be contributed by either parent is limited to the genes in 1 of the 2 HLA haplotypes in that parent. If an increased number or variety of affective disorder susceptibility genes in a person results in a greater probability of illness, then nuclear families with a higher proportion of affected family members are likely to have a greater number or variety of affective disorder susceptibility genes than families with a low proportion of affected members. Weitkamp and Stancer (1989) suggested that the HLA effect may be greater in unipolar than in bipolar disorders and more apparent in families with few affected members than in 'high-load' families.

Schulze et al. (2004) extended the study of Dick et al. (2003) to test for robustness of the linkage to differing analysis methods, genotyping error, and gender-specific maps; for parent-of-origin effects; and for interaction with markers within the schizophrenia linkage region on chromosome 6p (see SCZD3; 600511). Members of 245 families ascertained through a sib pair affected with bipolar I or schizoaffective-bipolar disorder were genotyped with 18 markers spanning chromosome 6, and nonparametric linkage analysis was performed. Linkage to 6q was robust to analysis methods, gender-specific map differences, and genotyping error. The locus conferred a 1.4-fold increased risk. Affected sibs shared the maternal more often than paternal chromosome (p = 0.006), which could reflect a maternal parent-of-origin effect. There was a positive correlation between family-specific linkage scores on 6q and those on 6p22.2 (p less than 0.0001). Linkage analysis for each locus conditioned on evidence of linkage to the other increased the evidence for linkage at both loci (p less than 0.0005). Lod scores increased from 2.26 to 5.42 on 6q and from 0.35 to 2.26 on 6p22.2. The results supported linkage of bipolar disorder to 6q, revealed a maternal parent-of-origin effect, and demonstrated an interaction of this locus with a locus on chromosome 6p22.2 linked to schizophrenia.

Linkage to Chromosome 6q22

Middleton et al. (2004) performed a linkage analysis on 25 extended multiplex Portuguese families, including 12 families previously reported by Pato et al. (2004), segregating for bipolar disorder using a high-density SNP genotyping assay with a 0.21-Mb intermarker spacing. The analysis revealed genomewide significance with a maximum NPL of 4.20 and a maximum lod score of 3.56 at 6q22 (125.8 Mb).

Linkage to Chromosome 8

Ophoff et al. (2002) performed a genomewide association study of severe bipolar disorder in the patients from the central valley of Costa Rica. They observed LD with severe bipolar disorder on several chromosomes; the most striking results were in proximal 8p, a region that had previously shown linkage to schizophrenia. Ophoff et al. (2002) suggested that this region could be important for severe psychiatric disorders rather than for a specific phenotype.

Cichon et al. (2001) conducted a complete genome screen with 382 markers in a sample of 75 BPAD families of German, Israeli, and Italian origin. Parametric and nonparametric linkage analysis was performed. The highest 2-point lod score was obtained on 8q24 (D8S514; lod score = 3.62), and the authors confirmed a putative locus on 10q25-q26 (D10S217; lod score = 2.86). By analyzing the autosomal genotype data, putative paternally imprinted loci were identified in chromosomal regions 2p24-p21 and 2q31-q32; maternally imprinted susceptibility genes may be located on 14q32 and 16q21-q23.

Park et al. (2004) genotyped 373 individuals from 40 extended pedigrees with high density bipolar disorder and found evidence for significant linkage for psychotic bipolar disorder (genomewide p less than 0.05) to chromosomes 9q31 (lod = 3.55) and 8p21 (lod = 3.46). Nine other sites obtained lod scores supportive of linkage. The highest lod scores occurred in the subgroup of families with the largest concentration of psychotic individuals. Seven of the loci identified in this study had previously been implicated in schizophrenia, suggesting that psychosis is a potentially useful phenotype in bipolar disorder for genetic studies.

Linkage to Chromosome 9

Sherrington et al. (1994) performed linkage analysis on 5 multigenerational families with bipolar and unipolar affective disorder, using highly polymorphic microsatellite markers from the ABO-AK1-ORM region at 9q34. The dopamine beta-hydroxylase locus (223360) is also at 9q34 and was considered to be a candidate gene. Their analyses provided strong evidence against a major susceptibility allele in this region, in contradistinction to the findings of Hill et al. (1988), Tanna et al. (1989), and Wilson et al. (1989, 1991).

Venken et al. (2005) conducted a genomewide scan to identify susceptibility loci for affective spectrum disorder (bipolar disorder and recurrent unipolar depression) in 9 families from an isolated population in Vasterbotten in northern Sweden. A region on chromosome 9q showed the highest 2-point and multipoint lod scores. A common ancestral haplotype was inherited by 18 of 21 patients from 3 families linked to 9q, which reduced the candidate region to 1.6 Mb on 9q31-q33. Further analysis identified the shared haplotype in 4.2% of 182 unrelated patients with bipolar disorder from the Vasterbotten isolate, but not in 182 control individuals. Venken et al. (2005) concluded that a susceptibility locus for affective disorder is located on chromosome 9q31-q33.

Park et al. (2004) genotyped 373 individuals from 40 extended pedigrees with high density bipolar disorder and found evidence for significant linkage for psychotic bipolar disorder (genomewide p less than 0.05) to chromosomes 9q31 (lod = 3.55) and 8p21 (lod = 3.46). Nine other sites obtained lod scores supportive of linkage. The highest lod scores occurred in the subgroup of families with the largest concentration of psychotic individuals. Seven of the loci identified in this study had previously been implicated in schizophrenia, suggesting that psychosis is a potentially useful phenotype in bipolar disorder for genetic studies.

Linkage to Chromosome 10q21

Ferreira et al. (2008) tested 1.8 million variants in 4,387 cases of bipolar disorder and 6,209 controls from 3 independent samples and identified a region of strong association with SNP rs10994336 in the ankyrin G gene (ANK3; 600465) on chromosome 10q21, with a p value of 9.1 x 10(-9). See MAFD8 (612357).

Linkage to Chromosome 11

A form of manic-depressive disorder in the Old Order Amish of Lancaster County, Pennsylvania, was thought by Egeland et al. (1987) to be tightly linked to INS (176730) and HRAS1 (190020). In linkage studies using RFLPs related to these genes on the tip of 11p, the maximum lod score was 4.5 at theta = 0.0. Of interest is the description by Joffe et al. (1986) of cosegregation of thalassemia and affective disorder in a non-Amish pedigree. Egeland et al. (1987) suggested that the tyrosine hydroxylase (TH; 191290) gene, which maps to 11p, should be considered as a candidate gene because this enzyme catalyzes an important step in the dopamine synthesis pathway. Gill et al. (1988) ruled out tight linkage between manic-depressive psychosis and the 11p markers HRAS1 and INS, however. Two other groups failed to find linkage of 11p markers to manic-depressive illness (Neiswanger et al., 1990). From another extension of the study of the original Amish pedigree, Pauls et al. (1991) likewise excluded linkage to 11p markers. Pakstis et al. (1991) found no evidence of linkage after screening 185 marker loci in the Old Order Amish. They estimated that roughly 23% of the autosomal genome had been excluded. Law et al. (1992) determined the INS and HRAS1 genotypes of 81 persons in this pedigree and excluded that region of chromosome 11 as the site of the gene, which they symbolized BAD (for bipolar affective disorder).

In 5 Icelandic pedigrees, Holmes et al. (1991) could find no evidence of linkage of manic depression to the dopamine D2 receptor (DRD2; 126450) or other markers in its vicinity on 11q.

Linkage to Chromosome 12p13

Ferreira et al. (2008) tested 1.8 million variants in 4,387 cases of bipolar disorder and 6,209 controls from 3 independent samples and identified association with SNP rs1006737 in the CACNA1C gene (114205) on chromosome 12p13, with a p value of 7.0 x 10(-8). See MAFD9 (612372).

Linkage to Chromosome 12q

By linkage analysis in 2 Danish families with bipolar affective disorder, Ewald et al. (1998) found that the microsatellite marker D12S1639 gave a significant lod score of 3.37. Earlier, Craddock et al. (1994) had suggested linkage between affective disorder and Darier disease (124200), which maps to 12q23-q24.1. Linkage results from independent studies in Canadian families (Morissette et al., 1999) supported the existence of a susceptibility locus on 12q23-q24. To take advantage of isolated populations for genetic mapping and disease gene identification, Degn et al. (2001) investigated a possible chromosomal segment shared among distantly related patients with bipolar affective disorder in the Faroe Islands, using 17 microsatellite markers covering 24 cM in the 12q24 region. The region of most interest contained the primary region suggested by the previously reported haplotypes in the 2 Danish families studied by Ewald et al. (1998).

Ewald et al. (2002) reported a genomewide scan for risk genes involved in bipolar disorder in 2 Danish Caucasian families with affected members in several generations. Ewald et al. (2002) used 613 microsatellite markers in a 2-stage approach. Linkage was obtained at 12q24.3 (D12S1639) with a 2-point parametric lod score of 3.42 (empirical P-value 0.00004, genomewide P-value 0.0417) in both families tested. The multipoint parametric lod score at D12S1639 was 3.63 (genomewide P-value 0.0265). At 1p22-p21 (D1S216), a parametric, affecteds-only 2-point lod score of 2.75 (empirical P-value 0.0002, genomewide P-value 0.1622) was found. A 3-point lod score of 2.98 (genomewide p value = 0.1022) was found at D1S216, and a multipoint nonparametric analysis yielded a maximum nonparametric linkage (NPL)-all score of 17.60 (p value = 0.00079) at D1S216.

In 2 cohorts of patients with bipolar affective disorder from Germany and Russia totaling 883 patients and 1,300 controls, Cichon et al. (2008) observed an association between disease and the minor alleles of 3 SNPs in haplotype 1 of the TPH2 gene on chromosome 12q21, (rs11178997, rs11178998, and rs7954758; odds ratio of 1.6, p value of 0.00073). Haplotype 1 covers part of the 5-prime regulatory region and exons 1 and 2 of the TPH2 gene. Cichon et al. (2008) also observed an association between bipolar disorder and a nonsynonymous SNP in the TPH2 gene (P206S; 607478.0003).

Linkage to Chromosome 16

See MAFD4 (611247) for a discussion of linkage of susceptibility to bipolar disorder to chromosome 16p12.

Linkage to Chromosome 17

Dick et al. (2003) performed genomewide linkage analyses on 1,152 individuals from a new sample of 250 families segregating for bipolar disorder and related affective illnesses, ascertained at 10 sites in the United States through a proband with BP I affective disorder and a sib with BP I or schizoaffective disorder, bipolar type. Suggestive evidence for linkage was found on chromosome 17q (peak maximum lod score = 2.4) at marker D17S928, and on 6q (peak maximum lod score = 2.2) near marker D6S1021. Suggestive evidence of linkage was observed in 3 other regions, on chromosomes 2p, 3q, and 8q. This study, based on a linkage sample for bipolar disorder larger than any previously analyzed, indicated that several genes contribute to bipolar disorder.

Linkage to Chromosome 20

In 9 Australian pedigrees, Le et al. (1994) excluded close linkage of bipolar disorder to the gene encoding the alpha subunit of the stimulatory form of G protein (139320), previously mapped to chromosomal region 20q13.2.

Radhakrishna et al. (2001) studied a large Turkish pedigree segregating apparently autosomal dominant BPAD, which contained 13 affected individuals. The age of onset ranged from 15 to 40 years with a mean age of 25 years. The phenotypes consisted of recurrent manic and major depressive episodes, including suicide attempts. There was usually full remission with lithium treatment. A genotyping of 230 highly informative polymorphic markers throughout the genome and subsequent linkage analysis using a dominant mode of inheritance showed strong evidence for a BPAD susceptibility locus on chromosome 20p11.2-q11.2. The highest 2-point lod score of 4.34 (theta = 0.0) was obtained with markers D20S604, D20S470, D20S836, and D20S838 (100% penetrance). Haplotype analysis using informative recombinants enabled the mapping of the BPAD locus in this family between markers D20S186 and D20S109 in a region of approximately 42 cM. The authors noted that the chromosome 20 BPAD susceptibility locus had not been identified in previous studies of common 'polygenic' small pedigrees, which could be explained by an absence of common deleterious mutations of the chromosome 20 BPAD locus in those pedigrees and/or by the presence of a severe mutation in the Turkish pedigree that by itself confers susceptibility to BPAD.

Linkage to Chromosome 21q22

In a preliminary genome screen of 47 bipolar disorder families, Straub et al. (1994) detected one in which a lod score of 3.41 was demonstrated for linkage with the PFKL (171860) locus on 21q22.3. Largely positive lod scores were obtained also with 14 other markers in 21q22.3 in this family. In a linkage analysis with an 'affecteds-only' method, Aita et al. (1999) found linkage to the 21q22 region, corroborating the findings of the earlier study by the same group (Straub et al., 1994).

Following up on the work of Straub et al. (1994) suggesting a susceptibility locus for bipolar affective disorder on the long arm of chromosome 21, Smyth et al. (1997) studied 23 multiply affected pedigrees collected from Iceland and the U.K., using the markers PFKL, D21S171, and D21S49. Positive lod scores were obtained with 3 Icelandic families. Affected sib pair analysis demonstrated increased allele sharing. The same set of pedigrees had previously been typed for a tyrosine hydroxylase gene (TH; 191290) polymorphism at 11p15 and had shown some evidence for linkage. When information from TH and the 21q markers was combined in a 2-locus admixture analysis, an overall admixture lod of 3.87 was obtained using the bipolar affection model. Thus the data of Smyth et al. (1997) were compatible with the hypothesis that a locus at or near TH influences susceptibility of some pedigrees, while a locus near D21S171 is active in others.

Age at onset (AAO) is a potential clinical marker of genetic heterogeneity in BP (Bellivier et al., 2001). Rates of comorbidity and clinical indicators of severity (e.g., suicide attempt) vary across different AAO subgroups, and AAO subgroups aggregate in families such that affected relatives typically have similar AAOs. Therefore, Lin et al. (2005) sought to incorporate AAO as a covariate in linkage analysis of BP using 2 different methods in genomewide scans of 150 multiplex pedigrees with 874 individuals. The LODPAL analysis identified 2 loci: one on 21q22.13 (MAFD3; 609633) and the other on 18p11.2 (MAFD1) for early onset (AAO = 21 years or younger) and later onset (AAO = older than 21 years), respectively. The finding on 21q22.13 was significant at the chromosome-wide level, even after correction for multiple testing. Moreover, a similar finding was observed in an independent sample of 65 pedigrees (lod = 2.88). The finding on 18p11.2 was only nominally significant and was not observed in the independent sample. However, 18p11.2 emerged as one of the strongest regions in the ordered-subset analysis (OSA) with a lod of 2.92, in which it was the only finding to meet chromosomewide levels of significance after correction for multiple testing. These results suggested that 21q22.13 and 18p11.2 may harbor genes that increase the risks for early-onset and later-onset forms of BP, respectively. Lin et al. (2005) suggested that previous inconsistent linkage findings may have been due to differences in the AAO characteristics of the samples examined.

Linkage to Chromosome 22q12

For information on linkage of susceptibility to bipolar disorder to chromosome 22q12, see MAFD7 (612371) and Kelsoe et al., 2001.

Other Genomewide Linkage Studies

In an 'Old Order Amish revisited' study, Ginns et al. (1996) performed a genomewide linkage analysis in the Lancaster County group. In addition to the so-called pedigree 110, which was used for reporting the original genetic linkage data by Egeland et al. (1987), 2 pedigrees closely related to pedigree 110 and 2 other pedigrees, 210 and 310, were studied; all 5 pedigrees traced back to a founder couple who immigrated to the U.S. around 1750. The diagnoses were broken down into BP I (bipolar disorder with mania) and BP II (bipolar disorder with hypomania). Ginns et al. (1996) found evidence that regions on chromosomes 6, 13, and 15 harbor susceptibility loci for bipolar affective disorder, suggesting to them that bipolar affective disorder in the Old Order Amish is inherited as a complex trait.

LaBuda et al. (1996) reported progress of a full genome screen for loci predisposing to affective disorder in the Old Order Amish. To the previously reported lod score results published by Gerhard et al. (1994), they added lod score results for an additional 367 markers distributed throughout the genome, along with allele- and haplotype-sharing analyses on those chromosomes sufficiently saturated with markers. No statistically significant lod scores resulted. Some degree of allele sharing was found at 74 loci, and 3.8% of all markers analyzed passed more stringent significance criteria suggestive of linkage. Although genomic areas were highlighted for further exploration, the studies of LaBuda et al. (1996) identified no region clearly involved in the etiology of affective disorder in this population.

Risch and Botstein (1996) reviewed 19 linkage studies in manic-depressive illness; the studies purported to identify loci on 10 different autosomes, including both the short arm and the long arm of chromosome 18; linkage in 3 different regions of distal Xq had been proposed.

Segurado et al. (2003) applied the rank-based genome scan metaanalysis (GSMA) method (Levinson et al., 2003) to 18 bipolar disorder genome scan datasets in an effort to identify regions with significant support for linkage in the combined data. No region achieved genomewide statistical significance by several simulation-based criteria. The most significant p values (less than 0.01) were observed on chromosomes 9p22.3-p21.1, 10q11.21-q22.1, and 14q24.1-q32.12. Nominally significant p values were observed in several other chromosomal regions.

In a study of Ashkenazi Jewish families, Fallin et al. (2004) identified 4 regions suggestive of linkage to bipolar disorder on chromosomes 1, 3, 11, and 18.

Pato et al. (2004) conducted a genomewide scan of 16 extended families from a Portuguese genetic isolate with bipolar disorder and identified 3 regions on chromosomes 2, 11, and 19 with genomewide suggestive linkage and several other regions, including chromosome 6q, that approached suggestive levels of significance. This research replicated the finding of an elevated lod score near marker D6S1021 on chromosome 6q (peak NPL at D6S1021 = 2.02; p = 0.025). Higher density mapping provided additional support for this locus (NPL = 2.59; p = 0.0068) and another marker, D6S1639 (NPL = 3.06; p = 0.0019). On chromosome 11, linkage was found to D11S1883 (NPL = 3.15; p = 0.0014).

Middleton et al. (2004) performed a linkage analysis on 25 extended multiplex Portuguese families, including 12 families previously reported by Pato et al. (2004), segregating for bipolar disorder using a high-density SNP genotyping assay with a 0.21-Mb intermarker spacing. The analysis revealed genomewide significance with a maximum NPL of 4.20 and a maximum lod score of 3.56 at 6q22 (125.8 Mb). Several other areas had suggestive linkage: 2 regions on chromosome 2 (57 Mb, NPL = 2.98; 145 Mb, NPL = 3.09), chromosome 4 (91 Mb, NPL = 2.97), chromosome 11 (45-68 Mb, NPL = 2.51), chromosome 16 (20 Mb, NPL = 2.89), and chromosome 20 (60 Mb, NPL = 2.99).

McQueen et al. (2005) hypothesized that combining original genotype data on linkage of bipolar disorder would provide benefits of increased power and control over sources of heterogeneity that outweigh the difficulty and potential pitfalls of the implementation. Thus, they conducted a combined analysis using the original genotype data from 11 bipolar disorder genomewide linkage scans comprising 5,179 individuals from 1,067 families. Heterogeneity among studies was minimized in the analyses by using uniform methods of analysis and a common, standardized marker map. They demonstrated that combining original genome-scan data is a powerful approach for the elucidation of linkage regions underlying complex disease. Their results established genomewide significant linkage to BP on chromosomes 6q and 8q, and provided solid information to guide future gene-finding efforts that rely on fine mapping and association approaches. McQueen et al. (2005) observed the most significant result for 'narrow' BP (BP type I-only phenotype) on chromosome 6q. When the analysis was expanded to include BP II, the linkage signal on 6q was attenuated, despite the increase in the number of affected relative pairs (ARPs). In contrast, removal of the individuals with BP II from the analysis reduced the evidence of linkage on 8q.

Maziade et al. (2005) performed a dense genome scan to identify susceptibility loci shared by schizophrenia and bipolar disorder. They used the same ascertainment, statistical, and molecular methods for 480 members from 21 multigenerational families from Eastern Quebec affected by schizophrenia, bipolar affective disorder, or both. Five genomewide significant linkages with maximized lod scores over 4.0 were observed: 3 for bipolar disorder (15q11.1, 16p12.3, 18q12-q21) and 2 for the shared 'common locus' phenotype (15q26, 18q12-q21). Nine maximized lod scores exceeded the suggestive threshold of 2.6: 3 for bipolar disorder (3q21, 10p13, 12q23), 3 for schizophrenia (6p22, 13q13, 18q21), and 3 for the combined locus phenotype (2q12.3, 13q14, 16p13). Maziade et al. (2005) noted that all of the linkage signals overlapped formerly reported susceptibility regions except the signal at 15q26.

Cheng et al. (2006) conducted a 9-cM genomewide scan in a large bipolar pedigree sample from the National Institute of Mental Health Genetics Initiative (1,060 individuals from 154 multiplex families). Parametric and nonparametric analyses using both standard diagnostic models and comorbid conditions thought to identify phenotypic subtypes were conducted. Genomewide significant linkage was observed on chromosomes 10q25, 10p12, 16q24, 16p13, and 16p12 using standard diagnostic models, and on 6q25 (suicidal behavior), 7q21 (panic disorder), and 16p12 (psychosis) using phenotypic subtypes. Several other regions were suggestive of linkage including 1p13 (psychosis), 1p21 (psychosis), 1q44, 2q24 (suicidal behavior), 2p25 (psychosis), 4p16 (psychosis, suicidal behavior), 5p15, 6p25 (psychosis), 8p22 (psychosis), 8q24, 10q21, 10q25 (suicidal behavior), 10p11 (psychosis), 13q32 and 19p13 (psychosis).

Kimmel et al. (2005) reported a large family in which bipolar disorder appeared to cosegregate with autosomal dominant medullary cystic kidney disease. Of the 7 members with kidney disease, 5 had bipolar I disorder, one had unipolar depression, and 1 had a hyperthymic phenotype. The authors noted that the 2 known loci of medullary cystic kidney disease are in regions of chromosome 1 (MCKD1; 174000) and 16 (MCDK2; 162000) had previously been linked to bipolar disorder and schizophrenia.

Exclusion Studies

Although corticotropin-releasing hormone (CRH; 122560) and its function in the hypothalamic-pituitary-adrenal axis had been implicated in depression (Stratakis and Chrousos, 1995), Stratakis et al. (1997) could demonstrate no linkage between the CRH gene and bipolar affective disorder.


Clinical Management

Genetic Variation in Lithium Response

Lithium has been a first-line choice for maintenance treatment of bipolar disorders to prevent relapse of mania and depression, but many patients do not have a response to lithium treatment. To discover genetic variation influencing response to lithium treatment, Chen et al. (2014) performed a discovery genomewide association study and 2 sets of replication in patients with bipolar I disorder from the Taiwan Bipolar Consortium who were receiving lithium treatment. Two SNPs in high linkage disequilibrium, and , located in the introns of GADL1 (615601) showed the strongest associations in the genomewide association study (p = 5.50 x 10(-37) and p = 2.52 x 10 (-37), respectively) and in the replication sample of 100 patients (p = 9.19 x 10(-15) for each SNP). These 2 SNPs had a sensitivity of 93% for predicting a response to lithium and differentiated between patients with a good response and those with a poor response in the follow-up cohort. Resequencing of GADL1 revealed a novel variant in GADL1 intron 8, IVS8+48delG, that is in complete linkage disequilibrium with rs17026688 and is predicted to affect splicing. These variants are rare in persons of European and African ancestry.

In a comment on the report of Chen et al. (2014), Birnbaum et al. (2014) stated that they found a very low level of GADL1 expression in the brain, and suggested that there was higher expression in the kidney; therefore, they concluded that the role of GADL1 is more likely related to taurine biosynthesis and kidney function than to brain function. Birnbaum et al. (2014) encouraged a retrospective review of kidney function and lithium levels in bipolar patients. Lee and Cheng (2014) responded to Birnbaum et al. (2014) that taurine may cross the blood-brain barrier to interact directly with the glutamate NMDA receptor, suggesting that the role of GADL1 in kidney function may be related to bipolar disorder.

Commenting on the report of Chen et al. (2014), Vlachadis et al. (2014) speculated that, given the magnitude of the association between the presence of the T allele and the response to lithium therapy, there might be a significant difference in the minimum efficacious serum lithium level between carriers and noncarriers of the 'response' allele. Lee and Cheng (2014) replied that they were unable to examine this issue in their retrospective study. Lee and Cheng (2014) also thanked Vlachadis et al. (2014) for identifying an error in Table 2 of their article. The odds ratio for the association between the presence of the T allele and a response to lithium therapy in the combined cohorts should be 88.5 (95% confidence interval, 41.4-198.0).

Ikeda et al. (2014) assessed 154 Japanese patients with bipolar disorder and did not observe an association for any criterion, even in a stringent phenotype analysis, as reported by Chen et al. (2014) for the rs17026688 allele. The Consortium on Lithium Genetics (2014) undertook a replication study in 218 samples that they collected from patients. Because the alleles reported by Chen et al. (2014) are common in Asians but rare in whites, the Consortium on Lithium Genetics (2014) studied only the Asian samples that they had obtained. In samples obtained from patients of Han Chinese or Japanese ancestry, the authors found no association between the variants and a response to lithium therapy at any threshold on the Alda scale. Lee and Cheng (2014) replied to Ikeda et al. (2014) and the Consortium on Lithium Genetics (2014) that the methods were not duplicated precisely, and offered to provide help with independent replication studies.

Anghelescu and Dettling (2014) questioned the finding of Chen et al. (2014) of better response to lithium therapy among patients with rapid cycling. Lee and Cheng (2014) replied that the relationship between rapid cycling and the effect of lithium is controversial. Anghelescu and Dettling (2014) also recommended that, based on the assumption that lack of efficacy leads to premature termination of therapy, genetic analysis be extended to include patients who terminated therapy. Lee and Cheng (2014) stated that because adherence to lithium maintenance treatment involves such factors as illness behavior and cognitive function, rapport with psychiatrists, serious adverse effects of treatment, and familial and financial support, the effect of lithium in patients with premature termination cannot be assessed.


Molecular Genetics

Association with the SLC6A3 Gene on Chromosome 5p15

Greenwood et al. (2001) reported evidence for an association between the DAT1 gene (SLC6A3; 126455) and bipolar disorder in a sample of 50 parent-proband trios. Using the transmission disequilibrium test (TDT), they showed an association between a haplotype composed of 5 SNPs in the 3-prime region of the DAT1 gene, exon 9 through exon 15, and bipolar disorder (allele-wise TDT empirical P = 0.001; genotype-wise TDT empirical P = 0.0004). Greenwood et al. (2006) analyzed a total of 22 SNPs in the 50 previously studied parent-proband trios and an independent set of 70 parent-proband trios. Using TDT analysis, an intron 8 SNP and an intron 13 SNP were found to be moderately associated with bipolar disorder, each in 1 of the 2 independent samples. Analysis of haplotypes of all 22 SNPs in sliding windows of 5 adjacent SNPs revealed an association to the region near intron 7 and 8 in both samples (empirical P values of 0.002 and 0.001, respectively, for the same window).

Association with the HTR4 Gene on Chromosome 5q32

Ohtsuki et al. (2002) performed mutation and association analyses of the HTR4 gene (602164) on 5q32, which encodes the serotonin 5-HT4 receptor, in 96 Japanese patients, 48 with mood disorders and 48 with schizophrenia. Eight polymorphisms and 4 rare variants were identified. Four polymorphisms at or in close proximity to exon d showed significant association with bipolar disorder with odds ratios of 1.5 to 2; these included g.83097C/T (HTR4-SVR (splice variant region) SNP1), g.83159G/A (HTR4-SVRSNP2), g.83164(T)9-10 (HTR4-SVRSNP3), and g.83198A/G (HTR4-SVRSNP4). These polymorphisms were in linkage disequilibrium, and only 3 common haplotypes were observed. One haplotype (SVRSNP1, SVRSNP4 C-A) was significantly associated with bipolar disorder (p = 0.002). The genotypic and haplotypic associations with bipolar disorder were confirmed by the transmission disequilibrium test in the NIMH Genetics Initiative bipolar pedigrees with ratios of transmitted to not transmitted alleles of 1.5 to 2.0 (p = 0.01). The same haplotype that showed association with bipolar disorder was suggested to be associated with schizophrenia in the case-control analysis (p = 0.003) but was not confirmed when Japanese schizophrenia families were tested. The polymorphisms associated with mood disorder were located within the region that encodes the divergent C-terminal tails of the 5-HT4 receptor.

Association with the ABCA13 Gene on Chromosome 7p12.3

Knight et al. (2009) reported evidence that ABCA13 (607807) is a susceptibility factor for both schizophrenia and bipolar disorder. After the initial discovery of its disruption by a chromosome abnormality in a person with schizophrenia, Knight et al. (2009) resequenced ABCA13 exons in 100 cases with schizophrenia and 100 controls. Multiple rare coding variants were identified including 1 nonsense and 9 missense mutations and compound heterozygosity/homozygosity in 6 cases. Variants were genotyped in more than 1,600 additional schizophrenia, bipolar, and depression cases and in more than 950 control cohorts, and the frequency of all rare variants combined was greater than controls in schizophrenia (odds ratio = 1.93, P = 0.0057) and bipolar disorder (odds ratio = 2.71, P = 0.00007). The population-attributable risk of these mutations was 2.2% for schizophrenia and 4.0% for bipolar disorder. In a study of 21 families of mutation carriers, Knight et al. (2009) genotyped affected and unaffected relatives and found significant linkage (lod = 4.3) of rare variants with a phenotype including schizophrenia, bipolar disorder, and major depression. Knight et al. (2009) concluded that their data identified a candidate gene (ABCA13), highlighted the genetic overlap between schizophrenia, bipolar disorder, and depression, and suggested that rare coding variants may contribute significantly to risk of these disorders.

Association with the DRD4 Gene on Chromosome 11p15

Lopez Leon et al. (2005) conducted a metaanalysis to reevaluate the role of the 48-bp repeat polymorphism of the dopamine D4 receptor gene (DRD4; 126452) on chromosome 11p15 in mood disorders by studying 917 patients with unipolar or bipolar affective disorder and 1,164 control subjects from 12 samples using the Cockrane Review Manager. An association was found between all mood disorder groups and the DRD4 2-repeat 48-bp (2R) polymorphism. After correcting for multiple testing, the association between this repeat and bipolar affective disorder dropped to insignificance; however, the evidence for an association between the 2R allele and unipolar depression (p less than 0.001) and the combined group (p less than 0.001) remained.

Association with the BDNF Gene on Chromosome 11p13

Geller et al. (2004) noted that Sklar et al. (2002) and Neves-Pereira et al. (2002), using family-based methods, had found that the BDNF val66 allele (113505.0002) was preferentially transmitted to predominantly Caucasian adult probands with bipolar disorder. Geller et al. (2004) reported that the val66 allele was also preferentially transmitted in children with bipolar disorder. Lohoff et al. (2005) studied the BDNF val66 allele in 621 European patients with bipolar I disorder and positive family histories of affective disorder and 998 European controls. The frequency of the val66 allele was significantly increased in the bipolar I patients when compared to controls (P = 0.028; OR of 1.22).

Rybakowski et al. (2006) studied 111 patients with bipolar disorder, 129 schizophrenia patients, and 92 healthy controls utilizing the Wisconsin Card Sorting Test in the context of the BDNF V66M polymorphism. They found that bipolar patients with the val/val genotype made significantly fewer perseverative errors, had more correctly completed categories and conceptual level responses compared to bipolar patients with the val/met or met/met genotypes. No differences were observed in schizophrenia patients and controls.

Association with the CUX2 Gene on Chromosome 12q

Glaser et al. (2005) performed linkage disequilibrium mapping with 17 microsatellite markers across a 1.6-Mb segment forming the central part of the chromosome 12q23-q24 region implicated in several linkage studies for bipolar affective disorder. In a U.K. Caucasian case-control sample of 347 cases and 374 controls, a significant signal was identified (p = 0.0016) for the microsatellite marker M19 at 12q24. Genes, including regulatory elements, around this marker were screened for mutations and the linkage disequilibrium structure of the region determined by genotyping 22 SNPs and insertion/deletion polymorphisms in 94 individuals. Eleven haplotypes and SNPs were genotyped and 3, an insertion/deletion and a SNP within FLJ32356 (rs3840795 and rs933399) and a SNP within CUX2 (rs3847953), showed significant or nearly significant association with bipolar disorder after Bonferroni-correction (p values from 0.002-0.005).

Association with the SLC6A4 Gene on Chromosome 17q11

Lasky-Su et al. (2005) conducted a metaanalysis on case-control studies of the association between 2 polymorphisms of the SLC6A4 gene (a 17-bp VNTR in intron 2, and a 44-bp insertion/deletion in the promoter region; see 182138.0001) and affective disorders (bipolar disorder and unipolar depression) resulting in 4 metaanalyses. For each polymorphism, the authors assessed the evidence for allelic association, heterogeneity among studies, the influence of individual studies, and the potential for publication bias. The short alleles of the 44-bp insertion/deletion polymorphism showed a significant association with bipolar disorder (OR = 1.13, p = 0.001) but not unipolar disorder. The VNTR had no association with either disorder.

Cho et al. (2005) performed 2 metaanalyses of published studies involving the SLC4A4 gene as a candidate for bipolar disorder. The studies were population-based and family-based studies investigating the association with the promoter polymorphism (5-HTTLPR) and the intron 2 VNTR. Seventeen population-based studies comprising 1,712 cases and 2,583 controls and 6 family-based studies comprising 587 trios were included in the 5-HTTLPR metaanalysis. Sixteen population-based studies comprising 1,764 cases and 2,703 controls as well as for family-based studies comprising 382 trios were included in the intron 2 VNTR metaanalysis. Meta-regression showed that neither study type nor ethnic sample significantly contributed to heterogeneity of the metaanalyses. Overall, odds ratios suggested a very small but detectable effect of the serotonin transporter in susceptibility to bipolar disorder.

Association with the BCR Gene on Chromosome 22q11

Hashimoto et al. (2005) studied 171 patients with bipolar disorder, 329 with major depressive disorder, and 351 controls, all of whom were Japanese, for genetic association using 11 single nucleotide polymorphisms, including a missense polymorphism (N796S; rs140504) in the region of the breakpoint cluster region gene (BCR; 151410) on chromosome 22q11. Significant allelic associations with bipolar disorder were observed for 3 single nucleotide polymorphisms and associations with bipolar II disorder were observed for 10 polymorphisms including N796S (bipolar disorder, p = 0.0054; bipolar II disorder, p = 0.0014). There was a significant association with major depression for 6 polymorphisms. S796 allele carriers were in excess in bipolar II patients (p = 0.0046; OR = 3.1, 95% CI, 1.53-8.76).

Association with the COMT Gene on Chromosome 22q11

Comorbid panic disorder may define a subtype of bipolar disorder and may influence the strength of association between bipolar disorder and candidate genes involved in monoamine neurotransmission. Rotondo et al. (2002) studied the frequency of the V158M polymorphism of catechol-O-methyltransferase (COMT; 116790.0001), the 5-HTTLPR polymorphism of the serotonin transporter SLC6A4 (182138.0001), and a splice site polymorphism (IVS7+218C-A) of tryptophan hydroxylase (TPH; 191060) in a case-control association study of bipolar disorder patients with or without lifetime panic disorder. They compared results from DNA extracted from blood leukocytes of 111 unrelated subjects of Italian descent meeting DSM-III-R criteria for bipolar disorder, including 49 with and 62 without comorbid lifetime panic disorder, with those of 127 healthy subjects. Relative to the comparison subjects, subjects with bipolar disorder without panic disorder, but not those with comorbid bipolar disorder and panic disorder, showed significantly higher frequencies of the COMT met158 and the short 5-HTTLPR alleles. No statistical significance was found between the bipolar disorder groups and the TPH polymorphism. Rotondo et al. (2002) concluded that bipolar disorder without panic disorder may represent a more homogeneous form of illness and that variants of the COMT and SLC6A4 genes may influence clinical features of bipolar disorder.

Association with the XBP1 Gene on Chromosome 22q12

See 194355 for discussion of an association between susceptibility to bipolar disorder and a polymorphism in the XBP1 gene.

Association with the TRPM2 Gene on Chromosome 21q22

McQuillin et al. (2006) fine mapped chromosome 21q22.3 using 30 genetic markers in 600 bipolar subjects and 450 ancestrally matched supernormal controls. Allelic association with D21S171 (p = 0.016), rs1556314 (p = 0.008), and rs1785467 (p = 0.025) was observed. A test of association with a 3-locus haplotype across a susceptibility region was significant with a permutation test (p = 0.011), and a 2-SNP haplotype was also significantly associated with bipolar disorder (p = 0.01). The 2 brain-expressed genes present in the associated region, TRPM2 (603749) and C21ORF29 (612920), were sequenced from subjects who had inherited the associated marker alleles. The rs1556314 polymorphism in exon 11 of TRPM2, which causes an asp543-to-glu (D543E) change, showed the strongest association with bipolar disorder (p = 0.008). McQuillin et al. (2006) noted that deletion of exon 11 is known to cause dysregulation of cellular calcium homeostasis in response to oxidative stress.

Association with Repeat Expansions

Del-Favero et al. (2002) studied the CTG repeat in the third intron of the SEF2-1B gene (602272) located at 18q21.1 and the CAG repeat at the ERDA1 locus (603279) located at 17q21.3 in a large combined European case-control sample of bipolar affective disorder. The sample consisted of 403 patients and 486 controls matched for age, gender, and ethnicity. The patients were consecutively recruited from 5 participating centers in Belgium, Croatia, Denmark, Scotland, and Sweden. Dichotomous analysis of the combined sample did not show a significant difference in expansion frequency between cases and controls at either of the 2 loci. Secondary analysis after stratification for family history of affective disorder in first-degree relatives and disease severity revealed a borderline significant difference (p = 0.03) with a relative risk of 2.43 of developing bipolar disorder in familial cases homozygous for the expanded SEF2-1B allele. This finding rendered further support to the hypothesis that SEF2-1B cannot be excluded as a susceptibility gene for bipolar disorder or that SEF2-1B is in linkage disequilibrium with a causal gene for bipolar disorder.

Tsutsumi et al. (2004) used a repeat expansion detection assay to examine genomic DNA from 100 unrelated probands with schizophrenia and 68 unrelated probands with bipolar affective disorder for the presence of CAG/CTG repeat expansions. They found that 28% of probands with schizophrenia and 38% of probands with bipolar disorder had a CAG/CTG repeat in the expanded range. Each expansion could be explained by 1 of 3 nonpathogenic repeat expansions known to exist in the general population. Thus, a novel CAG/CTG repeat expansion was not a common genetic risk factor for bipolar disorder or schizophrenia in this study.

Association with the Mitochondrial MTND1 Gene

Munakata et al. (2004) reported an association between bipolar disorder and a polymorphism in the mitochondrial MTND1 gene (516000).

Gene Interaction and Locus Heterogeneity

Jamra et al. (2007) presented the first genomewide interaction and locus heterogeneity linkage scan in bipolar affective disorder, using a large linkage dataset (52 families of European descent; 448 participants and 259 affected individuals). The results provided the strongest evidence of interaction between BPAD genes on chromosome 2q22-q24 and 6q23-q24, which was observed symmetrically in both directions; nonparametric lod (NPL) scores of 7.55 on 2q and 7.63 on 6q; P less than 0.0001 and P = 0.0001, respectively, after a genomewide permutation procedure. The second-best BPAD interaction evidence was observed between 2q22-q24 and 15q26. Here, Jamra et al. (2007) also observed a symmetric interaction. Heterogeneity analysis revealed locus heterogeneity at 2q, 6p, 11p, 13q, and 22q, which was supported by adjacent markers within each region and by previously reported BPAD linkage findings.

Epigenetic Theory of Major Psychosis

As a test of the hypothesis that epigenetic misregulation is consistent with various nonmendelian features of schizophrenia (181500) and bipolar disorder, Mill et al. (2008) used CpG island microarrays to identify DNA methylation changes in the frontal cortex and germline associated with schizophrenia and bipolar disorder. In the frontal cortex they found evidence for psychosis-associated DNA methylation differences in numerous loci, including several involved in glutamatergic and GABAergic neurotransmission, brain development, and other processes functionally linked to disease etiology. DNA methylation changes in a significant proportion of these loci corresponded to reported changes of steady-state mRNA levels associated with psychosis. Gene ontology analysis highlighted epigenetic disruption to loci involved in mitochondrial function, brain development, and stress response. Changes in both the brain and the germline of affected individuals suggested that systemic epigenetic dysfunction may be associated with major psychosis. Mill et al. (2008) observed that frontal cortex DNA methylation in the BDNF gene (113505) is correlated with genotype at a nearby nonsynonymous SNP (V66M) that had been associated with major psychosis.

Reviews

See Kato (2007) for a review of molecular genetic findings on bipolar disorder and major depression from 2004 to 2007. Also see the reviews on the genetics of bipolar disorder by Craddock and Sklar (2009, 2013).

Associations Pending Confirmation

For a discussion of a possible association between variation in the KCNH7 gene and susceptibility to bipolar spectrum disorder, see 608169.0001.


Animal Model

Maeng et al. (2008) found that transgenic mice with selective neuron-specific overexpression of Bag1 (601497) in the hippocampus did not have obvious motor, sensory, or learning impairments, but showed less anxious behavior and had higher spontaneous recovery rates from helplessness behavior compared to wildtype mice. These transgenic mice also recovered faster from tests designed to trigger hyperlocomotion or addictive behaviors. In contrast, heterozygous Bag1 +/- mice showed enhanced extreme behavioral responses and less recovery in similar tests. The data suggested that BAG1 may play a role in affective resilience, and perhaps regulates recovery from behavioral impairments observed in patients with bipolar affective disorder. Maeng et al. (2008) postulated that the effects are mediated by BAG1 regulation of glucocorticoid receptor function.


REFERENCES

  1. Aita, V. M., Liu, J., Knowles, J. A., Terwilliger, J. D., Baltazar, R., Grunn, A., Loth, J. E., Kanyas, K., Lerer, B., Endicott, J., Wang, Z., Penchaszadeh, G., Gilliam, T. C., Baron, M. A comprehensive linkage analysis of chromosome 21q22 supports prior evidence for a putative bipolar affective disorder locus. Am. J. Hum. Genet. 64: 210-217, 1999. [PubMed: 9915960, related citations] [Full Text]

  2. Anghelescu, I., Dettling, M. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1859 only, 2014. [PubMed: 24806177, related citations] [Full Text]

  3. Bellivier, F., Golmard, J.-L., Henry, C., Leboyer, M., Schurhoff, F. Admixture analysis of age at onset in bipolar I affective disorder. Arch. Gen. Psychiat. 58: 510-512, 2001. [PubMed: 11343533, related citations] [Full Text]

  4. Berrettini, W. H., Ferraro, T. N., Goldin, L. R., Weeks, D. E., Detera-Wadleigh, S., Nurnberger, J. I., Jr., Gershon, E. S. Chromosome 18 DNA markers and manic-depressive illness: evidence for a susceptibility gene. Proc. Nat. Acad. Sci. 91: 5918-5921, 1994. [PubMed: 8016089, related citations] [Full Text]

  5. Birnbaum, R., Shin, J. H., Weinberger, D. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1855-1856, 2014. [PubMed: 24806173, related citations] [Full Text]

  6. Blackwood, D. H. R., He, L., Morris, S. W., McLean, A., Whitton, C., Thomson, M., Walker, M. T., Woodburn, K., Sharp, C. M., Wright, A. F., Shibasaki, Y., St. Clair, D. M., Porteous, D. J., Muir, W. J. A locus for bipolar affective disorder on chromosome 4p. Nature Genet. 12: 427-430, 1996. [PubMed: 8630499, related citations] [Full Text]

  7. Brent, D. A., Mann, J. J. Family genetic studies, suicide, and suicidal behavior. Am. J. Med. Genet. C Semin. Med. Genet. 133C: 13-24, 2005. [PubMed: 15648081, related citations] [Full Text]

  8. Brent, D. A., Mann, J. J. Familial pathways to suicidal behavior--understanding and preventing suicide among adolescents. New Eng. J. Med. 355: 2719-2721, 2006. [PubMed: 17192535, related citations] [Full Text]

  9. Cadoret, R. J. Evidence for genetic inheritance of primary affective disorder in adoptees. Am. J. Psychiat. 135: 463-466, 1978. [PubMed: 637144, related citations] [Full Text]

  10. Chen, C.-H., Lee, C.-S., Lee, M.-T. M., Ouyang, W.-C., Chen, C.-C., Chong, M.-Y., Wu, J.-Y., Tan, H. K.-L., Lee, Y.-C., Chuo, L.-J., Chiu, N.-Y., Tsang, H.-Y., and 31 others. Variant GADL1 and response to lithium therapy in bipolar I disorder. New Eng. J. Med. 370: 119-128, 2014. Note: Erratum: N. Eng. J. Med. 370: 1866 only, 2014. [PubMed: 24369049, related citations] [Full Text]

  11. Cheng, R., Juo, S. H., Loth, J. E., Nee, J., Iossifov, I., Blumenthal, R., Sharpe, L., Kanyas, K., Lerer, B., Lilliston, B., Smith, M., Trautman, K., Gilliam, T. C., Endicott, J., Baron, M. Genome-wide linkage scan in a large bipolar disorder sample from the National Institute of Mental Health genetics initiative suggests putative loci for bipolar disorder, psychosis, suicide, and panic disorder. Molec. Psychiat. 11: 252-260, 2006. [PubMed: 16402137, related citations] [Full Text]

  12. Cho, H. J., Meira-Lima, I., Cordeiro, Q., Michelon, L., Sham, P., Vallada, H., Collier, D. A. Population-based and family-based studies on the serotonin transporter gene polymorphisms and bipolar disorder: a systematic review and meta-analysis. Molec. Psychiat. 10: 771-781, 2005. [PubMed: 15824745, related citations] [Full Text]

  13. Cichon, S., Schumacher, J., Muller, D. J., Hurter, M., Windemuth, C., Strauch, K., Hemmer, S., Schulze, T. G., Schmidt-Wolf, G., Albus, M., Borrmann-Hassenbach, M., Franzek, E., and 15 others. A genome screen for genes predisposing to bipolar affective disorder detects a new susceptibility locus on 8q. Hum. Molec. Genet. 10: 2933-2944, 2001. Note: Erratum: Hum. Molec. Genet. 11: 1685 only, 2002. [PubMed: 11741836, related citations] [Full Text]

  14. Cichon, S., Winge, I., Mattheisen, M., Georgi, A., Karpushova, A., Freudenberg, J., Freudenberg-Hua, Y., Babadjanova, G., Van Den Bogaert, A., Abramova, L. I., Kapiletti, S., Knappskog, P. M., and 9 others. Brain-specific tryptophan hydroxylase 2 (TPH2): a functional Pro206Ser substitution and variation in the 5-prime region are associated with bipolar affective disorder. Hum. Molec. Genet. 17: 87-97, 2008. [PubMed: 17905754, related citations] [Full Text]

  15. Consortium on Lithium Genetics. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1857-1859, 2014. [PubMed: 24806176, related citations] [Full Text]

  16. Coon, H., Jensen, S., Hoff, M., Holik, J., Plaetke, R., Reimherr, F., Wender, P., Leppert, M., Byerley, W. A genome-wide search for genes predisposing to manic-depression, assuming autosomal dominant inheritance. Am. J. Hum. Genet. 52: 1234-1249, 1993. [PubMed: 8503452, related citations]

  17. Craddock, N., Owen, M., Burge, S., Kurian, B., Thomas, P., McGuffin, P. Familial cosegregation of major affective disorder and Darier's disease (keratosis follicularis). Brit. J. Psychiat. 164: 355-358, 1994. [PubMed: 8199789, related citations] [Full Text]

  18. Craddock, N., Sklar, P. Genetics of bipolar disorder: successful start to a long journey. Trends Genet. 25: 99-105, 2009. [PubMed: 19144440, related citations] [Full Text]

  19. Craddock, N., Sklar, P. Genetics of bipolar disorder. Lancet 381: 1654-1662, 2013. [PubMed: 23663951, related citations] [Full Text]

  20. Degn, B., Lundorf, M. D., Wang, A., Vang, M., Mors, O., Kruse, T. A., Ewald, H. Further evidence for a bipolar risk gene on chromosome 12q24 suggested by investigation of haplotype sharing and allelic association in patients from the Faroe Islands. Molec. Psychiat. 6: 450-455, 2001. [PubMed: 11443532, related citations] [Full Text]

  21. Del-Favero, J., Van Gestel, S., Borglum, A. D., Muir, W., Ewald, H., Mors, O., Ivezic, S., Oruc, L., Adolfsson, R., Blackwood, D., Kruse, T., Mendlewicz, J., Schalling, M., Van Broeckhoven, C. European combined analysis of the CTG18.1 and the ERDA1 CAG/CTG repeats in bipolar disorder. Europ. J. Hum. Genet. 10: 276-280, 2002. [PubMed: 12032737, related citations] [Full Text]

  22. Dick, D. M., Foroud, T., Flury, L., Bowman, E. S., Miller, M. J., Rau, N. L., Moe, P. R., Samavedy, N., El-Mallakh, R., Manji, H., Glitz, D. A., Meyer, E. T., and 20 others. Genomewide linkage analyses of bipolar disorder: a new sample of 250 pedigrees from the National Institute of Mental Health Genetics Initiative. Am. J. Hum. Genet. 73: 107-114, 2003. Note: Erratum: Am. J. Hum. Genet. 73: 979 only, 2003. [PubMed: 12772088, related citations] [Full Text]

  23. Drevets, W. C., Price, J. L., Simpson, J. R., Jr., Todd, R. D., Reich, T., Vannier, M., Ralchle, M. E. Subgenual prefrontal cortex abnormalities in mood disorders. Nature 386: 824-827, 1997. [PubMed: 9126739, related citations] [Full Text]

  24. Egeland, J. A., Gerhard, D. S., Pauls, D. L., Sussex, J. N., Kidd, K. K., Allen, C. R., Hostetter, A. M., Housman, D. E. Bipolar affective disorders linked to DNA markers on chromosome 11. Nature 325: 783-787, 1987. [PubMed: 2881209, related citations] [Full Text]

  25. Ekholm, J. M., Kieseppa, T., Hiekkalinna, T., Partonen, T., Paunio, T., Perola, M., Ekelund, J., Lonnqvist, J., Pekkarinen-Ijas, P., Peltonen, L. Evidence of susceptibility loci on 4q32 and 16p12 for bipolar disorder. Hum. Molec. Genet. 12: 1907-1915, 2003. [PubMed: 12874110, related citations] [Full Text]

  26. Escamilla, M. A., McInnes, L. A., Spesny, M., Reus, V. I., Service, S. K., Shimayoshi, N., Tyler, D. J., Silva, S., Molina, J., Gallegos, A., Meza, L., Cruz, M. L., and 10 others. Assessing the feasibility of linkage disequilibrium methods for mapping complex traits: an initial screen for bipolar disorder loci on chromosome 18. Am. J. Hum. Genet. 64: 1670-1678, 1999. [PubMed: 10330354, related citations] [Full Text]

  27. Escamilla, M. A., Spesny, M., Reus, V. I., Gallegos, A., Meza, L., Molina, J., Sandkuijl, L. A., Fournier, E., Leon, P. E., Smith, L. B., Freimer, N. B. Use of linkage disequilibrium approaches to map genes for bipolar disorder in the Costa Rican population. Am. J. Med. Genet. 67: 244-253, 1996. [PubMed: 8725743, related citations] [Full Text]

  28. Etain, B., Mathieu, F., Rietschel, M., Maier, W., Albus, M., McKeon, P., Roche, S., Kealey, C., Blackwood, D., Muir, W., Bellivier, F., Henry, C. {and 24 others}: Genome-wide scan for genes involved in bipolar affective disorder in 70 European families ascertained through a bipolar type I early-onset proband: supportive evidence for linkage at 3p14. Molec. Psychiat. 11: 685-694, 2006. [PubMed: 16534504, related citations] [Full Text]

  29. Ewald, H., Degn, B., Mors, O., Kruse, T. A. Significant linkage between bipolar affective disorder and chromosome 12q24. Psychiat. Genet. 8: 131-140, 1998. [PubMed: 9800214, related citations] [Full Text]

  30. Ewald, H., Flint, T., Kruse, T. A., Mors, O. A genome-wide scan shows significant linkage between bipolar disorder and chromosome 12q24.3 and suggestive linkage to chromosomes 1p22-21, 4p16, 6q14-22, 10q26 and 16p13.3. Molec. Psychiat. 7: 734-744, 2002. [PubMed: 12192618, related citations] [Full Text]

  31. Fallin, M. D., Lasseter, V. K., Wolyniec, P. S., McGrath, J. A., Nestadt, G., Valle, D., Liang, K.-Y., Pulver, A. E. Genomewide linkage scan for bipolar-disorder susceptibility loci among Ashkenazi Jewish families. Am. J. Hum. Genet. 75: 204-219, 2004. [PubMed: 15208783, images, related citations] [Full Text]

  32. Ferreira, M. A. R., O'Donovan, M. C., Meng, Y. A., Jones, I. R., Ruderfer, D. M., Jones, L., Fan, J., Kirov, G., Perlis, R. H., Green, E. K., Smoller, J. W., Grozeva, D., and 51 others. Collaborative genome-wide association analysis supports a role for ANK3 and CACNA1C in bipolar disorder. Nature Genet. 40: 1056-1058, 2008. [PubMed: 18711365, related citations] [Full Text]

  33. Freimer, N. B., Reus, V. I., Escamilla, M. A., McInnes, L. A., Spesny, M., Leon, P., Service, S. K., Smith, L. B., Silva, S., Rojas, E., Gallegos, A., Meza, L., Fournier, E., Baharloo, S., Blankenship, K., Tyler, D. J., Batki, S., Vinogradov, S., Weissenbach, J., Barondes, S. H., Sandkuijl, L. A. Genetic mapping using haplotype, association and linkage methods suggests a locus for severe bipolar disorder (BPI) at 18q22-q23. Nature Genet. 12: 436-441, 1996. [PubMed: 8630501, related citations] [Full Text]

  34. Garner, C., McInnes, L. A., Service, S. K., Spesny, M., Fournier, E., Leon, P., Freimer, N. B. Linkage analysis of a complex pedigree with severe bipolar disorder, using a Markov chain Monte Carlo method. Am. J. Hum. Genet. 68: 1061-1064, 2001. [PubMed: 11222106, related citations] [Full Text]

  35. Gelernter, J. Genetics of bipolar affective disorder: time for another reinvention? (Editorial) Am. J. Hum. Genet. 56: 1262-1266, 1995. [PubMed: 7762549, related citations]

  36. Geller, B., Badner, J. A., Tillman, R., Christian, S. L., Bolhofner, K., Cook, E. H., Jr. Linkage disequilibrium of the brain-derived neurotrophic factor val66met polymorphism in children with prepubertal and early adolescent bipolar disorder phenotype. Am. J. Psychiat. 161: 1698-1700, 2004. [PubMed: 15337662, related citations] [Full Text]

  37. Gerhard, D. S., LaBuda, M. C., Bland, S. D., Allen, C., Egeland, J. A., Pauls, D. L. Initial report of a genome search for the affective disorder predisposition gene in the old order Amish pedigrees: chromosomes 1 and 11. Am. J. Med. Genet. 54: 398-404, 1994. [PubMed: 7726215, related citations] [Full Text]

  38. Gill, M., McKeon, P., Humphries, P. Linkage analysis of manic depression in an Irish family using H-RAS-1 and INS DNA markers. J. Med. Genet. 25: 634-637, 1988. [PubMed: 3184144, related citations] [Full Text]

  39. Ginns, E. I., Ott, J., Egeland, J. A., Allen, C. R., Fann, C. S. J., Pauls, D. L., Weissenbach, J., Carulli, J. P., Falls, K. M., Keith, T. P., Paul, S. M. A genome-wide search for chromosomal loci linked to bipolar affective disorder in the Old Order Amish. Nature Genet. 12: 431-435, 1996. [PubMed: 8630500, related citations] [Full Text]

  40. Ginns, E. I., St. Jean, P., Philibert, R. A., Galdzicka, M., Damschroder-Williams, P., Thiel, B., Long, R. T., Ingraham, L. J., Dalwaldi, H., Murray, M. A., Ehlert, M., Paul, S., and 19 others. A genome-wide search for chromosomal loci linked to mental health wellness in relatives at high risk for bipolar affective disorder among the Old Order Amish. Proc. Nat. Acad. Sci. 95: 15531-15536, 1998. [PubMed: 9861003, images, related citations] [Full Text]

  41. Glaser, B., Kirov, G., Green, E., Craddock, N., Owen, M. J. Linkage disequilibrium mapping of bipolar affective disorder at 12q23-q24 provides evidence for association at CUX2 and FLJ32356. Am. J. Med. Genet. 132B: 38-45, 2005. [PubMed: 15389760, related citations] [Full Text]

  42. Goodwin, F. K., Jamison, K. R. Manic-Depressive Illness. New York: Oxford University Press 1990.

  43. Greenwood, T. A., Alexander, M., Keck, P. E., McElroy, S., Sadovnick, A. D., Remick, R. A., Kelsoe, J. R. Evidence for linkage disequilibrium between the dopamine transporter and bipolar disorder. Am. J. Med. Genet. 105: 145-151, 2001. [PubMed: 11304827, related citations] [Full Text]

  44. Greenwood, T. A., Schork, N. J., Eskin, E., Kelsoe, J. R. Identification of additional variants within the human dopamine transporter gene provides further evidence for an association with bipolar disorder in two independent samples. Molec. Psychiat. 11: 125-133, 2006. [PubMed: 16261167, related citations] [Full Text]

  45. Grof, P., Alda, M., Grof, E., Zvolsky, P., Walsh, M. Lithium response and genetics of affective disorders. J. Affect. Dis. 32: 85-95, 1994. [PubMed: 7829768, related citations] [Full Text]

  46. Hashimoto, R., Okada, T., Kato, T., Kosuga, A., Tatsumi, M., Kamijima, K., Kunugi, H. The breakpoint cluster region gene on chromosome 22q11 is associated with bipolar disorder. Biol. Psychiat. 57: 1097-1102, 2005. [PubMed: 15866548, related citations] [Full Text]

  47. Hill, E. M., Wilson, A. F., Elston, R. C., Winokur, G. Evidence for possible linkage between genetic markers and affective disorders. Biol. Psychiat. 24: 903-917, 1988. [PubMed: 3233232, related citations] [Full Text]

  48. Hodge, S. E., Wickramaratne, P. Statistical pitfalls in detecting age-of-onset anticipation: the role of correlation in studying anticipation and detecting ascertainment bias. Psychiat. Genet. 5: 43-47, 1995. [PubMed: 7582880, related citations] [Full Text]

  49. Holmes, D., Brynjolfsson, J., Brett, P., Curtis, D., Petursson, H., Sherrington, R., Gurling, H. No evidence for a susceptibility locus predisposing to manic depression in the region of the dopamine (D2) receptor gene. Brit. J. Psychiat. 158: 635-641, 1991. [PubMed: 1677600, related citations] [Full Text]

  50. Hong, K. S., McInnes, L. A., Service, S. K., Song, T., Lucas, J., Silva, S., Fournier, E., Leon, P., Molina, J., Reus, V. I., Sandkuijl, L. A., Freimer, N. B. Genetic mapping using haplotype and model-free linkage analysis supports previous evidence for a locus predisposing to severe bipolar disorder at 5q31-33. Am. J. Med. Genet. 125B: 83-86, 2004. [PubMed: 14755450, related citations] [Full Text]

  51. Ikeda, M., Kondo, K., Iwata, N. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1856-1857, 2014. [PubMed: 24806175, related citations] [Full Text]

  52. Jamra, R. A., Fuerst, R., Kaneva, R., Diaz, G. O., Rivas, F., Mayoral, F., Gay, E., Sans, S., Gonzalez, M. J., Gil, S., Cabaleiro, F., del Rio, F. {and 18 others}: The first genomewide interaction and locus-heterogeneity linkage scan in bipolar affective disorder: strong evidence of epistatic effects between loci on chromosome 2q and 6q. Am. J. Hum. Genet. 81: 974-986, 2007. [PubMed: 17924339, images, related citations] [Full Text]

  53. Joffe, R. T., Horvath, Z., Tarvydas, I. Bipolar affective disorder and thalassemia minor. (Letter) Am. J. Psychiat. 143: 933, 1986. [PubMed: 3717440, related citations] [Full Text]

  54. Kato, T. Molecular genetics of bipolar disorder and depression. Psychiat. Clin. Neurosci. 61: 3-19, 2007. [PubMed: 17239033, related citations] [Full Text]

  55. Kelsoe, J. R., Spence, M. A., Loetscher, E., Foguet, M., Sadovnick, A. D., Remick, R. A., Flodman, P., Khristich, J., Mroczkowski-Parker, Z., Brown, J. L., Masser, D., Ungerleider, S., Rapaport, M. H., Wishart, W. L., Luebbert, H. A genome survey indicates a possible susceptibility locus for bipolar disorder on chromosome 22. Proc. Nat. Acad. Sci. 98: 585-590, 2001. [PubMed: 11149935, images, related citations] [Full Text]

  56. Kimmel, R. J., Kovacs, I., Vrabel, C., Wood, B., Schalling, M., Kelsoe, J. R. Cosegregation of bipolar disorder and autosomal-dominant medullary cystic kidney disease in a large family. Am. J. Psychiat. 162: 1972-1974, 2005. [PubMed: 16199849, related citations] [Full Text]

  57. Kirk, R., Furlong, R. A., Amos, W., Cooper, G., Rubinsztein, J. S., Walsh, C., Paykel, E. S., Rubinsztein, D. C. Mitochondrial genetic analyses suggest selection against maternal lineages in bipolar affective disorder. Am. J. Hum. Genet. 65: 508-518, 1999. [PubMed: 10417293, related citations] [Full Text]

  58. Klein, P. S., Melton, D. A. A molecular mechanism for the effect of lithium on development. Proc. Nat. Acad. Sci. 93: 8455-8459, 1996. [PubMed: 8710892, related citations] [Full Text]

  59. Knight, H. M., Pickard, B. S., Maclean, A., Malloy, M. P., Soares, D. C., McRae, A. F., Condie, A., White, A., Hawkins, W., McGhee, K., van Beck, M., MacIntyre, D. J., Starr, J. M., Deary, I. J., Visscher, P. M., Porteous, D. J., Cannon, R. E., St Clair, D., Muir, W. J., Blackwood, D. H. R. A cytogenetic abnormality and rare coding variants identify ABCA13 as a candidate gene in schizophrenia, bipolar disorder, and depression. Am. J. Hum. Genet. 85: 833-846, 2009. [PubMed: 19944402, images, related citations] [Full Text]

  60. Knowles, J. A., Rao, P. A., Cox-Matise, T., Loth, J. E., de Jesus, G. M., Levine, L., Das, K., Penchaszadeh, G. K., Alexander, J. R., Lerer, B., Endicott, J., Ott, J., Gilliam, T. C., Baron, M. No evidence for significant linkage between bipolar affective disorder and chromosome 18 pericentromeric markers in a large series of multiplex extended pedigrees. Am. J. Hum. Genet. 62: 916-924, 1998. [PubMed: 9529343, related citations] [Full Text]

  61. LaBuda, M. C., Maldonado, M., Marshall, D., Otten, K., Gerhard, D. S. A follow-up report of a genome search for affective disorder predisposition loci in the Old Order Amish. Am. J. Hum. Genet. 59: 1343-1362, 1996. [PubMed: 8940281, related citations]

  62. Lasky-Su, J. A., Faraone, S. V., Glatt, S. J., Tsuang, M. T. Meta-analysis of the association between two polymorphisms in the serotonin transporter gene and affective disorders. Am. J. Med. Genet. 133B: 110-115, 2005. [PubMed: 15578606, related citations] [Full Text]

  63. Law, A., Richard, C. W., III, Cottingham, R. W., Jr., Lathrop, G. M., Cox, D. R., Myers, R. M. Genetic linkage analysis of bipolar affective disorder in an Old Order Amish pedigree. Hum. Genet. 88: 562-568, 1992. [PubMed: 1551659, related citations] [Full Text]

  64. Le, F., Mitchell, P., Vivero, C., Waters, B., Donald, J., Selbie, L. A., Shine, J., Schofield, P. Exclusion of close linkage of bipolar disorder to the GS-alpha subunit gene in nine Australian pedigrees. J. Affect. Disord. 32: 187-195, 1994. [PubMed: 7531727, related citations] [Full Text]

  65. Lee, C.-S., Cheng, A. T.-A. Reply to variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1859-1860, 2014. [PubMed: 24806172, related citations] [Full Text]

  66. Levinson, D. F., Levinson, M. D., Segurado, R., Lewis, C. M. Genome scan meta-analysis of schizophrenia and bipolar disorder. Part I: methods and power analysis. Am. J. Hum. Genet. 73: 17-33, 2003. [PubMed: 12802787, images, related citations] [Full Text]

  67. Lin, P.-I., McInnis, M. G., Potash, J. B., Willour, V. L., MacKinnon, D. F., Miao, K., DePaulo, J. R., Zandi, P. P. Assessment of the effect of age at onset on linkage to bipolar disorder: evidence on chromosomes 18p and 21q. Am. J. Hum. Genet. 77: 545-555, 2005. [PubMed: 16175501, images, related citations] [Full Text]

  68. Lohoff, F. W., Sander, T., Ferraro, T. N., Dahl, J. P., Gallinat, J., Berrettini, W. H. Confirmation of association between the val66-to-met polymorphism in the brain-derived neurotrophic factor (BDNF) gene and bipolar I disorder. Am. J. Med. Genet. 139B: 51-53, 2005. [PubMed: 16152572, related citations] [Full Text]

  69. Lombroso, C. Crime: Its Causes and Remedies. Boston: Little, Brown 1911.

  70. Lopez Leon, S., Croes, E. A., Sayed-Tabatabaei, F. A., Claes, S., Van Broeckhoven, C., van Duijn, C. M. The dopamine D4 receptor gene 48-base-pair-repeat polymorphism and mood disorders: a meta-analysis. Biol. Psychiat. 57: 999-1003, 2005. [PubMed: 15860340, related citations] [Full Text]

  71. Maeng, S., Hunsberger, J. G., Pearson, B., Yuan, P., Wang, Y., Wei, Y., McCammon, J., Schloesser, R. J., Zhou, R., Du, J., Chen, G., McEwen, B., Reed, J. C., Manji, H. K. BAG1 plays a critical role in regulating recovery from both manic-like and depression-like behavioral impairments. Proc. Nat. Acad. Sci. 105: 8766-8771, 2008. [PubMed: 18562287, images, related citations] [Full Text]

  72. Maziade, M., Roy, M.-A., Chagnon, Y. C., Cliche, D., Fournier, J.-P., Montgrain, N., Dion, C., Lavallee, J.-C., Garneau, Y., Gingras, N., Nicole, L., Pires, A., Ponton, A.-M., Potvin, A., Wallot, H., Merette, C. Shared and specific susceptibility loci for schizophrenia and bipolar disorder: a dense genome scan in Eastern Quebec families. Molec. Psychiat. 10: 486-499, 2005. [PubMed: 15534619, related citations] [Full Text]

  73. McInnes, L. A., Escamilla, M. A., Service, S. K., Reus, V. I., Leon, P., Silva, S., Rojas, E., Spesny, M., Baharloo, S., Blankenship, K., Peterson, A., Tyler, D., and 11 others. A complete genome screen for genes predisposing to severe bipolar disorder in two Costa Rican pedigrees. Proc. Nat. Acad. Sci. 93: 13060-13065, 1996. [PubMed: 8917544, images, related citations] [Full Text]

  74. McInnes, L. A., Service, S. K., Reus, V. I., Barnes, G., Charlat, O., Jawahar, S., Lewitzky, S., Yang, Q., Duong, Q., Spesny, M., Araya, C., Araya, X., and 23 others. Fine-scale mapping of a locus for severe bipolar mood disorder on chromosome 18p11.3 in the Costa Rican population. Proc. Nat. Acad. Sci. 98: 11485-11490, 2001. [PubMed: 11572994, images, related citations] [Full Text]

  75. McInnis, M. G., McMahon, F. J., Chase, G. A., Simpson, S. G., Ross, C. A., DePaulo, J. R., Jr. Anticipation in bipolar affective disorder. Am. J. Hum. Genet. 53: 385-390, 1993. [PubMed: 8328456, related citations]

  76. McInnis, M. G. Anticipation: an old idea in new genes. (Editorial) Am. J. Hum. Genet. 59: 973-979, 1996. [PubMed: 8900222, related citations]

  77. McMahon, F. J., Simpson, S. G., McInnis, M. G., Badner, J. A., MacKinnon, D. F., DePaulo, J. R. Linkage of bipolar disorder to chromosome 18q and the validity of bipolar II disorder. Arch. Gen. Psychiat. 58: 1025-1031, 2001. [PubMed: 11695948, related citations] [Full Text]

  78. McMahon, F. J., Stine, O. C., Meyers, D. A., Simpson, S. G., DePaulo, J. R. Patterns of maternal transmission in bipolar affective disorder. Am. J. Hum. Genet. 56: 1277-1286, 1995. [PubMed: 7762550, related citations]

  79. McQueen, M. B., Devlin, B., Faraone, S. V., Nimgaonkar, V. L., Sklar, P., Smoller, J. W., Jamra, R. A., Albus, M., Bacanu, S.-A., Baron, M., Barrett, T. B., Berrettini, W., and 41 others. Combined analysis from eleven linkage studies of bipolar disorder provides strong evidence of susceptibility loci on chromosomes 6q and 8q. Am. J. Hum. Genet. 77: 582-595, 2005. [PubMed: 16175504, images, related citations] [Full Text]

  80. McQuillin, A., Bass, N. J., Kalsi, G., Lawrence, J., Puri, V., Choudhury, K., Detera-Wadleigh, S. D., Curtis, D., Gurling, H. M. D. Fine mapping of a susceptibility locus for bipolar and genetically related unipolar affective disorders, to a region containing the C21ORF29 and TRPM2 genes on chromosome 21q22.3. Molec. Psychiat. 11: 134-142, 2006. [PubMed: 16205735, related citations] [Full Text]

  81. Middleton, F. A., Pato, M. T., Gentile, K. L., Morley, C. P., Zhao, X., Eisener, A. F., Brown, A., Petryshen, T. L., Kirby, A. N., Medeiros, H., Carvalho, C., Macedo, A., and 11 others. Genomewide linkage analysis of bipolar disorder by use of a high-density single-nucleotide-polymorphism (SNP) genotype assay: a comparison with microsatellite marker assays and finding of significant linkage to chromosome 6q22. Am. J. Hum. Genet. 74: 886-897, 2004. [PubMed: 15060841, images, related citations] [Full Text]

  82. Mill, J., Tang, T., Kaminsky, Z., Khare, T., Yazdanpanah, S., Bouchard, L., Jia, P., Assadzadeh, A., Flanagan, J., Schumacher, A., Wang, S.-C., Petronis, A. Epigenetic profiling reveals DNA-methylation changes associated with major psychosis. Am. J. Hum. Genet. 82: 696-711, 2008. [PubMed: 18319075, images, related citations] [Full Text]

  83. Morel, B. A. Traite des Degenerescences. Paris: JB Bailliere 1857.

  84. Morissette, J., Villeneuve, A., Bordeleau, L., Rochette, D., Laberge, C., Gagne, B. et al. Genome-wide search for linkage of bipolar affective disorders in a very large pedigree derived from a homogeneous population in Quebec points to a locus of major effect on chromosome 12q23-q24. Am. J. Med. Genet. 88: 567-587, 1999. [PubMed: 10490718, related citations] [Full Text]

  85. Munakata, K., Tanaka, M., Mori, K., Washizuka, S., Yoneda, M., Tajima, O., Akiyama, T., Nanko, S., Kunugi, H., Tadokoro, K., Ozaki, N., Inada, T., Sakamoto, K., Fukunaga, T., Iijima, Y., Iwata, N., Tatsumi, M., Yamada, K., Yoshikawa, T., Kato, T. Mitochondrial DNA 3644T-C mutation associated with bipolar disorder. Genomics 84: 1041-1050, 2004. [PubMed: 15533721, related citations] [Full Text]

  86. Neiswanger, K., Slaugenhaupt, S. A., Hughes, H. B., Frank, E., Frankel, D. R., McCarty, M. J., Chakravarti, A., Zubenko, G. S., Kupfer, D. J., Kaplan, B. B. Evidence against close linkage of unipolar affective illness to human chromosome 11p markers HRAS1 and INS and chromosome Xq marker DXS52. Biol. Psychiat. 28: 63-72, 1990. [PubMed: 1973904, related citations] [Full Text]

  87. Neves-Pereira, M., Mundo, E., Muglia, P., King, N., Macciardi, F., Kennedy, J. L. The brain-derived neurotrophic factor gene confers susceptibility to bipolar disorder: evidence from a family-based association study. Am. J. Hum. Genet. 71: 651-655, 2002. [PubMed: 12161822, related citations] [Full Text]

  88. O'Reilly, R. L., Bogue, L., Singh, S. M. Pharmacogenetic response to antidepressants in a multicase family with affective disorder. Biol. Psychiat. 36: 467-471, 1994. [PubMed: 7811843, related citations] [Full Text]

  89. Ohtsuki, T., Ishiguro, H., Detera-Wadleigh, S. D., Toyota, T., Shimizu, H., Yamada, K., Yoshitsugu, K., Hattori, E., Yoshikawa, T., Arinami, T. Association between serotonin 4 receptor gene polymorphisms and bipolar disorder in Japanese case-control samples and the NIMH Genetics Initiative bipolar pedigrees. Molec. Psychiat. 7: 954-961, 2002. [PubMed: 12399948, related citations] [Full Text]

  90. Ophoff, R. A., Escamilla, M. A., Service, S. K., Spesny, M., Meshi, D. B., Poon, W., Molina, J., Fournier, E., Gallegos, A., Mathews, C., Neylan, T., Batki, S. L., Roche, E., Ramirez, M., Silva, S., De Mille, M. C., Dong, P., Leon, P. E., Reus, V. I., Sandkuijl, L. A., Freimer, N. B. Genomewide linkage disequilibrium mapping of severe bipolar disorder in a population isolate. Am. J. Hum. Genet. 71: 565-574, 2002. [PubMed: 12119601, images, related citations] [Full Text]

  91. Pakstis, A. J., Kidd, J. R., Castiglione, C. M., Kidd, K. K. Status of the search for a major genetic locus for affective disorder in the Old Order Amish. Hum. Genet. 87: 475-483, 1991. [PubMed: 1879834, related citations] [Full Text]

  92. Pardue, L. H. Familial unipolar depressive illness: a pedigree study. Am. J. Psychiat. 132: 970-972, 1975. [PubMed: 1155636, related citations] [Full Text]

  93. Park, N., Juo, S. H., Cheng, R., Liu, J., Loth, J. E., Lilliston, B., Nee, J., Grunn, A., Kanyas, K., Lerer, B., Endicott, J., Gilliam, T. C., Baron, M. Linkage analysis of psychosis in bipolar pedigrees suggests novel putative loci for bipolar disorder and shared susceptibility with schizophrenia. Molec. Psychiat. 9: 1091-1099, 2004. [PubMed: 15241432, related citations] [Full Text]

  94. Pato, C. N., Pato, M. T., Kirby, A., Petryshen, T. L., Medeiros, H., Carvalho, C., Macedo, A., Dourado, A., Coelho, I., Valente, J., Soares, M. J., Ferreira, C. P., Lei, M., Verner, A., Hudson, T. J., Morley, C. P., Kennedy, J. L., Azevedo, M. H., Daly, M. J., Sklar, P. Genome-wide scan in Portuguese Island families implicates multiple loci in bipolar disorder: fine mapping adds support on chromosomes 6 and 11. Am. J. Med. Genet. 127B: 30-34, 2004. [PubMed: 15108176, related citations] [Full Text]

  95. Pauls, D. L., Gerhard, D. S., Lacy, L. G., Hostetter, A. M., Allen, C. R., Bland, S. D., LaBuda, M. C., Egeland, J. A. Linkage of bipolar affective disorders to markers on chromosome 11p is excluded in a second lateral extension of Amish pedigree 110. Genomics 11: 730-736, 1991. [PubMed: 1774072, related citations] [Full Text]

  96. Pauls, D. L., Ott, J., Paul, S. M., Allen, C. R., Fann, C. S. J., Carulli, J. P., Falls, K. M., Bouthillier, C. A., Gravius, T. C., Keith, T. P., Egeland, J. A., Ginns, E. I. Linkage analysis of chromosome 18 markers do not identify a major susceptibility locus for bipolar affective disorder in the Old Order Amish. Am. J. Hum. Genet. 57: 636-643, 1995. [PubMed: 7668292, related citations]

  97. Radhakrishna, U., Senol, S., Herken, H., Gucuyener, K., Gehrig, C., Blouin, J.-L., Akarsu, N. A., Antonarakis, S. E. An apparently dominant bipolar affective disorder (BPAD) locus on chromosome 20p11.2-q11.2 in a large Turkish pedigree. Europ. J. Hum. Genet. 9: 39-44, 2001. [PubMed: 11175298, related citations] [Full Text]

  98. Risch, N., Botstein, D. A manic depressive history. Nature Genet. 12: 351-352, 1996. [PubMed: 8630482, related citations] [Full Text]

  99. Rotondo, A., Mazzanti, C., Dell'Osso, L., Rucci, P., Sullivan, P., Bouanani, S., Gonnelli, C., Goldman, D., Cassano, G. B. Catechol O-methyltransferase, serotonin transporter, and tryptophan hydroxylase gene polymorphisms in bipolar disorder patients with and without comorbid panic disorder. Am. J. Psychiat. 159: 23-29, 2002. [PubMed: 11772685, related citations] [Full Text]

  100. Rybakowski, J., Borkowska, A., Skibinska, M., Hauser, J. Illness-specific association of val66met BDNF polymorphism with performance on Wisconsin Card Sorting Test in bipolar disorder. (Letter) Molec. Psychiat. 11: 122-124, 2006. [PubMed: 16222333, related citations] [Full Text]

  101. Schulze, T. G., Buervenich, S., Badner, J. A., Steele, C. J. M., Detera-Wadleigh, S. D., Dick, D., Foroud, T., Cox, N. J., MacKinnon, D. F., Potash, J. B., Berrettini, W. H., Byerley, W., Coryell, W., DePaulo, J. R., Jr., Gershon, E. S., Kelsoe, J. R., McInnis, M. G., Murphy, D. L., Reich, T., Scheftner, W., Nurnberger, J. I., Jr., McMahon, F. J. Loci on chromosomes 6q and 6p interact to increase susceptibility to bipolar affective disorder in the National Institute of Mental Health Genetics Initiative pedigrees. Biol. Psychiat. 56: 18-23, 2004. [PubMed: 15219468, related citations] [Full Text]

  102. Segurado, R., Detera-Wadleigh, S. D., Levinson, D. F., Lewis, C. M., Gill, M., Nurnberger, J. I., Jr., Craddock, N., DePaulo, J. R., Baron, M., Gershon, E. S., Ekholm, J., Cichon, S., and 66 others. Genome scan meta-analysis of schizophrenia and bipolar disorder, Part III: bipolar disorder. Am. J. Hum. Genet. 73: 49-62, 2003. [PubMed: 12802785, images, related citations] [Full Text]

  103. Sherrington, R., Curtis, D., Brynjolfsson, J., Moloney, E., Rifkin, L., Petursson, H., Gurling, H. A linkage study of affective disorder with DNA markers for the ABO-AK1-ORM linkage group near the dopamine beta hydroxylase gene. Biol. Psychiat. 36: 434-442, 1994. [PubMed: 7811839, related citations] [Full Text]

  104. Sklar, P., Gabriel, S. B., McInnis, M. G., Bennett, P., Lim, Y.-M., Tsan, G., Schaffner, S., Kirov, G., Jones, I., Owen, M., Craddock, N., DePaulo, J. R., Lander, E. S. Family-based association study of 76 candidate genes in bipolar disorder: BDNF is a potential risk locus. Molec. Psychiat. 7: 579-593, 2002. [PubMed: 12140781, related citations] [Full Text]

  105. Smeraldi, E., Negri, F., Melica, A. M., Scorza-Smeraldi, R. HLA system and affective disorders: a sibship genetic study. Tissue Antigens 12: 270-274, 1978. [PubMed: 725908, related citations] [Full Text]

  106. Smyth, C., Kalsi, G., Curtis, D., Brynjolfsson, J., O'Neill, J., Rifkin, L., Moloney, E., Murphy, P., Petursson, H., Gurling, H. Two-locus admixture linkage analysis of bipolar and unipolar affective disorder supports the presence of susceptibility loci on chromosomes 11p15 and 21q22. Genomics 39: 271-278, 1997. [PubMed: 9119364, related citations] [Full Text]

  107. Stancer, H. C., Weitkamp, L. R., Persad, E., Flood, C., Jorna, T., Guttormsen, S. A., Yagnow, R. L. Confirmation of the relationship of HLA (chromosome 6) genes to depression and manic depression. II. The Ontario follow-up and analysis of 117 kindreds. Ann. Hum. Genet. 52: 279-298, 1988. [PubMed: 3268040, related citations] [Full Text]

  108. Stine, O. C., Xu, J., Koskela, R., McMahon, F. J., Gschwend, M., Friddle, C., Clark, C. D., McInnis, M. G., Simpson, S. G., Breschel, T. S., Vishio, E., Riskin, K., Feilotter, H., Chen, E., Shen, S., Folstein, S., Meyers, D. A., Botstein, D., Marr, T. G., DePaulo, J. R. Evidence for linkage of bipolar disorder to chromosome 18 with a parent-of-origin effect. Am. J. Hum. Genet. 57: 1384-1394, 1995. [PubMed: 8533768, related citations]

  109. Stratakis, C. A., Chrousos, G. P. Neuroendocrinology and pathophysiology of the stress system. Ann. N.Y. Acad. Sci. 771: 1-18, 1995. [PubMed: 8597390, related citations] [Full Text]

  110. Stratakis, C. A., Sarlis, N. J., Berrettini, W. H., Badner, J. A., Chrousos, G. P., Gershon, E. S., Detera-Wadleigh, S. D. Lack of linkage between the corticotropin-releasing hormone (CRH) gene and bipolar affective disorder. Molec. Psychiat. 2: 483-485, 1997. [PubMed: 9399692, related citations] [Full Text]

  111. Straub, R. E., Lehner, T., Luo, Y., Loth, J. E., Shao, W., Sharpe, L., Alexander, J. R., Das, K., Simon, R., Fieve, R. R., Lerer, B., Endicott, J., Ott, J., Gilliam, T. C., Baron, M. A possible vulnerability locus for bipolar affective disorder on chromosome 21q22.3. Nature Genet. 8: 291-296, 1994. [PubMed: 7874172, related citations] [Full Text]

  112. Tanna, V. L., Wilson, A. F., Winokur, G., Elston, R. C. Linkage analysis of pure depressive disease. J. Psychiat. Res. 23: 99-107, 1989. [PubMed: 2585349, related citations] [Full Text]

  113. Tsutsumi, T., Holmes, S. E., McInnis, M. G., Sawa, A., Callahan, C., DePaulo, J. R., Ross, C. A., DeLisi, L. E., Margolis, R. L. Novel CAG/CTG repeat expansion mutations do not contribute to the genetic risk for most cases of bipolar disorder or schizophrenia. Am. J. Med. Genet. 124B: 15-19, 2004. [PubMed: 14681907, related citations] [Full Text]

  114. Venken, T., Claes, S., Sluijs, S., Paterson, A. D., van Duijn, C., Adolfsson, R., Del-Favero, J., Van Broeckhoven, C. Genomewide scan for affective disorder susceptibility loci in families of a northern Swedish isolated population. Am. J. Hum. Genet. 76: 237-248, 2005. [PubMed: 15614721, images, related citations] [Full Text]

  115. Vlachadis, N., Vrachnis, N., Economou, E. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1856 only, 2014. [PubMed: 24806174, related citations] [Full Text]

  116. Weitkamp, L. R., Stancer, H. C., Persad, E., Flood, C., Guttormsen, S. Depressive disorders and HLA: a gene on chromosome 6 that can affect behavior. New Eng. J. Med. 305: 1301-1306, 1981. [PubMed: 6945479, related citations] [Full Text]

  117. Weitkamp, L. R., Stancer, H. C. Analysis of the Toronto-Rochester depression study follow-up data confirms an HLA-region gene contribution to susceptibility to affective disorder. Genet. Epidemiol. 6: 305-310, 1989. [PubMed: 2499504, related citations] [Full Text]

  118. Weitkamp, L. R. HLA and disease: predictions for HLA haplotype sharing in families. Am. J. Hum. Genet. 33: 776-784, 1981. [PubMed: 7294026, related citations]

  119. Weitkamp, L. R. Genetic analysis of multifactorial diseases. (Letter) Am. J. Hum. Genet. 35: 130-133, 1983. [PubMed: 6823971, related citations]

  120. Williams, R. S. B., Cheng, L., Mudge, A. W., Harwood, A. J. A common mechanism of action for three mood-stabilizing drugs. Nature 417: 292-295, 2002. [PubMed: 12015604, related citations] [Full Text]

  121. Wilson, A. F., Elston, R. C., Mallott, D. B., Tran, L. D., Winokur, G. The current status of genetic linkage studies of alcoholism and unipolar depression. Psychiat. Genet. 2: 107-124, 1991.

  122. Wilson, A. F., Tanna, V. L., Winokur, G., Elston, R. C., Hill, E. M. Linkage analysis of depression spectrum disease. Biol. Psychiat. 26: 163-175, 1989. [PubMed: 2736265, related citations] [Full Text]

  123. Wingerson, L. Searching for depression genes. Discover 3(2): 60-64, 1982.

  124. Wright, A. F., Crichton, D. N., Loudon, J. B., Morten, J. E. N., Steel, C. M. Beta-adrenoceptor binding defects in cell lines from families with manic-depressive disorder. Ann. Hum. Genet. 48: 201-214, 1984. [PubMed: 6087716, related citations] [Full Text]


Cassandra L. Kniffin - updated : 1/28/2015
Ada Hamosh - updated : 6/10/2014
Ada Hamosh - updated : 2/6/2014
Ada Hamosh - updated : 3/16/2010
Cassandra L. Kniffin - updated : 8/25/2009
Cassandra L. Kniffin - updated : 7/29/2009
Ada Hamosh - updated : 10/22/2008
Carol A. Bocchini - updated : 9/8/2008
Victor A. McKusick - updated : 4/14/2008
Victor A. McKusick - updated : 10/10/2007
Ada Hamosh - updated : 7/19/2007
Victor A. McKusick - updated : 2/26/2007
John Logan Black, III - updated : 1/23/2007
John Logan Black, III - updated : 11/16/2006
John Logan Black, III - updated : 11/15/2006
John Logan Black, III - updated : 6/7/2006
John Logan Black, III - updated : 6/2/2006
John Logan Black, III - updated : 5/25/2006
John Logan Black, III - updated : 5/16/2006
John Logan Black, III - updated : 4/21/2006
John Logan Black, III - updated : 4/6/2006
John Logan Black, III - updated : 4/6/2006
John Logan Black, III - updated : 12/6/2005
Victor A. McKusick - updated : 10/5/2005
John Logan Black, III - updated : 8/8/2005
John Logan Black, III - updated : 7/26/2005
John Logan Black, III - updated : 7/22/2005
John Logan Black, III - updated : 7/21/2005
John Logan Black, III - updated : 7/20/2005
George E. Tiller - updated : 5/5/2005
John Logan Black, III - updated : 4/8/2005
John Logan Black, III - updated : 3/2/2005
Cassandra L. Kniffin - updated : 10/28/2004
Victor A. McKusick - updated : 8/12/2004
John Logan Black, III - updated : 10/30/2003
John Logan Black, III - updated : 8/19/2003
Victor A. McKusick - updated : 7/9/2003
John Logan Black, III - updated : 11/8/2002
Michael B. Petersen - updated : 11/7/2002
Victor A. McKusick - updated : 10/9/2002
Dawn Watkins-Chow - updated : 8/23/2002
George E. Tiller - updated : 8/15/2002
Victor A. McKusick - updated : 7/31/2002
Ada Hamosh - updated : 5/29/2002
Victor A. McKusick - updated : 1/30/2002
Victor A. McKusick - updated : 11/1/2001
Victor A. McKusick - updated : 9/12/2001
Michael B. Petersen - updated : 7/17/2001
Victor A. McKusick - updated : 2/26/2001
Victor A. McKusick - updated : 1/12/2000
Victor A. McKusick - updated : 3/1/1999
Victor A. McKusick - updated : 2/8/1999
Victor A. McKusick - updated : 5/13/1998
Victor A. McKusick - updated : 4/23/1997
Victor A. McKusick - updated : 2/20/1997
Creation Date:
Victor A. McKusick : 6/4/1986
alopez : 09/30/2022
carol : 09/25/2020
carol : 01/22/2020
carol : 07/18/2017
alopez : 07/17/2017
carol : 07/21/2015
alopez : 1/30/2015
ckniffin : 1/28/2015
alopez : 6/10/2014
carol : 4/11/2014
alopez : 2/6/2014
carol : 9/23/2011
alopez : 11/10/2010
wwang : 4/6/2010
alopez : 3/18/2010
alopez : 3/18/2010
terry : 3/16/2010
wwang : 9/8/2009
wwang : 9/4/2009
ckniffin : 8/25/2009
ckniffin : 7/29/2009
mgross : 7/17/2009
carol : 5/1/2009
carol : 12/2/2008
carol : 10/28/2008
terry : 10/22/2008
carol : 9/8/2008
carol : 5/7/2008
carol : 5/7/2008
alopez : 4/30/2008
terry : 4/14/2008
alopez : 10/15/2007
terry : 10/10/2007
alopez : 7/24/2007
terry : 7/19/2007
carol : 5/10/2007
terry : 5/9/2007
alopez : 3/8/2007
terry : 2/26/2007
carol : 1/23/2007
carol : 1/23/2007
carol : 11/16/2006
terry : 11/16/2006
carol : 11/16/2006
terry : 11/15/2006
carol : 11/13/2006
wwang : 6/7/2006
wwang : 6/7/2006
terry : 6/2/2006
wwang : 5/25/2006
wwang : 5/16/2006
carol : 4/25/2006
terry : 4/21/2006
carol : 4/11/2006
wwang : 4/10/2006
terry : 4/6/2006
terry : 4/6/2006
terry : 12/13/2005
carol : 12/6/2005
wwang : 11/7/2005
alopez : 10/6/2005
terry : 10/5/2005
carol : 10/4/2005
terry : 8/8/2005
carol : 7/28/2005
terry : 7/26/2005
carol : 7/25/2005
terry : 7/22/2005
carol : 7/21/2005
terry : 7/21/2005
carol : 7/21/2005
terry : 7/20/2005
tkritzer : 5/5/2005
mgross : 4/8/2005
mgross : 4/5/2005
tkritzer : 3/2/2005
tkritzer : 11/1/2004
ckniffin : 10/28/2004
terry : 8/12/2004
alopez : 3/17/2004
carol : 10/30/2003
carol : 10/13/2003
carol : 9/9/2003
carol : 8/20/2003
terry : 8/19/2003
terry : 8/15/2003
tkritzer : 7/23/2003
tkritzer : 7/15/2003
terry : 7/9/2003
cwells : 5/21/2003
carol : 11/21/2002
carol : 11/8/2002
cwells : 11/7/2002
carol : 10/11/2002
tkritzer : 10/10/2002
terry : 10/9/2002
mgross : 10/9/2002
tkritzer : 8/23/2002
cwells : 8/15/2002
mgross : 8/5/2002
terry : 7/31/2002
alopez : 5/30/2002
terry : 5/29/2002
terry : 3/8/2002
alopez : 2/5/2002
terry : 1/30/2002
carol : 11/5/2001
terry : 11/1/2001
mcapotos : 9/18/2001
mcapotos : 9/12/2001
carol : 7/17/2001
mgross : 7/17/2001
mgross : 7/17/2001
mgross : 2/26/2001
carol : 5/4/2000
terry : 4/24/2000
mgross : 2/7/2000
terry : 1/12/2000
carol : 8/13/1999
carol : 7/16/1999
carol : 6/25/1999
carol : 6/25/1999
terry : 4/30/1999
carol : 3/22/1999
terry : 3/1/1999
carol : 2/14/1999
terry : 2/8/1999
dkim : 12/10/1998
alopez : 5/19/1998
terry : 5/13/1998
terry : 11/11/1997
terry : 11/11/1997
terry : 11/11/1997
terry : 11/10/1997
alopez : 7/29/1997
terry : 7/28/1997
alopez : 7/7/1997
carol : 6/23/1997
alopez : 4/23/1997
terry : 4/23/1997
alopez : 3/20/1997
alopez : 3/19/1997
mark : 2/20/1997
terry : 2/12/1997
terry : 1/16/1997
jamie : 1/15/1997
terry : 1/8/1997
terry : 12/30/1996
terry : 12/20/1996
terry : 12/10/1996
terry : 12/5/1996
mark : 4/3/1996
terry : 4/3/1996
mark : 3/21/1996
terry : 3/19/1996
mark : 3/15/1996
mark : 3/15/1996
mark : 3/14/1996
mark : 3/1/1996
terry : 2/29/1996
mark : 12/15/1995
terry : 12/14/1995
mark : 9/10/1995
terry : 3/3/1995
carol : 3/1/1995
jason : 7/12/1994
mimadm : 6/25/1994
warfield : 4/8/1994

% 125480

MAJOR AFFECTIVE DISORDER 1; MAFD1


Alternative titles; symbols

MANIC-DEPRESSIVE PSYCHOSIS; MD1
BIPOLAR AFFECTIVE DISORDER; BPAD
MANIC-DEPRESSIVE PSYCHOSIS, AUTOSOMAL


ICD10CM: F31;  


Cytogenetic location: 18p     Genomic coordinates (GRCh38): 18:1-18,500,001


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
18p {Major affective disorder 1} 125480 Autosomal dominant 2

TEXT

Bipolar affective disorder is a genetically heterogeneous complex trait. One susceptibility locus for bipolar disorder, MAFD1, has been mapped to chromosome 18p.

Other mapped loci include MAFD2 (309200) on chromosome Xq28, MAFD3 (609633) on chromosome 21q22, MAFD4 (611247) on chromosome 16p12, MAFD7 (612371) on chromosome 22q12, MAFD8 (612357) on chromosome 10q21, and MAFD9 (612372) on chromosome 12p13.

Epistatic interaction has been postulated for loci designated MAFD5 (611535) on chromosome 2q22-q24 and MAFD6 (611536) on chromosome 6q23-q24.


Clinical Features

Depressive disorders represent a prevalent (1 to 2%) and major illness characterized by episodes of dysphoria that are associated with somatic symptoms. It may have a manic-depressive (bipolar) or purely depressive (unipolar) course. The most characteristic features of bipolar affective disorder (BPAD) are episodes of mania (bipolar I, BP I) or hypomania (bipolar II, BP II) interspersed with periods of depression (Goodwin and Jamison, 1990). If untreated, manic-depressive illness is associated with a suicide rate of approximately 20%.


Biochemical Features

Wright et al. (1984) studied binding of radioiodine-labeled hydroxybenzylpindolol to beta-adrenoceptors in lymphoblastoid cell lines from members of 5 families affected by manic-depressive disorder. Binding was reduced to less than half of control values in cell lines from 4 of 6 patients with manic-depressive disorder and only 1 of 18 unaffected relatives or controls. All the cell lines with reduced binding came from 3 families; members of 2 remaining families showed normal binding. The findings were interpreted as indicating genetic heterogeneity in manic-depressive disorder and a role played by a beta-adrenoceptor defect in genetic susceptibility to the disorder in some cases.

O'Reilly et al. (1994) presented a 2-generation family in which 8 members met DSM-III-R criteria for major depression. Four of the affected individuals failed to respond to standard therapeutic doses of tricyclic and new generation antidepressants but subsequently responded to the monoamine oxidase inhibitor tranylcypromine. O'Reilly et al. (1994) suggested that families demonstrating preferential response to a particular psychotropic drug may be a more homogeneous group in which to perform linkage analysis. Grof et al. (1994) also suggested that a helpful criterion for selecting more homogeneous groups of patients for linkage analysis would be the presence of a response to specific treatment. To that end, they studied 121 probands with primary affective disorders and 903 first-degree relatives and spouses. Seventy-one probands responded to lithium treatment and 50 were nonresponders. Study of the first-degree relatives of lithium responders revealed that 3.8% had bipolar disorder, whereas none of the relatives of the nonresponders were so affected. Schizophrenia was more common in the families of nonresponders (2.4% vs 0.3%).

In studies of the effect of lithium on Xenopus morphogenesis, Klein and Melton (1996) determined that lithium acts through inhibition of glycogen synthase kinase 3-beta (GSK3B; 605004) and not through inhibition of inositol monophosphatase (602064). They suggested that their observations may provide insights into the pathogenesis and treatment of bipolar disorder.

Using positron emission tomographic (PET) images of cerebral blood flow and rate of glucose metabolism to measure brain activity, Drevets et al. (1997) localized an area of abnormally decreased activity in the prefrontal cortex ventral to the genu of the corpus callosum in both familial bipolar depression patients and familial unipolar depression patients. The decrement in activity was at least partly explained by a corresponding reduction in cortical volume, as demonstrated by magnetic resonance imaging (MRI). This region had previously been implicated in the mediation of emotional and autonomic responses to socially significant or provocative stimuli, and in the modulation of the neurotransmitter systems targeted by antidepressant drugs.

Williams et al. (2002) demonstrated that lithium, carbamazepine, and valproic acid, all drugs used to treat bipolar affective disorder, inhibit the collapse of sensory neuron growth cones and increase growth cone area. These effects do not depend on glycogen synthase kinase-3 (see 606784) or histone deacetylase (see 601241) inhibition. Inositol, however, reverses the effects of the drugs on growth cones, thus implicating inositol depletion in their action. Moreover, the development of Dictyostelium is sensitive to lithium and to valproic acid, but resistance to both is conferred by deletion of the gene that codes for prolyl oligopeptidase (600400), which also regulates inositol metabolism. Inhibitors of prolyl oligopeptidase reversed the effects of all 3 drugs on sensory neuron growth cone area and collapse. Williams et al. (2002) concluded that their results suggest a molecular basis for both bipolar affective disorder and its treatment.


Inheritance

The role of genetic factors in bipolar disorder is indicated by concordance in monozygotic and dizygotic twins, respectively, of 57% and 14%, and the correlation between adopted persons and their biologic relatives (Cadoret, 1978).

Brent and Mann (2005, 2006) noted that in studies of adopted children and twins, familial concordance for suicidal behavior is explained by both genetic and environmental factors. Suicidal behavior that begins before 25 years of age is highly familial, and having a greater number of affected family members is associated with an earlier age.

Craddock and Sklar (2013) noted that strong evidence exists for a polygenic contribution to risk of bipolar disorder (i.e., many risk alleles of small effect). They stated that most cases of bipolar disorder involve the interplay of several genes or more complex genetic mechanisms, together with the effects of nongenetic (environmental) risk factors and stochastic factors.

Anticipation

In 34 families unilineal for bipolar affective disorder, McInnis et al. (1993) compared age of onset and disease severity between 2 generations. They found that the second generation experienced onset 8.9 to 13.5 years earlier and illness 1.8 to 3.4 times more severe than did the first generation. They concluded that genetic anticipation was demonstrated and they suggested that genes with expanding trinucleotide repeats might be involved in the genetic etiology of the disorder.

Gelernter (1995) reviewed the genetics of BPAD and other behavioral disorders and discussed specifically the difficulties associated with possible anticipation in bipolar affective disorder as discussed by McInnis et al. (1993). McInnis (1996) traced the history of anticipation. He suggested that it has its roots in the social chaos of the French Revolution, as well as in the medical observations of the French 'alienist' (an early term for psychiatrist) (Morel, 1857) and in the theories of atavism advocated by the Italian psychiatrist of the late 19th century (Lombroso, 1911). As to the validity of anticipation in bipolar disorder and schizophrenia, McInnis (1996) quoted Hodge and Wickramaratne (1995) as concluding that 'in psychiatric disorders, bias of ascertainment is pervasive and that there is no simple way to circumvent it.'

Parent-of-Origin Effect

McMahon et al. (1995) tested for parent-of-origin effect on the transmission of bipolar affective disorder, which might reflect either imprinting or mitochondrial inheritance. They examined the frequency and risk of BPAD among relatives in a sample of 31 families ascertained through treated probands with BPAD and selected for the presence of affected phenotypes in only 1 parental lineage. They observed a higher than expected frequency of affected mothers (p less than 0.04), a 2.3- to 2.8-fold increased risk of illness for maternal relatives (P less than 0.006), and a 1.3- to 2.5-fold increased risk of illness for the offspring of affected mothers (p less than 0.017). In 7 large pedigrees, fathers repeatedly failed to transmit the affected phenotype to daughters or sons. Taken together, these findings were interpreted as indicating a maternal effect in the transmission in BPAD susceptibility and suggested that molecular studies of mtDNA and imprinted DNA are warranted in patients with BPAD.

A number of family studies had reported increased morbid risk to the mothers, relative to the fathers, of probands with bipolar affective disorder. An excess of mother-offspring pairs had also been reported. These observations suggested that bipolar affective disorder may be caused by mitochondrial DNA mutations. Kirk et al. (1999) sequenced the mitochondrial genome in 25 bipolar patients with family histories of psychiatric disorder that suggested matrilineal inheritance. No polymorphism identified more than once in this sequencing showed any significant association with bipolar affective disorder in association studies using 94 cases and 94 controls. To determine whether their sample of patients showed evidence of selection against the maternal lineage, Kirk et al. (1999) determined genetic distances between all possible pairwise comparisons within the bipolar and control groups, based on multilocus mitochondrial polymorphism haplotypes. These analyses revealed fewer closely related haplotypes in the bipolar group than in the matched control group, suggesting selection against maternal lineages in this disease. Such selection was considered compatible with recurrent mitochondrial mutations, which are associated with slightly decreased fitness.


Mapping

Linkage to Chromosome 18 (MAFD1)

In the course of a systematic genomic survey, Berrettini et al. (1994) examined 22 manic-depressive (bipolar) families for linkage to 11 chromosome 18 pericentromeric marker loci, under dominant and recessive models. Overall lod score analysis for the pedigrees was not significant under either model, but several families yielded scores consistent with linkage under dominant or recessive models. Affected sib pair analysis of these data yielded evidence for linkage (p less than 0.001) with D18S21. Affected pedigree member analysis also suggested linkage. The results were interpreted as suggesting a susceptibility gene in the pericentromeric region of chromosome 18, with a complex mode of inheritance.

Pursuant to the report by Berrettini et al. (1994) of a BPAD susceptibility locus on chromosome 18 and the report of a parent-of-origin effect by McMahon et al. (1995), Stine et al. (1995) undertook a linkage study in 28 nuclear families selected for apparent unilineal transmission of the BPAD phenotype. They used 31 polymorphic markers spanning chromosome 18. The study was distinguished by relatively small, densely affected families with apparent unilineal transmission and direct clinical evaluation of family members by psychiatrists. Evidence for linkage was tested with affected-sib-pair and lod score methods under 2 definitions of the affected phenotype. The affected-sib-pair analyses indicated excess allele sharing for markers on 18p within the region reported previously. The greatest sharing was at D18S37. In addition, excess sharing of the paternally, but not maternally, transmitted alleles was observed at 3 markers on 18q. The evidence for linkage to loci on both 18p and 18q was strongest in the 11 paternal pedigrees, i.e., those in which the father or one of the father's sibs was affected. The results were interpreted by Stine et al. (1995) as providing further support for linkage of BPAD to chromosome 18.

Prompted by the report by Berrettini et al. (1994) of a bipolar susceptibility locus in the region of the centromere of chromosome 18, Pauls et al. (1995) studied markers from this region in Old Order Amish families. Although linkage findings were replicated in 1 previously studied Amish pedigree containing 4 affected individuals, linkage to this region was excluded in the larger sample. Pauls et al. (1995) concluded that if a susceptibility locus for bipolar disorder is located in this region of chromosome 18, it is of minor significance among the Amish.

In a genetically isolated population of the central valley of Costa Rica (CVCR), Freimer et al. (1996) undertook a linkage study of severe bipolar disorder in 2 pedigrees. The 2.6 million residents of the CVCR are descended mainly from a small group of Spanish and Amerindian founders who lived in the 16th and 17th centuries; by the beginning of the 18th century, the CVCR had a single population that then grew rapidly, without subsequent immigration, for almost 200 years (Escamilla et al., 1996). Freimer et al. (1996) found strongest evidence for a specific locus on chromosome 18q22-q23 where 7 of 16 markers yielded peak lod scores over 1.0. This localization was supported by marker haplotypes shared by 23 of 26 affected individuals studied. As a continuation of that study, McInnes et al. (1996) performed a complete genome screen for genes predisposing to severe bipolar disorder. They considered as affected only individuals with bipolar mood disorder and screened the genome for linkage with 473 microsatellite markers. They used a model for linkage analysis that incorporated a high phenocopy rate and a conservative estimate of penetrance. They suggested on the basis of their results that 18q, 18p, and 11p deserve further study; in these regions suggestive lod scores were observed for 2 or more contiguous markers. Isolated lod scores that exceeded the threshold in 1 or both families studied also occurred on 10 other chromosomes. Additional linkage studies on 2 extended BP I pedigrees from the CVCR implicated a candidate region on 18p11.3 (Escamilla et al., 1999).

Knowles et al. (1998) could find no evidence for significant linkage between bipolar affective disorder and chromosome 18 pericentromeric markers in a large series of multiplex extended pedigrees. This was one of the largest samples reported to date: 1,013 genotyped individuals in 53 unilineal multiplex pedigrees. Ten highly polymorphic markers and a range of parametric and nonparametric analyses were used. Not only was there no evidence for linkage, but there was also no evidence for significant parent-of-origin effect.

McInnes et al. (2001) further investigated the 18p11.3 region by creating a physical map and developing 4 new microsatellite and 26 single-nucleotide polymorphism (SNP) markers for typing in the Costa Rican pedigree and population samples. The results of fine-scale association analyses in the population sample, as well as evaluation of haplotypes in 1 of the large pedigrees, suggested a candidate region containing 6 genes but also highlighted the complexities of linkage disequilibrium mapping of common disorders.

To clarify the issue of genetic linkage between bipolar affective disorder and 18q, McMahon et al. (2001) analyzed the relationship between clinical features and allele sharing. Relatives ascertained through a proband who had BP I disorder were interviewed by a psychiatrist, assigned an all-sources diagnosis, and genotyped with 32 markers on 18q21-q23. The authors found that paternal allele sharing on 18q21 was significantly associated with a diagnostic subtype, and was greatest in pairs where both sibs had BP II. Paternal allele sharing across 18q21-q23 was also significantly greater in families with at least 1 sib pair in which both had BP II. In these families, multipoint affected sib-pair linkage analysis produced a peak paternal lod score of 4.67 versus 1.53 in all families. Thus, affected sib pairs with BP II discriminated between families who showed evidence of linkage to 18q and families who did not. Families with a BP II sib pair produced an increased lod score and improved linkage resolution. These findings strengthened the evidence of genetic linkage between BPAD and 18q, and provided preliminary support for BP II as a genetically valid subtype of BPAD. Bipolar type II disorder is characterized by hypomania that is so brief or so slight as to cause no significant problems in functioning. BP I, on the other hand, is the diagnosis attached to anyone with a significantly problematic manic state (extreme symptoms of grandiosity, poor social judgment or functional impairment due to distractibility at work, etc.).

Linkage to Chromosome 2q22-q24

See MAFD5 (611535) for a discussion of linkage of susceptibility to bipolar disorder to chromosome 2q22-q24.

Linkage to Chromosome 3p

Etain et al. (2006) conducted a genomewide search with 384 microsatellite markers using nonparametric linkage (NPL) analysis in 87 sib pairs ascertained as part of the European Collaborative Study of Early Onset Bipolar Affective Disorder. Early-onset patients (age at onset of 21 years or below) were studied because age of onset may help to define homogeneous bipolar affective disorder subtypes. The 3p14 region showed the most significant linkage in the first phase of analysis with an NPL score of 3.51. Additional linkage analysis with increased marker density revealed an NPL score of 3.83 at chromosome 3p14.

Linkage to Chromosome 4

Blackwood et al. (1996) carried out a linkage study in Scotland in 12 bipolar families. In a single family, a genome search using 193 markers indicated linkage on 4p where D4S394 generated a 2-point lod score of 4.1 under a dominant model of inheritance. With 3-point analyses using neighboring markers, they obtained a maximum lod score of 4.8. (Eleven other bipolar families were typed using D4S394 and in all families combined there was evidence of linkage with heterogeneity with a maximum 2-point lod score of 4.1 (theta = 0.0; alpha = 0.35).)

Ginns et al. (1998) reported a different approach to linkage study of BPAD in the Old Order Amish. To determine whether there could be protective alleles that prevent or reduce the risk of developing BPAD, similar to what is observed in other genetic disorders, they used 'mental health wellness' (absence of any psychiatric disorder) as the phenotype in their genomewide linkage scan of several large multigeneration Old Order Amish pedigrees exhibiting a high incidence of BPAD. They found strong evidence for linkage of mental health wellness to a locus on 4p, designated MHW1 (603663), at D4S2949; maximum nonparametric linkage score = 4.05, p = 5.22 x 10(-4). They also found suggestive evidence for a locus on 4q, designated MHW2 (603664), at D4S397; maximum nonparametric linkage score = 3.29, p = 2.57 x 10(-3). Findings were consistent with the hypothesis that certain alleles can prevent or modify the clinical manifestations of BPAD and perhaps other related affective disorders.

Ekholm et al. (2003) performed a genomewide scan for susceptibility loci in bipolar disorder in 41 Finnish families with at least 2 affected sibs. They identified a distinct locus on 16p12 (see MAFD4, 611247) and observed 3 additional loci with a 2-point lod score greater than 2.0, at markers on 4q32, 12q23, and Xq25. After fine mapping these chromosomal regions and genotyping additional family members, 4q32 provided significant evidence of linkage for the 3-point analyses (maximum lod = 3.6 between D4S3049 and D4S1629).

Linkage to Chromosome 5

Garner et al. (2001) used an algorithm that permitted nonparametric linkage analysis of large, complex pedigrees with multiple inbreeding loops to reanalyze the genome-screen data from the Costa Rican kindred segregating severe bipolar disorder (Freimer et al., 1996; McInnes et al., 1996). The results were consistent with previous linkage findings on chromosome 18 and also suggested a novel locus on chromosome 5 that was not identified using traditional linkage analysis.

Hong et al. (2004) performed linkage analysis using 74 individuals from the Costa Rican pedigree and found evidence for a 3.2-Mb region between markers D5S1480 and D5S2090 on chromosome 5q31-q33. The authors suggested that conflicting haplotype data reflected incomplete penetrance, phenocopies, or locus/allelic heterogeneity. The authors noted that Freimer et al. (1996) described a conserved haplotype on 18q22-q23 in the same kindred. Hong et al. (2004) found that 12 of 20 affected individuals shared both haplotypes, suggesting that both loci are important in conferring disease risk.

Coon et al. (1993) carried out an extensive linkage analysis in 8 moderate-sized families with manic-depression. When autosomal dominant inheritance was assumed, 273 DNA markers gave lod scores less than -2.0 at theta = 0.0, 174 DNA loci produced lod scores less than -2.0 at theta = 0.05, and 4 DNA marker loci yielded lod scores greater than 1. Of the markers giving lod scores greater than 1, only D5S62 continued to show evidence for linkage when the affected-pedigree-member method was used. D5S62 maps to distal 5q, a region containing neurotransmitter receptor genes for dopamine (e.g., 126449), gamma-aminobutyric acid (e.g., 137160, 137164), glutamate (e.g., 138248), and norepinephrine (e.g., 109690, 104219, 104220).

Linkage to Chromosome 6p

Smeraldi et al. (1978) first suggested linkage between HLA on chromosome 6p21.3 and affective disorders on the basis of the finding that pairs of affected sibs shared HLA haplotypes more often than would be predicted by chance. Weitkamp et al. (1981) likewise found evidence of a susceptibility gene or genes linked to HLA. Neither group subdivided the depressive disorders into bipolar and unipolar subtypes. Stronger evidence of linkage might be found in 1 subtype, or it may turn out that both are linked to HLA, suggesting that they are different forms of the same illness. One of Weitkamp's study families was that reported earlier by Pardue (1975)--in fact, Pardue's own kindred (Wingerson, 1982). Weitkamp et al. (1981) found that HLA haplotype identity in pairs of affected sibs and in pairs of unaffected older sibs deviated markedly from expected (p less than 0.005). Perhaps surprisingly, no increase in HLA haplotype identity was found in sibships with more than 2 affected members. When parents had a difference in load of genes for susceptibility (as estimated by the occurrence of affective illness in themselves and their relatives), HLA haplotypes were randomly transmitted to unaffected or affected children from the affected, 'high-load' parent, but not randomly from the unaffected, 'low-load' parent (p less than 0.001), suggesting a recessive effect, i.e., greater chance of illness in homozygotes.

Stancer et al. (1988) published data apparently confirming the relationship between HLA and manic depression. When combined with their previous data, the total number of families analyzed was 117. As in the previous study, the increase in HLA haplotype sharing over random expectation was greater if 'high-load' sibships, i.e., sibships with 3 or more affected sibs, were omitted from the analysis. Weitkamp (1981, 1983) suggested that the extent of HLA haplotype sharing among affected sib pairs should decrease as the number of parental HLA haplotypes containing susceptibility genes increases from 1 to 4. Thus, he reasoned that there may actually be less HLA haplotype sharing among sibs when the parents have maximum genetic susceptibility ('high load') compared with families in which the genetic susceptibility that could be contributed by either parent is limited to the genes in 1 of the 2 HLA haplotypes in that parent. If an increased number or variety of affective disorder susceptibility genes in a person results in a greater probability of illness, then nuclear families with a higher proportion of affected family members are likely to have a greater number or variety of affective disorder susceptibility genes than families with a low proportion of affected members. Weitkamp and Stancer (1989) suggested that the HLA effect may be greater in unipolar than in bipolar disorders and more apparent in families with few affected members than in 'high-load' families.

Schulze et al. (2004) extended the study of Dick et al. (2003) to test for robustness of the linkage to differing analysis methods, genotyping error, and gender-specific maps; for parent-of-origin effects; and for interaction with markers within the schizophrenia linkage region on chromosome 6p (see SCZD3; 600511). Members of 245 families ascertained through a sib pair affected with bipolar I or schizoaffective-bipolar disorder were genotyped with 18 markers spanning chromosome 6, and nonparametric linkage analysis was performed. Linkage to 6q was robust to analysis methods, gender-specific map differences, and genotyping error. The locus conferred a 1.4-fold increased risk. Affected sibs shared the maternal more often than paternal chromosome (p = 0.006), which could reflect a maternal parent-of-origin effect. There was a positive correlation between family-specific linkage scores on 6q and those on 6p22.2 (p less than 0.0001). Linkage analysis for each locus conditioned on evidence of linkage to the other increased the evidence for linkage at both loci (p less than 0.0005). Lod scores increased from 2.26 to 5.42 on 6q and from 0.35 to 2.26 on 6p22.2. The results supported linkage of bipolar disorder to 6q, revealed a maternal parent-of-origin effect, and demonstrated an interaction of this locus with a locus on chromosome 6p22.2 linked to schizophrenia.

Linkage to Chromosome 6q22

Middleton et al. (2004) performed a linkage analysis on 25 extended multiplex Portuguese families, including 12 families previously reported by Pato et al. (2004), segregating for bipolar disorder using a high-density SNP genotyping assay with a 0.21-Mb intermarker spacing. The analysis revealed genomewide significance with a maximum NPL of 4.20 and a maximum lod score of 3.56 at 6q22 (125.8 Mb).

Linkage to Chromosome 8

Ophoff et al. (2002) performed a genomewide association study of severe bipolar disorder in the patients from the central valley of Costa Rica. They observed LD with severe bipolar disorder on several chromosomes; the most striking results were in proximal 8p, a region that had previously shown linkage to schizophrenia. Ophoff et al. (2002) suggested that this region could be important for severe psychiatric disorders rather than for a specific phenotype.

Cichon et al. (2001) conducted a complete genome screen with 382 markers in a sample of 75 BPAD families of German, Israeli, and Italian origin. Parametric and nonparametric linkage analysis was performed. The highest 2-point lod score was obtained on 8q24 (D8S514; lod score = 3.62), and the authors confirmed a putative locus on 10q25-q26 (D10S217; lod score = 2.86). By analyzing the autosomal genotype data, putative paternally imprinted loci were identified in chromosomal regions 2p24-p21 and 2q31-q32; maternally imprinted susceptibility genes may be located on 14q32 and 16q21-q23.

Park et al. (2004) genotyped 373 individuals from 40 extended pedigrees with high density bipolar disorder and found evidence for significant linkage for psychotic bipolar disorder (genomewide p less than 0.05) to chromosomes 9q31 (lod = 3.55) and 8p21 (lod = 3.46). Nine other sites obtained lod scores supportive of linkage. The highest lod scores occurred in the subgroup of families with the largest concentration of psychotic individuals. Seven of the loci identified in this study had previously been implicated in schizophrenia, suggesting that psychosis is a potentially useful phenotype in bipolar disorder for genetic studies.

Linkage to Chromosome 9

Sherrington et al. (1994) performed linkage analysis on 5 multigenerational families with bipolar and unipolar affective disorder, using highly polymorphic microsatellite markers from the ABO-AK1-ORM region at 9q34. The dopamine beta-hydroxylase locus (223360) is also at 9q34 and was considered to be a candidate gene. Their analyses provided strong evidence against a major susceptibility allele in this region, in contradistinction to the findings of Hill et al. (1988), Tanna et al. (1989), and Wilson et al. (1989, 1991).

Venken et al. (2005) conducted a genomewide scan to identify susceptibility loci for affective spectrum disorder (bipolar disorder and recurrent unipolar depression) in 9 families from an isolated population in Vasterbotten in northern Sweden. A region on chromosome 9q showed the highest 2-point and multipoint lod scores. A common ancestral haplotype was inherited by 18 of 21 patients from 3 families linked to 9q, which reduced the candidate region to 1.6 Mb on 9q31-q33. Further analysis identified the shared haplotype in 4.2% of 182 unrelated patients with bipolar disorder from the Vasterbotten isolate, but not in 182 control individuals. Venken et al. (2005) concluded that a susceptibility locus for affective disorder is located on chromosome 9q31-q33.

Park et al. (2004) genotyped 373 individuals from 40 extended pedigrees with high density bipolar disorder and found evidence for significant linkage for psychotic bipolar disorder (genomewide p less than 0.05) to chromosomes 9q31 (lod = 3.55) and 8p21 (lod = 3.46). Nine other sites obtained lod scores supportive of linkage. The highest lod scores occurred in the subgroup of families with the largest concentration of psychotic individuals. Seven of the loci identified in this study had previously been implicated in schizophrenia, suggesting that psychosis is a potentially useful phenotype in bipolar disorder for genetic studies.

Linkage to Chromosome 10q21

Ferreira et al. (2008) tested 1.8 million variants in 4,387 cases of bipolar disorder and 6,209 controls from 3 independent samples and identified a region of strong association with SNP rs10994336 in the ankyrin G gene (ANK3; 600465) on chromosome 10q21, with a p value of 9.1 x 10(-9). See MAFD8 (612357).

Linkage to Chromosome 11

A form of manic-depressive disorder in the Old Order Amish of Lancaster County, Pennsylvania, was thought by Egeland et al. (1987) to be tightly linked to INS (176730) and HRAS1 (190020). In linkage studies using RFLPs related to these genes on the tip of 11p, the maximum lod score was 4.5 at theta = 0.0. Of interest is the description by Joffe et al. (1986) of cosegregation of thalassemia and affective disorder in a non-Amish pedigree. Egeland et al. (1987) suggested that the tyrosine hydroxylase (TH; 191290) gene, which maps to 11p, should be considered as a candidate gene because this enzyme catalyzes an important step in the dopamine synthesis pathway. Gill et al. (1988) ruled out tight linkage between manic-depressive psychosis and the 11p markers HRAS1 and INS, however. Two other groups failed to find linkage of 11p markers to manic-depressive illness (Neiswanger et al., 1990). From another extension of the study of the original Amish pedigree, Pauls et al. (1991) likewise excluded linkage to 11p markers. Pakstis et al. (1991) found no evidence of linkage after screening 185 marker loci in the Old Order Amish. They estimated that roughly 23% of the autosomal genome had been excluded. Law et al. (1992) determined the INS and HRAS1 genotypes of 81 persons in this pedigree and excluded that region of chromosome 11 as the site of the gene, which they symbolized BAD (for bipolar affective disorder).

In 5 Icelandic pedigrees, Holmes et al. (1991) could find no evidence of linkage of manic depression to the dopamine D2 receptor (DRD2; 126450) or other markers in its vicinity on 11q.

Linkage to Chromosome 12p13

Ferreira et al. (2008) tested 1.8 million variants in 4,387 cases of bipolar disorder and 6,209 controls from 3 independent samples and identified association with SNP rs1006737 in the CACNA1C gene (114205) on chromosome 12p13, with a p value of 7.0 x 10(-8). See MAFD9 (612372).

Linkage to Chromosome 12q

By linkage analysis in 2 Danish families with bipolar affective disorder, Ewald et al. (1998) found that the microsatellite marker D12S1639 gave a significant lod score of 3.37. Earlier, Craddock et al. (1994) had suggested linkage between affective disorder and Darier disease (124200), which maps to 12q23-q24.1. Linkage results from independent studies in Canadian families (Morissette et al., 1999) supported the existence of a susceptibility locus on 12q23-q24. To take advantage of isolated populations for genetic mapping and disease gene identification, Degn et al. (2001) investigated a possible chromosomal segment shared among distantly related patients with bipolar affective disorder in the Faroe Islands, using 17 microsatellite markers covering 24 cM in the 12q24 region. The region of most interest contained the primary region suggested by the previously reported haplotypes in the 2 Danish families studied by Ewald et al. (1998).

Ewald et al. (2002) reported a genomewide scan for risk genes involved in bipolar disorder in 2 Danish Caucasian families with affected members in several generations. Ewald et al. (2002) used 613 microsatellite markers in a 2-stage approach. Linkage was obtained at 12q24.3 (D12S1639) with a 2-point parametric lod score of 3.42 (empirical P-value 0.00004, genomewide P-value 0.0417) in both families tested. The multipoint parametric lod score at D12S1639 was 3.63 (genomewide P-value 0.0265). At 1p22-p21 (D1S216), a parametric, affecteds-only 2-point lod score of 2.75 (empirical P-value 0.0002, genomewide P-value 0.1622) was found. A 3-point lod score of 2.98 (genomewide p value = 0.1022) was found at D1S216, and a multipoint nonparametric analysis yielded a maximum nonparametric linkage (NPL)-all score of 17.60 (p value = 0.00079) at D1S216.

In 2 cohorts of patients with bipolar affective disorder from Germany and Russia totaling 883 patients and 1,300 controls, Cichon et al. (2008) observed an association between disease and the minor alleles of 3 SNPs in haplotype 1 of the TPH2 gene on chromosome 12q21, (rs11178997, rs11178998, and rs7954758; odds ratio of 1.6, p value of 0.00073). Haplotype 1 covers part of the 5-prime regulatory region and exons 1 and 2 of the TPH2 gene. Cichon et al. (2008) also observed an association between bipolar disorder and a nonsynonymous SNP in the TPH2 gene (P206S; 607478.0003).

Linkage to Chromosome 16

See MAFD4 (611247) for a discussion of linkage of susceptibility to bipolar disorder to chromosome 16p12.

Linkage to Chromosome 17

Dick et al. (2003) performed genomewide linkage analyses on 1,152 individuals from a new sample of 250 families segregating for bipolar disorder and related affective illnesses, ascertained at 10 sites in the United States through a proband with BP I affective disorder and a sib with BP I or schizoaffective disorder, bipolar type. Suggestive evidence for linkage was found on chromosome 17q (peak maximum lod score = 2.4) at marker D17S928, and on 6q (peak maximum lod score = 2.2) near marker D6S1021. Suggestive evidence of linkage was observed in 3 other regions, on chromosomes 2p, 3q, and 8q. This study, based on a linkage sample for bipolar disorder larger than any previously analyzed, indicated that several genes contribute to bipolar disorder.

Linkage to Chromosome 20

In 9 Australian pedigrees, Le et al. (1994) excluded close linkage of bipolar disorder to the gene encoding the alpha subunit of the stimulatory form of G protein (139320), previously mapped to chromosomal region 20q13.2.

Radhakrishna et al. (2001) studied a large Turkish pedigree segregating apparently autosomal dominant BPAD, which contained 13 affected individuals. The age of onset ranged from 15 to 40 years with a mean age of 25 years. The phenotypes consisted of recurrent manic and major depressive episodes, including suicide attempts. There was usually full remission with lithium treatment. A genotyping of 230 highly informative polymorphic markers throughout the genome and subsequent linkage analysis using a dominant mode of inheritance showed strong evidence for a BPAD susceptibility locus on chromosome 20p11.2-q11.2. The highest 2-point lod score of 4.34 (theta = 0.0) was obtained with markers D20S604, D20S470, D20S836, and D20S838 (100% penetrance). Haplotype analysis using informative recombinants enabled the mapping of the BPAD locus in this family between markers D20S186 and D20S109 in a region of approximately 42 cM. The authors noted that the chromosome 20 BPAD susceptibility locus had not been identified in previous studies of common 'polygenic' small pedigrees, which could be explained by an absence of common deleterious mutations of the chromosome 20 BPAD locus in those pedigrees and/or by the presence of a severe mutation in the Turkish pedigree that by itself confers susceptibility to BPAD.

Linkage to Chromosome 21q22

In a preliminary genome screen of 47 bipolar disorder families, Straub et al. (1994) detected one in which a lod score of 3.41 was demonstrated for linkage with the PFKL (171860) locus on 21q22.3. Largely positive lod scores were obtained also with 14 other markers in 21q22.3 in this family. In a linkage analysis with an 'affecteds-only' method, Aita et al. (1999) found linkage to the 21q22 region, corroborating the findings of the earlier study by the same group (Straub et al., 1994).

Following up on the work of Straub et al. (1994) suggesting a susceptibility locus for bipolar affective disorder on the long arm of chromosome 21, Smyth et al. (1997) studied 23 multiply affected pedigrees collected from Iceland and the U.K., using the markers PFKL, D21S171, and D21S49. Positive lod scores were obtained with 3 Icelandic families. Affected sib pair analysis demonstrated increased allele sharing. The same set of pedigrees had previously been typed for a tyrosine hydroxylase gene (TH; 191290) polymorphism at 11p15 and had shown some evidence for linkage. When information from TH and the 21q markers was combined in a 2-locus admixture analysis, an overall admixture lod of 3.87 was obtained using the bipolar affection model. Thus the data of Smyth et al. (1997) were compatible with the hypothesis that a locus at or near TH influences susceptibility of some pedigrees, while a locus near D21S171 is active in others.

Age at onset (AAO) is a potential clinical marker of genetic heterogeneity in BP (Bellivier et al., 2001). Rates of comorbidity and clinical indicators of severity (e.g., suicide attempt) vary across different AAO subgroups, and AAO subgroups aggregate in families such that affected relatives typically have similar AAOs. Therefore, Lin et al. (2005) sought to incorporate AAO as a covariate in linkage analysis of BP using 2 different methods in genomewide scans of 150 multiplex pedigrees with 874 individuals. The LODPAL analysis identified 2 loci: one on 21q22.13 (MAFD3; 609633) and the other on 18p11.2 (MAFD1) for early onset (AAO = 21 years or younger) and later onset (AAO = older than 21 years), respectively. The finding on 21q22.13 was significant at the chromosome-wide level, even after correction for multiple testing. Moreover, a similar finding was observed in an independent sample of 65 pedigrees (lod = 2.88). The finding on 18p11.2 was only nominally significant and was not observed in the independent sample. However, 18p11.2 emerged as one of the strongest regions in the ordered-subset analysis (OSA) with a lod of 2.92, in which it was the only finding to meet chromosomewide levels of significance after correction for multiple testing. These results suggested that 21q22.13 and 18p11.2 may harbor genes that increase the risks for early-onset and later-onset forms of BP, respectively. Lin et al. (2005) suggested that previous inconsistent linkage findings may have been due to differences in the AAO characteristics of the samples examined.

Linkage to Chromosome 22q12

For information on linkage of susceptibility to bipolar disorder to chromosome 22q12, see MAFD7 (612371) and Kelsoe et al., 2001.

Other Genomewide Linkage Studies

In an 'Old Order Amish revisited' study, Ginns et al. (1996) performed a genomewide linkage analysis in the Lancaster County group. In addition to the so-called pedigree 110, which was used for reporting the original genetic linkage data by Egeland et al. (1987), 2 pedigrees closely related to pedigree 110 and 2 other pedigrees, 210 and 310, were studied; all 5 pedigrees traced back to a founder couple who immigrated to the U.S. around 1750. The diagnoses were broken down into BP I (bipolar disorder with mania) and BP II (bipolar disorder with hypomania). Ginns et al. (1996) found evidence that regions on chromosomes 6, 13, and 15 harbor susceptibility loci for bipolar affective disorder, suggesting to them that bipolar affective disorder in the Old Order Amish is inherited as a complex trait.

LaBuda et al. (1996) reported progress of a full genome screen for loci predisposing to affective disorder in the Old Order Amish. To the previously reported lod score results published by Gerhard et al. (1994), they added lod score results for an additional 367 markers distributed throughout the genome, along with allele- and haplotype-sharing analyses on those chromosomes sufficiently saturated with markers. No statistically significant lod scores resulted. Some degree of allele sharing was found at 74 loci, and 3.8% of all markers analyzed passed more stringent significance criteria suggestive of linkage. Although genomic areas were highlighted for further exploration, the studies of LaBuda et al. (1996) identified no region clearly involved in the etiology of affective disorder in this population.

Risch and Botstein (1996) reviewed 19 linkage studies in manic-depressive illness; the studies purported to identify loci on 10 different autosomes, including both the short arm and the long arm of chromosome 18; linkage in 3 different regions of distal Xq had been proposed.

Segurado et al. (2003) applied the rank-based genome scan metaanalysis (GSMA) method (Levinson et al., 2003) to 18 bipolar disorder genome scan datasets in an effort to identify regions with significant support for linkage in the combined data. No region achieved genomewide statistical significance by several simulation-based criteria. The most significant p values (less than 0.01) were observed on chromosomes 9p22.3-p21.1, 10q11.21-q22.1, and 14q24.1-q32.12. Nominally significant p values were observed in several other chromosomal regions.

In a study of Ashkenazi Jewish families, Fallin et al. (2004) identified 4 regions suggestive of linkage to bipolar disorder on chromosomes 1, 3, 11, and 18.

Pato et al. (2004) conducted a genomewide scan of 16 extended families from a Portuguese genetic isolate with bipolar disorder and identified 3 regions on chromosomes 2, 11, and 19 with genomewide suggestive linkage and several other regions, including chromosome 6q, that approached suggestive levels of significance. This research replicated the finding of an elevated lod score near marker D6S1021 on chromosome 6q (peak NPL at D6S1021 = 2.02; p = 0.025). Higher density mapping provided additional support for this locus (NPL = 2.59; p = 0.0068) and another marker, D6S1639 (NPL = 3.06; p = 0.0019). On chromosome 11, linkage was found to D11S1883 (NPL = 3.15; p = 0.0014).

Middleton et al. (2004) performed a linkage analysis on 25 extended multiplex Portuguese families, including 12 families previously reported by Pato et al. (2004), segregating for bipolar disorder using a high-density SNP genotyping assay with a 0.21-Mb intermarker spacing. The analysis revealed genomewide significance with a maximum NPL of 4.20 and a maximum lod score of 3.56 at 6q22 (125.8 Mb). Several other areas had suggestive linkage: 2 regions on chromosome 2 (57 Mb, NPL = 2.98; 145 Mb, NPL = 3.09), chromosome 4 (91 Mb, NPL = 2.97), chromosome 11 (45-68 Mb, NPL = 2.51), chromosome 16 (20 Mb, NPL = 2.89), and chromosome 20 (60 Mb, NPL = 2.99).

McQueen et al. (2005) hypothesized that combining original genotype data on linkage of bipolar disorder would provide benefits of increased power and control over sources of heterogeneity that outweigh the difficulty and potential pitfalls of the implementation. Thus, they conducted a combined analysis using the original genotype data from 11 bipolar disorder genomewide linkage scans comprising 5,179 individuals from 1,067 families. Heterogeneity among studies was minimized in the analyses by using uniform methods of analysis and a common, standardized marker map. They demonstrated that combining original genome-scan data is a powerful approach for the elucidation of linkage regions underlying complex disease. Their results established genomewide significant linkage to BP on chromosomes 6q and 8q, and provided solid information to guide future gene-finding efforts that rely on fine mapping and association approaches. McQueen et al. (2005) observed the most significant result for 'narrow' BP (BP type I-only phenotype) on chromosome 6q. When the analysis was expanded to include BP II, the linkage signal on 6q was attenuated, despite the increase in the number of affected relative pairs (ARPs). In contrast, removal of the individuals with BP II from the analysis reduced the evidence of linkage on 8q.

Maziade et al. (2005) performed a dense genome scan to identify susceptibility loci shared by schizophrenia and bipolar disorder. They used the same ascertainment, statistical, and molecular methods for 480 members from 21 multigenerational families from Eastern Quebec affected by schizophrenia, bipolar affective disorder, or both. Five genomewide significant linkages with maximized lod scores over 4.0 were observed: 3 for bipolar disorder (15q11.1, 16p12.3, 18q12-q21) and 2 for the shared 'common locus' phenotype (15q26, 18q12-q21). Nine maximized lod scores exceeded the suggestive threshold of 2.6: 3 for bipolar disorder (3q21, 10p13, 12q23), 3 for schizophrenia (6p22, 13q13, 18q21), and 3 for the combined locus phenotype (2q12.3, 13q14, 16p13). Maziade et al. (2005) noted that all of the linkage signals overlapped formerly reported susceptibility regions except the signal at 15q26.

Cheng et al. (2006) conducted a 9-cM genomewide scan in a large bipolar pedigree sample from the National Institute of Mental Health Genetics Initiative (1,060 individuals from 154 multiplex families). Parametric and nonparametric analyses using both standard diagnostic models and comorbid conditions thought to identify phenotypic subtypes were conducted. Genomewide significant linkage was observed on chromosomes 10q25, 10p12, 16q24, 16p13, and 16p12 using standard diagnostic models, and on 6q25 (suicidal behavior), 7q21 (panic disorder), and 16p12 (psychosis) using phenotypic subtypes. Several other regions were suggestive of linkage including 1p13 (psychosis), 1p21 (psychosis), 1q44, 2q24 (suicidal behavior), 2p25 (psychosis), 4p16 (psychosis, suicidal behavior), 5p15, 6p25 (psychosis), 8p22 (psychosis), 8q24, 10q21, 10q25 (suicidal behavior), 10p11 (psychosis), 13q32 and 19p13 (psychosis).

Kimmel et al. (2005) reported a large family in which bipolar disorder appeared to cosegregate with autosomal dominant medullary cystic kidney disease. Of the 7 members with kidney disease, 5 had bipolar I disorder, one had unipolar depression, and 1 had a hyperthymic phenotype. The authors noted that the 2 known loci of medullary cystic kidney disease are in regions of chromosome 1 (MCKD1; 174000) and 16 (MCDK2; 162000) had previously been linked to bipolar disorder and schizophrenia.

Exclusion Studies

Although corticotropin-releasing hormone (CRH; 122560) and its function in the hypothalamic-pituitary-adrenal axis had been implicated in depression (Stratakis and Chrousos, 1995), Stratakis et al. (1997) could demonstrate no linkage between the CRH gene and bipolar affective disorder.


Clinical Management

Genetic Variation in Lithium Response

Lithium has been a first-line choice for maintenance treatment of bipolar disorders to prevent relapse of mania and depression, but many patients do not have a response to lithium treatment. To discover genetic variation influencing response to lithium treatment, Chen et al. (2014) performed a discovery genomewide association study and 2 sets of replication in patients with bipolar I disorder from the Taiwan Bipolar Consortium who were receiving lithium treatment. Two SNPs in high linkage disequilibrium, and , located in the introns of GADL1 (615601) showed the strongest associations in the genomewide association study (p = 5.50 x 10(-37) and p = 2.52 x 10 (-37), respectively) and in the replication sample of 100 patients (p = 9.19 x 10(-15) for each SNP). These 2 SNPs had a sensitivity of 93% for predicting a response to lithium and differentiated between patients with a good response and those with a poor response in the follow-up cohort. Resequencing of GADL1 revealed a novel variant in GADL1 intron 8, IVS8+48delG, that is in complete linkage disequilibrium with rs17026688 and is predicted to affect splicing. These variants are rare in persons of European and African ancestry.

In a comment on the report of Chen et al. (2014), Birnbaum et al. (2014) stated that they found a very low level of GADL1 expression in the brain, and suggested that there was higher expression in the kidney; therefore, they concluded that the role of GADL1 is more likely related to taurine biosynthesis and kidney function than to brain function. Birnbaum et al. (2014) encouraged a retrospective review of kidney function and lithium levels in bipolar patients. Lee and Cheng (2014) responded to Birnbaum et al. (2014) that taurine may cross the blood-brain barrier to interact directly with the glutamate NMDA receptor, suggesting that the role of GADL1 in kidney function may be related to bipolar disorder.

Commenting on the report of Chen et al. (2014), Vlachadis et al. (2014) speculated that, given the magnitude of the association between the presence of the T allele and the response to lithium therapy, there might be a significant difference in the minimum efficacious serum lithium level between carriers and noncarriers of the 'response' allele. Lee and Cheng (2014) replied that they were unable to examine this issue in their retrospective study. Lee and Cheng (2014) also thanked Vlachadis et al. (2014) for identifying an error in Table 2 of their article. The odds ratio for the association between the presence of the T allele and a response to lithium therapy in the combined cohorts should be 88.5 (95% confidence interval, 41.4-198.0).

Ikeda et al. (2014) assessed 154 Japanese patients with bipolar disorder and did not observe an association for any criterion, even in a stringent phenotype analysis, as reported by Chen et al. (2014) for the rs17026688 allele. The Consortium on Lithium Genetics (2014) undertook a replication study in 218 samples that they collected from patients. Because the alleles reported by Chen et al. (2014) are common in Asians but rare in whites, the Consortium on Lithium Genetics (2014) studied only the Asian samples that they had obtained. In samples obtained from patients of Han Chinese or Japanese ancestry, the authors found no association between the variants and a response to lithium therapy at any threshold on the Alda scale. Lee and Cheng (2014) replied to Ikeda et al. (2014) and the Consortium on Lithium Genetics (2014) that the methods were not duplicated precisely, and offered to provide help with independent replication studies.

Anghelescu and Dettling (2014) questioned the finding of Chen et al. (2014) of better response to lithium therapy among patients with rapid cycling. Lee and Cheng (2014) replied that the relationship between rapid cycling and the effect of lithium is controversial. Anghelescu and Dettling (2014) also recommended that, based on the assumption that lack of efficacy leads to premature termination of therapy, genetic analysis be extended to include patients who terminated therapy. Lee and Cheng (2014) stated that because adherence to lithium maintenance treatment involves such factors as illness behavior and cognitive function, rapport with psychiatrists, serious adverse effects of treatment, and familial and financial support, the effect of lithium in patients with premature termination cannot be assessed.


Molecular Genetics

Association with the SLC6A3 Gene on Chromosome 5p15

Greenwood et al. (2001) reported evidence for an association between the DAT1 gene (SLC6A3; 126455) and bipolar disorder in a sample of 50 parent-proband trios. Using the transmission disequilibrium test (TDT), they showed an association between a haplotype composed of 5 SNPs in the 3-prime region of the DAT1 gene, exon 9 through exon 15, and bipolar disorder (allele-wise TDT empirical P = 0.001; genotype-wise TDT empirical P = 0.0004). Greenwood et al. (2006) analyzed a total of 22 SNPs in the 50 previously studied parent-proband trios and an independent set of 70 parent-proband trios. Using TDT analysis, an intron 8 SNP and an intron 13 SNP were found to be moderately associated with bipolar disorder, each in 1 of the 2 independent samples. Analysis of haplotypes of all 22 SNPs in sliding windows of 5 adjacent SNPs revealed an association to the region near intron 7 and 8 in both samples (empirical P values of 0.002 and 0.001, respectively, for the same window).

Association with the HTR4 Gene on Chromosome 5q32

Ohtsuki et al. (2002) performed mutation and association analyses of the HTR4 gene (602164) on 5q32, which encodes the serotonin 5-HT4 receptor, in 96 Japanese patients, 48 with mood disorders and 48 with schizophrenia. Eight polymorphisms and 4 rare variants were identified. Four polymorphisms at or in close proximity to exon d showed significant association with bipolar disorder with odds ratios of 1.5 to 2; these included g.83097C/T (HTR4-SVR (splice variant region) SNP1), g.83159G/A (HTR4-SVRSNP2), g.83164(T)9-10 (HTR4-SVRSNP3), and g.83198A/G (HTR4-SVRSNP4). These polymorphisms were in linkage disequilibrium, and only 3 common haplotypes were observed. One haplotype (SVRSNP1, SVRSNP4 C-A) was significantly associated with bipolar disorder (p = 0.002). The genotypic and haplotypic associations with bipolar disorder were confirmed by the transmission disequilibrium test in the NIMH Genetics Initiative bipolar pedigrees with ratios of transmitted to not transmitted alleles of 1.5 to 2.0 (p = 0.01). The same haplotype that showed association with bipolar disorder was suggested to be associated with schizophrenia in the case-control analysis (p = 0.003) but was not confirmed when Japanese schizophrenia families were tested. The polymorphisms associated with mood disorder were located within the region that encodes the divergent C-terminal tails of the 5-HT4 receptor.

Association with the ABCA13 Gene on Chromosome 7p12.3

Knight et al. (2009) reported evidence that ABCA13 (607807) is a susceptibility factor for both schizophrenia and bipolar disorder. After the initial discovery of its disruption by a chromosome abnormality in a person with schizophrenia, Knight et al. (2009) resequenced ABCA13 exons in 100 cases with schizophrenia and 100 controls. Multiple rare coding variants were identified including 1 nonsense and 9 missense mutations and compound heterozygosity/homozygosity in 6 cases. Variants were genotyped in more than 1,600 additional schizophrenia, bipolar, and depression cases and in more than 950 control cohorts, and the frequency of all rare variants combined was greater than controls in schizophrenia (odds ratio = 1.93, P = 0.0057) and bipolar disorder (odds ratio = 2.71, P = 0.00007). The population-attributable risk of these mutations was 2.2% for schizophrenia and 4.0% for bipolar disorder. In a study of 21 families of mutation carriers, Knight et al. (2009) genotyped affected and unaffected relatives and found significant linkage (lod = 4.3) of rare variants with a phenotype including schizophrenia, bipolar disorder, and major depression. Knight et al. (2009) concluded that their data identified a candidate gene (ABCA13), highlighted the genetic overlap between schizophrenia, bipolar disorder, and depression, and suggested that rare coding variants may contribute significantly to risk of these disorders.

Association with the DRD4 Gene on Chromosome 11p15

Lopez Leon et al. (2005) conducted a metaanalysis to reevaluate the role of the 48-bp repeat polymorphism of the dopamine D4 receptor gene (DRD4; 126452) on chromosome 11p15 in mood disorders by studying 917 patients with unipolar or bipolar affective disorder and 1,164 control subjects from 12 samples using the Cockrane Review Manager. An association was found between all mood disorder groups and the DRD4 2-repeat 48-bp (2R) polymorphism. After correcting for multiple testing, the association between this repeat and bipolar affective disorder dropped to insignificance; however, the evidence for an association between the 2R allele and unipolar depression (p less than 0.001) and the combined group (p less than 0.001) remained.

Association with the BDNF Gene on Chromosome 11p13

Geller et al. (2004) noted that Sklar et al. (2002) and Neves-Pereira et al. (2002), using family-based methods, had found that the BDNF val66 allele (113505.0002) was preferentially transmitted to predominantly Caucasian adult probands with bipolar disorder. Geller et al. (2004) reported that the val66 allele was also preferentially transmitted in children with bipolar disorder. Lohoff et al. (2005) studied the BDNF val66 allele in 621 European patients with bipolar I disorder and positive family histories of affective disorder and 998 European controls. The frequency of the val66 allele was significantly increased in the bipolar I patients when compared to controls (P = 0.028; OR of 1.22).

Rybakowski et al. (2006) studied 111 patients with bipolar disorder, 129 schizophrenia patients, and 92 healthy controls utilizing the Wisconsin Card Sorting Test in the context of the BDNF V66M polymorphism. They found that bipolar patients with the val/val genotype made significantly fewer perseverative errors, had more correctly completed categories and conceptual level responses compared to bipolar patients with the val/met or met/met genotypes. No differences were observed in schizophrenia patients and controls.

Association with the CUX2 Gene on Chromosome 12q

Glaser et al. (2005) performed linkage disequilibrium mapping with 17 microsatellite markers across a 1.6-Mb segment forming the central part of the chromosome 12q23-q24 region implicated in several linkage studies for bipolar affective disorder. In a U.K. Caucasian case-control sample of 347 cases and 374 controls, a significant signal was identified (p = 0.0016) for the microsatellite marker M19 at 12q24. Genes, including regulatory elements, around this marker were screened for mutations and the linkage disequilibrium structure of the region determined by genotyping 22 SNPs and insertion/deletion polymorphisms in 94 individuals. Eleven haplotypes and SNPs were genotyped and 3, an insertion/deletion and a SNP within FLJ32356 (rs3840795 and rs933399) and a SNP within CUX2 (rs3847953), showed significant or nearly significant association with bipolar disorder after Bonferroni-correction (p values from 0.002-0.005).

Association with the SLC6A4 Gene on Chromosome 17q11

Lasky-Su et al. (2005) conducted a metaanalysis on case-control studies of the association between 2 polymorphisms of the SLC6A4 gene (a 17-bp VNTR in intron 2, and a 44-bp insertion/deletion in the promoter region; see 182138.0001) and affective disorders (bipolar disorder and unipolar depression) resulting in 4 metaanalyses. For each polymorphism, the authors assessed the evidence for allelic association, heterogeneity among studies, the influence of individual studies, and the potential for publication bias. The short alleles of the 44-bp insertion/deletion polymorphism showed a significant association with bipolar disorder (OR = 1.13, p = 0.001) but not unipolar disorder. The VNTR had no association with either disorder.

Cho et al. (2005) performed 2 metaanalyses of published studies involving the SLC4A4 gene as a candidate for bipolar disorder. The studies were population-based and family-based studies investigating the association with the promoter polymorphism (5-HTTLPR) and the intron 2 VNTR. Seventeen population-based studies comprising 1,712 cases and 2,583 controls and 6 family-based studies comprising 587 trios were included in the 5-HTTLPR metaanalysis. Sixteen population-based studies comprising 1,764 cases and 2,703 controls as well as for family-based studies comprising 382 trios were included in the intron 2 VNTR metaanalysis. Meta-regression showed that neither study type nor ethnic sample significantly contributed to heterogeneity of the metaanalyses. Overall, odds ratios suggested a very small but detectable effect of the serotonin transporter in susceptibility to bipolar disorder.

Association with the BCR Gene on Chromosome 22q11

Hashimoto et al. (2005) studied 171 patients with bipolar disorder, 329 with major depressive disorder, and 351 controls, all of whom were Japanese, for genetic association using 11 single nucleotide polymorphisms, including a missense polymorphism (N796S; rs140504) in the region of the breakpoint cluster region gene (BCR; 151410) on chromosome 22q11. Significant allelic associations with bipolar disorder were observed for 3 single nucleotide polymorphisms and associations with bipolar II disorder were observed for 10 polymorphisms including N796S (bipolar disorder, p = 0.0054; bipolar II disorder, p = 0.0014). There was a significant association with major depression for 6 polymorphisms. S796 allele carriers were in excess in bipolar II patients (p = 0.0046; OR = 3.1, 95% CI, 1.53-8.76).

Association with the COMT Gene on Chromosome 22q11

Comorbid panic disorder may define a subtype of bipolar disorder and may influence the strength of association between bipolar disorder and candidate genes involved in monoamine neurotransmission. Rotondo et al. (2002) studied the frequency of the V158M polymorphism of catechol-O-methyltransferase (COMT; 116790.0001), the 5-HTTLPR polymorphism of the serotonin transporter SLC6A4 (182138.0001), and a splice site polymorphism (IVS7+218C-A) of tryptophan hydroxylase (TPH; 191060) in a case-control association study of bipolar disorder patients with or without lifetime panic disorder. They compared results from DNA extracted from blood leukocytes of 111 unrelated subjects of Italian descent meeting DSM-III-R criteria for bipolar disorder, including 49 with and 62 without comorbid lifetime panic disorder, with those of 127 healthy subjects. Relative to the comparison subjects, subjects with bipolar disorder without panic disorder, but not those with comorbid bipolar disorder and panic disorder, showed significantly higher frequencies of the COMT met158 and the short 5-HTTLPR alleles. No statistical significance was found between the bipolar disorder groups and the TPH polymorphism. Rotondo et al. (2002) concluded that bipolar disorder without panic disorder may represent a more homogeneous form of illness and that variants of the COMT and SLC6A4 genes may influence clinical features of bipolar disorder.

Association with the XBP1 Gene on Chromosome 22q12

See 194355 for discussion of an association between susceptibility to bipolar disorder and a polymorphism in the XBP1 gene.

Association with the TRPM2 Gene on Chromosome 21q22

McQuillin et al. (2006) fine mapped chromosome 21q22.3 using 30 genetic markers in 600 bipolar subjects and 450 ancestrally matched supernormal controls. Allelic association with D21S171 (p = 0.016), rs1556314 (p = 0.008), and rs1785467 (p = 0.025) was observed. A test of association with a 3-locus haplotype across a susceptibility region was significant with a permutation test (p = 0.011), and a 2-SNP haplotype was also significantly associated with bipolar disorder (p = 0.01). The 2 brain-expressed genes present in the associated region, TRPM2 (603749) and C21ORF29 (612920), were sequenced from subjects who had inherited the associated marker alleles. The rs1556314 polymorphism in exon 11 of TRPM2, which causes an asp543-to-glu (D543E) change, showed the strongest association with bipolar disorder (p = 0.008). McQuillin et al. (2006) noted that deletion of exon 11 is known to cause dysregulation of cellular calcium homeostasis in response to oxidative stress.

Association with Repeat Expansions

Del-Favero et al. (2002) studied the CTG repeat in the third intron of the SEF2-1B gene (602272) located at 18q21.1 and the CAG repeat at the ERDA1 locus (603279) located at 17q21.3 in a large combined European case-control sample of bipolar affective disorder. The sample consisted of 403 patients and 486 controls matched for age, gender, and ethnicity. The patients were consecutively recruited from 5 participating centers in Belgium, Croatia, Denmark, Scotland, and Sweden. Dichotomous analysis of the combined sample did not show a significant difference in expansion frequency between cases and controls at either of the 2 loci. Secondary analysis after stratification for family history of affective disorder in first-degree relatives and disease severity revealed a borderline significant difference (p = 0.03) with a relative risk of 2.43 of developing bipolar disorder in familial cases homozygous for the expanded SEF2-1B allele. This finding rendered further support to the hypothesis that SEF2-1B cannot be excluded as a susceptibility gene for bipolar disorder or that SEF2-1B is in linkage disequilibrium with a causal gene for bipolar disorder.

Tsutsumi et al. (2004) used a repeat expansion detection assay to examine genomic DNA from 100 unrelated probands with schizophrenia and 68 unrelated probands with bipolar affective disorder for the presence of CAG/CTG repeat expansions. They found that 28% of probands with schizophrenia and 38% of probands with bipolar disorder had a CAG/CTG repeat in the expanded range. Each expansion could be explained by 1 of 3 nonpathogenic repeat expansions known to exist in the general population. Thus, a novel CAG/CTG repeat expansion was not a common genetic risk factor for bipolar disorder or schizophrenia in this study.

Association with the Mitochondrial MTND1 Gene

Munakata et al. (2004) reported an association between bipolar disorder and a polymorphism in the mitochondrial MTND1 gene (516000).

Gene Interaction and Locus Heterogeneity

Jamra et al. (2007) presented the first genomewide interaction and locus heterogeneity linkage scan in bipolar affective disorder, using a large linkage dataset (52 families of European descent; 448 participants and 259 affected individuals). The results provided the strongest evidence of interaction between BPAD genes on chromosome 2q22-q24 and 6q23-q24, which was observed symmetrically in both directions; nonparametric lod (NPL) scores of 7.55 on 2q and 7.63 on 6q; P less than 0.0001 and P = 0.0001, respectively, after a genomewide permutation procedure. The second-best BPAD interaction evidence was observed between 2q22-q24 and 15q26. Here, Jamra et al. (2007) also observed a symmetric interaction. Heterogeneity analysis revealed locus heterogeneity at 2q, 6p, 11p, 13q, and 22q, which was supported by adjacent markers within each region and by previously reported BPAD linkage findings.

Epigenetic Theory of Major Psychosis

As a test of the hypothesis that epigenetic misregulation is consistent with various nonmendelian features of schizophrenia (181500) and bipolar disorder, Mill et al. (2008) used CpG island microarrays to identify DNA methylation changes in the frontal cortex and germline associated with schizophrenia and bipolar disorder. In the frontal cortex they found evidence for psychosis-associated DNA methylation differences in numerous loci, including several involved in glutamatergic and GABAergic neurotransmission, brain development, and other processes functionally linked to disease etiology. DNA methylation changes in a significant proportion of these loci corresponded to reported changes of steady-state mRNA levels associated with psychosis. Gene ontology analysis highlighted epigenetic disruption to loci involved in mitochondrial function, brain development, and stress response. Changes in both the brain and the germline of affected individuals suggested that systemic epigenetic dysfunction may be associated with major psychosis. Mill et al. (2008) observed that frontal cortex DNA methylation in the BDNF gene (113505) is correlated with genotype at a nearby nonsynonymous SNP (V66M) that had been associated with major psychosis.

Reviews

See Kato (2007) for a review of molecular genetic findings on bipolar disorder and major depression from 2004 to 2007. Also see the reviews on the genetics of bipolar disorder by Craddock and Sklar (2009, 2013).

Associations Pending Confirmation

For a discussion of a possible association between variation in the KCNH7 gene and susceptibility to bipolar spectrum disorder, see 608169.0001.


Animal Model

Maeng et al. (2008) found that transgenic mice with selective neuron-specific overexpression of Bag1 (601497) in the hippocampus did not have obvious motor, sensory, or learning impairments, but showed less anxious behavior and had higher spontaneous recovery rates from helplessness behavior compared to wildtype mice. These transgenic mice also recovered faster from tests designed to trigger hyperlocomotion or addictive behaviors. In contrast, heterozygous Bag1 +/- mice showed enhanced extreme behavioral responses and less recovery in similar tests. The data suggested that BAG1 may play a role in affective resilience, and perhaps regulates recovery from behavioral impairments observed in patients with bipolar affective disorder. Maeng et al. (2008) postulated that the effects are mediated by BAG1 regulation of glucocorticoid receptor function.


REFERENCES

  1. Aita, V. M., Liu, J., Knowles, J. A., Terwilliger, J. D., Baltazar, R., Grunn, A., Loth, J. E., Kanyas, K., Lerer, B., Endicott, J., Wang, Z., Penchaszadeh, G., Gilliam, T. C., Baron, M. A comprehensive linkage analysis of chromosome 21q22 supports prior evidence for a putative bipolar affective disorder locus. Am. J. Hum. Genet. 64: 210-217, 1999. [PubMed: 9915960] [Full Text: https://doi.org/10.1086/302185]

  2. Anghelescu, I., Dettling, M. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1859 only, 2014. [PubMed: 24806177] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  3. Bellivier, F., Golmard, J.-L., Henry, C., Leboyer, M., Schurhoff, F. Admixture analysis of age at onset in bipolar I affective disorder. Arch. Gen. Psychiat. 58: 510-512, 2001. [PubMed: 11343533] [Full Text: https://doi.org/10.1001/archpsyc.58.5.510]

  4. Berrettini, W. H., Ferraro, T. N., Goldin, L. R., Weeks, D. E., Detera-Wadleigh, S., Nurnberger, J. I., Jr., Gershon, E. S. Chromosome 18 DNA markers and manic-depressive illness: evidence for a susceptibility gene. Proc. Nat. Acad. Sci. 91: 5918-5921, 1994. [PubMed: 8016089] [Full Text: https://doi.org/10.1073/pnas.91.13.5918]

  5. Birnbaum, R., Shin, J. H., Weinberger, D. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1855-1856, 2014. [PubMed: 24806173] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  6. Blackwood, D. H. R., He, L., Morris, S. W., McLean, A., Whitton, C., Thomson, M., Walker, M. T., Woodburn, K., Sharp, C. M., Wright, A. F., Shibasaki, Y., St. Clair, D. M., Porteous, D. J., Muir, W. J. A locus for bipolar affective disorder on chromosome 4p. Nature Genet. 12: 427-430, 1996. [PubMed: 8630499] [Full Text: https://doi.org/10.1038/ng0496-427]

  7. Brent, D. A., Mann, J. J. Family genetic studies, suicide, and suicidal behavior. Am. J. Med. Genet. C Semin. Med. Genet. 133C: 13-24, 2005. [PubMed: 15648081] [Full Text: https://doi.org/10.1002/ajmg.c.30042]

  8. Brent, D. A., Mann, J. J. Familial pathways to suicidal behavior--understanding and preventing suicide among adolescents. New Eng. J. Med. 355: 2719-2721, 2006. [PubMed: 17192535] [Full Text: https://doi.org/10.1056/NEJMp068195]

  9. Cadoret, R. J. Evidence for genetic inheritance of primary affective disorder in adoptees. Am. J. Psychiat. 135: 463-466, 1978. [PubMed: 637144] [Full Text: https://doi.org/10.1176/ajp.135.4.463]

  10. Chen, C.-H., Lee, C.-S., Lee, M.-T. M., Ouyang, W.-C., Chen, C.-C., Chong, M.-Y., Wu, J.-Y., Tan, H. K.-L., Lee, Y.-C., Chuo, L.-J., Chiu, N.-Y., Tsang, H.-Y., and 31 others. Variant GADL1 and response to lithium therapy in bipolar I disorder. New Eng. J. Med. 370: 119-128, 2014. Note: Erratum: N. Eng. J. Med. 370: 1866 only, 2014. [PubMed: 24369049] [Full Text: https://doi.org/10.1056/NEJMoa1212444]

  11. Cheng, R., Juo, S. H., Loth, J. E., Nee, J., Iossifov, I., Blumenthal, R., Sharpe, L., Kanyas, K., Lerer, B., Lilliston, B., Smith, M., Trautman, K., Gilliam, T. C., Endicott, J., Baron, M. Genome-wide linkage scan in a large bipolar disorder sample from the National Institute of Mental Health genetics initiative suggests putative loci for bipolar disorder, psychosis, suicide, and panic disorder. Molec. Psychiat. 11: 252-260, 2006. [PubMed: 16402137] [Full Text: https://doi.org/10.1038/sj.mp.4001778]

  12. Cho, H. J., Meira-Lima, I., Cordeiro, Q., Michelon, L., Sham, P., Vallada, H., Collier, D. A. Population-based and family-based studies on the serotonin transporter gene polymorphisms and bipolar disorder: a systematic review and meta-analysis. Molec. Psychiat. 10: 771-781, 2005. [PubMed: 15824745] [Full Text: https://doi.org/10.1038/sj.mp.4001663]

  13. Cichon, S., Schumacher, J., Muller, D. J., Hurter, M., Windemuth, C., Strauch, K., Hemmer, S., Schulze, T. G., Schmidt-Wolf, G., Albus, M., Borrmann-Hassenbach, M., Franzek, E., and 15 others. A genome screen for genes predisposing to bipolar affective disorder detects a new susceptibility locus on 8q. Hum. Molec. Genet. 10: 2933-2944, 2001. Note: Erratum: Hum. Molec. Genet. 11: 1685 only, 2002. [PubMed: 11741836] [Full Text: https://doi.org/10.1093/hmg/10.25.2933]

  14. Cichon, S., Winge, I., Mattheisen, M., Georgi, A., Karpushova, A., Freudenberg, J., Freudenberg-Hua, Y., Babadjanova, G., Van Den Bogaert, A., Abramova, L. I., Kapiletti, S., Knappskog, P. M., and 9 others. Brain-specific tryptophan hydroxylase 2 (TPH2): a functional Pro206Ser substitution and variation in the 5-prime region are associated with bipolar affective disorder. Hum. Molec. Genet. 17: 87-97, 2008. [PubMed: 17905754] [Full Text: https://doi.org/10.1093/hmg/ddm286]

  15. Consortium on Lithium Genetics. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1857-1859, 2014. [PubMed: 24806176] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  16. Coon, H., Jensen, S., Hoff, M., Holik, J., Plaetke, R., Reimherr, F., Wender, P., Leppert, M., Byerley, W. A genome-wide search for genes predisposing to manic-depression, assuming autosomal dominant inheritance. Am. J. Hum. Genet. 52: 1234-1249, 1993. [PubMed: 8503452]

  17. Craddock, N., Owen, M., Burge, S., Kurian, B., Thomas, P., McGuffin, P. Familial cosegregation of major affective disorder and Darier's disease (keratosis follicularis). Brit. J. Psychiat. 164: 355-358, 1994. [PubMed: 8199789] [Full Text: https://doi.org/10.1192/bjp.164.3.355]

  18. Craddock, N., Sklar, P. Genetics of bipolar disorder: successful start to a long journey. Trends Genet. 25: 99-105, 2009. [PubMed: 19144440] [Full Text: https://doi.org/10.1016/j.tig.2008.12.002]

  19. Craddock, N., Sklar, P. Genetics of bipolar disorder. Lancet 381: 1654-1662, 2013. [PubMed: 23663951] [Full Text: https://doi.org/10.1016/S0140-6736(13)60855-7]

  20. Degn, B., Lundorf, M. D., Wang, A., Vang, M., Mors, O., Kruse, T. A., Ewald, H. Further evidence for a bipolar risk gene on chromosome 12q24 suggested by investigation of haplotype sharing and allelic association in patients from the Faroe Islands. Molec. Psychiat. 6: 450-455, 2001. [PubMed: 11443532] [Full Text: https://doi.org/10.1038/sj.mp.4000882]

  21. Del-Favero, J., Van Gestel, S., Borglum, A. D., Muir, W., Ewald, H., Mors, O., Ivezic, S., Oruc, L., Adolfsson, R., Blackwood, D., Kruse, T., Mendlewicz, J., Schalling, M., Van Broeckhoven, C. European combined analysis of the CTG18.1 and the ERDA1 CAG/CTG repeats in bipolar disorder. Europ. J. Hum. Genet. 10: 276-280, 2002. [PubMed: 12032737] [Full Text: https://doi.org/10.1038/sj.ejhg.5200803]

  22. Dick, D. M., Foroud, T., Flury, L., Bowman, E. S., Miller, M. J., Rau, N. L., Moe, P. R., Samavedy, N., El-Mallakh, R., Manji, H., Glitz, D. A., Meyer, E. T., and 20 others. Genomewide linkage analyses of bipolar disorder: a new sample of 250 pedigrees from the National Institute of Mental Health Genetics Initiative. Am. J. Hum. Genet. 73: 107-114, 2003. Note: Erratum: Am. J. Hum. Genet. 73: 979 only, 2003. [PubMed: 12772088] [Full Text: https://doi.org/10.1086/376562]

  23. Drevets, W. C., Price, J. L., Simpson, J. R., Jr., Todd, R. D., Reich, T., Vannier, M., Ralchle, M. E. Subgenual prefrontal cortex abnormalities in mood disorders. Nature 386: 824-827, 1997. [PubMed: 9126739] [Full Text: https://doi.org/10.1038/386824a0]

  24. Egeland, J. A., Gerhard, D. S., Pauls, D. L., Sussex, J. N., Kidd, K. K., Allen, C. R., Hostetter, A. M., Housman, D. E. Bipolar affective disorders linked to DNA markers on chromosome 11. Nature 325: 783-787, 1987. [PubMed: 2881209] [Full Text: https://doi.org/10.1038/325783a0]

  25. Ekholm, J. M., Kieseppa, T., Hiekkalinna, T., Partonen, T., Paunio, T., Perola, M., Ekelund, J., Lonnqvist, J., Pekkarinen-Ijas, P., Peltonen, L. Evidence of susceptibility loci on 4q32 and 16p12 for bipolar disorder. Hum. Molec. Genet. 12: 1907-1915, 2003. [PubMed: 12874110] [Full Text: https://doi.org/10.1093/hmg/ddg199]

  26. Escamilla, M. A., McInnes, L. A., Spesny, M., Reus, V. I., Service, S. K., Shimayoshi, N., Tyler, D. J., Silva, S., Molina, J., Gallegos, A., Meza, L., Cruz, M. L., and 10 others. Assessing the feasibility of linkage disequilibrium methods for mapping complex traits: an initial screen for bipolar disorder loci on chromosome 18. Am. J. Hum. Genet. 64: 1670-1678, 1999. [PubMed: 10330354] [Full Text: https://doi.org/10.1086/302400]

  27. Escamilla, M. A., Spesny, M., Reus, V. I., Gallegos, A., Meza, L., Molina, J., Sandkuijl, L. A., Fournier, E., Leon, P. E., Smith, L. B., Freimer, N. B. Use of linkage disequilibrium approaches to map genes for bipolar disorder in the Costa Rican population. Am. J. Med. Genet. 67: 244-253, 1996. [PubMed: 8725743] [Full Text: https://doi.org/10.1002/(SICI)1096-8628(19960531)67:3<244::AID-AJMG2>3.0.CO;2-N]

  28. Etain, B., Mathieu, F., Rietschel, M., Maier, W., Albus, M., McKeon, P., Roche, S., Kealey, C., Blackwood, D., Muir, W., Bellivier, F., Henry, C. {and 24 others}: Genome-wide scan for genes involved in bipolar affective disorder in 70 European families ascertained through a bipolar type I early-onset proband: supportive evidence for linkage at 3p14. Molec. Psychiat. 11: 685-694, 2006. [PubMed: 16534504] [Full Text: https://doi.org/10.1038/sj.mp.4001815]

  29. Ewald, H., Degn, B., Mors, O., Kruse, T. A. Significant linkage between bipolar affective disorder and chromosome 12q24. Psychiat. Genet. 8: 131-140, 1998. [PubMed: 9800214] [Full Text: https://doi.org/10.1097/00041444-199800830-00002]

  30. Ewald, H., Flint, T., Kruse, T. A., Mors, O. A genome-wide scan shows significant linkage between bipolar disorder and chromosome 12q24.3 and suggestive linkage to chromosomes 1p22-21, 4p16, 6q14-22, 10q26 and 16p13.3. Molec. Psychiat. 7: 734-744, 2002. [PubMed: 12192618] [Full Text: https://doi.org/10.1038/sj.mp.4001074]

  31. Fallin, M. D., Lasseter, V. K., Wolyniec, P. S., McGrath, J. A., Nestadt, G., Valle, D., Liang, K.-Y., Pulver, A. E. Genomewide linkage scan for bipolar-disorder susceptibility loci among Ashkenazi Jewish families. Am. J. Hum. Genet. 75: 204-219, 2004. [PubMed: 15208783] [Full Text: https://doi.org/10.1086/422474]

  32. Ferreira, M. A. R., O'Donovan, M. C., Meng, Y. A., Jones, I. R., Ruderfer, D. M., Jones, L., Fan, J., Kirov, G., Perlis, R. H., Green, E. K., Smoller, J. W., Grozeva, D., and 51 others. Collaborative genome-wide association analysis supports a role for ANK3 and CACNA1C in bipolar disorder. Nature Genet. 40: 1056-1058, 2008. [PubMed: 18711365] [Full Text: https://doi.org/10.1038/ng.209]

  33. Freimer, N. B., Reus, V. I., Escamilla, M. A., McInnes, L. A., Spesny, M., Leon, P., Service, S. K., Smith, L. B., Silva, S., Rojas, E., Gallegos, A., Meza, L., Fournier, E., Baharloo, S., Blankenship, K., Tyler, D. J., Batki, S., Vinogradov, S., Weissenbach, J., Barondes, S. H., Sandkuijl, L. A. Genetic mapping using haplotype, association and linkage methods suggests a locus for severe bipolar disorder (BPI) at 18q22-q23. Nature Genet. 12: 436-441, 1996. [PubMed: 8630501] [Full Text: https://doi.org/10.1038/ng0496-436]

  34. Garner, C., McInnes, L. A., Service, S. K., Spesny, M., Fournier, E., Leon, P., Freimer, N. B. Linkage analysis of a complex pedigree with severe bipolar disorder, using a Markov chain Monte Carlo method. Am. J. Hum. Genet. 68: 1061-1064, 2001. [PubMed: 11222106] [Full Text: https://doi.org/10.1086/319517]

  35. Gelernter, J. Genetics of bipolar affective disorder: time for another reinvention? (Editorial) Am. J. Hum. Genet. 56: 1262-1266, 1995. [PubMed: 7762549]

  36. Geller, B., Badner, J. A., Tillman, R., Christian, S. L., Bolhofner, K., Cook, E. H., Jr. Linkage disequilibrium of the brain-derived neurotrophic factor val66met polymorphism in children with prepubertal and early adolescent bipolar disorder phenotype. Am. J. Psychiat. 161: 1698-1700, 2004. [PubMed: 15337662] [Full Text: https://doi.org/10.1176/appi.ajp.161.9.1698]

  37. Gerhard, D. S., LaBuda, M. C., Bland, S. D., Allen, C., Egeland, J. A., Pauls, D. L. Initial report of a genome search for the affective disorder predisposition gene in the old order Amish pedigrees: chromosomes 1 and 11. Am. J. Med. Genet. 54: 398-404, 1994. [PubMed: 7726215] [Full Text: https://doi.org/10.1002/ajmg.1320540421]

  38. Gill, M., McKeon, P., Humphries, P. Linkage analysis of manic depression in an Irish family using H-RAS-1 and INS DNA markers. J. Med. Genet. 25: 634-637, 1988. [PubMed: 3184144] [Full Text: https://doi.org/10.1136/jmg.25.9.634]

  39. Ginns, E. I., Ott, J., Egeland, J. A., Allen, C. R., Fann, C. S. J., Pauls, D. L., Weissenbach, J., Carulli, J. P., Falls, K. M., Keith, T. P., Paul, S. M. A genome-wide search for chromosomal loci linked to bipolar affective disorder in the Old Order Amish. Nature Genet. 12: 431-435, 1996. [PubMed: 8630500] [Full Text: https://doi.org/10.1038/ng0496-431]

  40. Ginns, E. I., St. Jean, P., Philibert, R. A., Galdzicka, M., Damschroder-Williams, P., Thiel, B., Long, R. T., Ingraham, L. J., Dalwaldi, H., Murray, M. A., Ehlert, M., Paul, S., and 19 others. A genome-wide search for chromosomal loci linked to mental health wellness in relatives at high risk for bipolar affective disorder among the Old Order Amish. Proc. Nat. Acad. Sci. 95: 15531-15536, 1998. [PubMed: 9861003] [Full Text: https://doi.org/10.1073/pnas.95.26.15531]

  41. Glaser, B., Kirov, G., Green, E., Craddock, N., Owen, M. J. Linkage disequilibrium mapping of bipolar affective disorder at 12q23-q24 provides evidence for association at CUX2 and FLJ32356. Am. J. Med. Genet. 132B: 38-45, 2005. [PubMed: 15389760] [Full Text: https://doi.org/10.1002/ajmg.b.30081]

  42. Goodwin, F. K., Jamison, K. R. Manic-Depressive Illness. New York: Oxford University Press 1990.

  43. Greenwood, T. A., Alexander, M., Keck, P. E., McElroy, S., Sadovnick, A. D., Remick, R. A., Kelsoe, J. R. Evidence for linkage disequilibrium between the dopamine transporter and bipolar disorder. Am. J. Med. Genet. 105: 145-151, 2001. [PubMed: 11304827] [Full Text: https://doi.org/10.1002/1096-8628(2001)9999:9999<::aid-ajmg1161>3.0.co;2-8]

  44. Greenwood, T. A., Schork, N. J., Eskin, E., Kelsoe, J. R. Identification of additional variants within the human dopamine transporter gene provides further evidence for an association with bipolar disorder in two independent samples. Molec. Psychiat. 11: 125-133, 2006. [PubMed: 16261167] [Full Text: https://doi.org/10.1038/sj.mp.4001764]

  45. Grof, P., Alda, M., Grof, E., Zvolsky, P., Walsh, M. Lithium response and genetics of affective disorders. J. Affect. Dis. 32: 85-95, 1994. [PubMed: 7829768] [Full Text: https://doi.org/10.1016/0165-0327(94)90066-3]

  46. Hashimoto, R., Okada, T., Kato, T., Kosuga, A., Tatsumi, M., Kamijima, K., Kunugi, H. The breakpoint cluster region gene on chromosome 22q11 is associated with bipolar disorder. Biol. Psychiat. 57: 1097-1102, 2005. [PubMed: 15866548] [Full Text: https://doi.org/10.1016/j.biopsych.2005.02.019]

  47. Hill, E. M., Wilson, A. F., Elston, R. C., Winokur, G. Evidence for possible linkage between genetic markers and affective disorders. Biol. Psychiat. 24: 903-917, 1988. [PubMed: 3233232] [Full Text: https://doi.org/10.1016/0006-3223(88)90225-9]

  48. Hodge, S. E., Wickramaratne, P. Statistical pitfalls in detecting age-of-onset anticipation: the role of correlation in studying anticipation and detecting ascertainment bias. Psychiat. Genet. 5: 43-47, 1995. [PubMed: 7582880] [Full Text: https://doi.org/10.1097/00041444-199521000-00007]

  49. Holmes, D., Brynjolfsson, J., Brett, P., Curtis, D., Petursson, H., Sherrington, R., Gurling, H. No evidence for a susceptibility locus predisposing to manic depression in the region of the dopamine (D2) receptor gene. Brit. J. Psychiat. 158: 635-641, 1991. [PubMed: 1677600] [Full Text: https://doi.org/10.1192/bjp.158.5.635]

  50. Hong, K. S., McInnes, L. A., Service, S. K., Song, T., Lucas, J., Silva, S., Fournier, E., Leon, P., Molina, J., Reus, V. I., Sandkuijl, L. A., Freimer, N. B. Genetic mapping using haplotype and model-free linkage analysis supports previous evidence for a locus predisposing to severe bipolar disorder at 5q31-33. Am. J. Med. Genet. 125B: 83-86, 2004. [PubMed: 14755450] [Full Text: https://doi.org/10.1002/ajmg.b.20091]

  51. Ikeda, M., Kondo, K., Iwata, N. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1856-1857, 2014. [PubMed: 24806175] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  52. Jamra, R. A., Fuerst, R., Kaneva, R., Diaz, G. O., Rivas, F., Mayoral, F., Gay, E., Sans, S., Gonzalez, M. J., Gil, S., Cabaleiro, F., del Rio, F. {and 18 others}: The first genomewide interaction and locus-heterogeneity linkage scan in bipolar affective disorder: strong evidence of epistatic effects between loci on chromosome 2q and 6q. Am. J. Hum. Genet. 81: 974-986, 2007. [PubMed: 17924339] [Full Text: https://doi.org/10.1086/521690]

  53. Joffe, R. T., Horvath, Z., Tarvydas, I. Bipolar affective disorder and thalassemia minor. (Letter) Am. J. Psychiat. 143: 933, 1986. [PubMed: 3717440] [Full Text: https://doi.org/10.1176/ajp.143.7.933a]

  54. Kato, T. Molecular genetics of bipolar disorder and depression. Psychiat. Clin. Neurosci. 61: 3-19, 2007. [PubMed: 17239033] [Full Text: https://doi.org/10.1111/j.1440-1819.2007.01604.x]

  55. Kelsoe, J. R., Spence, M. A., Loetscher, E., Foguet, M., Sadovnick, A. D., Remick, R. A., Flodman, P., Khristich, J., Mroczkowski-Parker, Z., Brown, J. L., Masser, D., Ungerleider, S., Rapaport, M. H., Wishart, W. L., Luebbert, H. A genome survey indicates a possible susceptibility locus for bipolar disorder on chromosome 22. Proc. Nat. Acad. Sci. 98: 585-590, 2001. [PubMed: 11149935] [Full Text: https://doi.org/10.1073/pnas.98.2.585]

  56. Kimmel, R. J., Kovacs, I., Vrabel, C., Wood, B., Schalling, M., Kelsoe, J. R. Cosegregation of bipolar disorder and autosomal-dominant medullary cystic kidney disease in a large family. Am. J. Psychiat. 162: 1972-1974, 2005. [PubMed: 16199849] [Full Text: https://doi.org/10.1176/appi.ajp.162.10.1972]

  57. Kirk, R., Furlong, R. A., Amos, W., Cooper, G., Rubinsztein, J. S., Walsh, C., Paykel, E. S., Rubinsztein, D. C. Mitochondrial genetic analyses suggest selection against maternal lineages in bipolar affective disorder. Am. J. Hum. Genet. 65: 508-518, 1999. [PubMed: 10417293] [Full Text: https://doi.org/10.1086/302507]

  58. Klein, P. S., Melton, D. A. A molecular mechanism for the effect of lithium on development. Proc. Nat. Acad. Sci. 93: 8455-8459, 1996. [PubMed: 8710892] [Full Text: https://doi.org/10.1073/pnas.93.16.8455]

  59. Knight, H. M., Pickard, B. S., Maclean, A., Malloy, M. P., Soares, D. C., McRae, A. F., Condie, A., White, A., Hawkins, W., McGhee, K., van Beck, M., MacIntyre, D. J., Starr, J. M., Deary, I. J., Visscher, P. M., Porteous, D. J., Cannon, R. E., St Clair, D., Muir, W. J., Blackwood, D. H. R. A cytogenetic abnormality and rare coding variants identify ABCA13 as a candidate gene in schizophrenia, bipolar disorder, and depression. Am. J. Hum. Genet. 85: 833-846, 2009. [PubMed: 19944402] [Full Text: https://doi.org/10.1016/j.ajhg.2009.11.003]

  60. Knowles, J. A., Rao, P. A., Cox-Matise, T., Loth, J. E., de Jesus, G. M., Levine, L., Das, K., Penchaszadeh, G. K., Alexander, J. R., Lerer, B., Endicott, J., Ott, J., Gilliam, T. C., Baron, M. No evidence for significant linkage between bipolar affective disorder and chromosome 18 pericentromeric markers in a large series of multiplex extended pedigrees. Am. J. Hum. Genet. 62: 916-924, 1998. [PubMed: 9529343] [Full Text: https://doi.org/10.1086/301785]

  61. LaBuda, M. C., Maldonado, M., Marshall, D., Otten, K., Gerhard, D. S. A follow-up report of a genome search for affective disorder predisposition loci in the Old Order Amish. Am. J. Hum. Genet. 59: 1343-1362, 1996. [PubMed: 8940281]

  62. Lasky-Su, J. A., Faraone, S. V., Glatt, S. J., Tsuang, M. T. Meta-analysis of the association between two polymorphisms in the serotonin transporter gene and affective disorders. Am. J. Med. Genet. 133B: 110-115, 2005. [PubMed: 15578606] [Full Text: https://doi.org/10.1002/ajmg.b.30104]

  63. Law, A., Richard, C. W., III, Cottingham, R. W., Jr., Lathrop, G. M., Cox, D. R., Myers, R. M. Genetic linkage analysis of bipolar affective disorder in an Old Order Amish pedigree. Hum. Genet. 88: 562-568, 1992. [PubMed: 1551659] [Full Text: https://doi.org/10.1007/BF00219345]

  64. Le, F., Mitchell, P., Vivero, C., Waters, B., Donald, J., Selbie, L. A., Shine, J., Schofield, P. Exclusion of close linkage of bipolar disorder to the GS-alpha subunit gene in nine Australian pedigrees. J. Affect. Disord. 32: 187-195, 1994. [PubMed: 7531727] [Full Text: https://doi.org/10.1016/0165-0327(94)90017-5]

  65. Lee, C.-S., Cheng, A. T.-A. Reply to variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1859-1860, 2014. [PubMed: 24806172] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  66. Levinson, D. F., Levinson, M. D., Segurado, R., Lewis, C. M. Genome scan meta-analysis of schizophrenia and bipolar disorder. Part I: methods and power analysis. Am. J. Hum. Genet. 73: 17-33, 2003. [PubMed: 12802787] [Full Text: https://doi.org/10.1086/376548]

  67. Lin, P.-I., McInnis, M. G., Potash, J. B., Willour, V. L., MacKinnon, D. F., Miao, K., DePaulo, J. R., Zandi, P. P. Assessment of the effect of age at onset on linkage to bipolar disorder: evidence on chromosomes 18p and 21q. Am. J. Hum. Genet. 77: 545-555, 2005. [PubMed: 16175501] [Full Text: https://doi.org/10.1086/491602]

  68. Lohoff, F. W., Sander, T., Ferraro, T. N., Dahl, J. P., Gallinat, J., Berrettini, W. H. Confirmation of association between the val66-to-met polymorphism in the brain-derived neurotrophic factor (BDNF) gene and bipolar I disorder. Am. J. Med. Genet. 139B: 51-53, 2005. [PubMed: 16152572] [Full Text: https://doi.org/10.1002/ajmg.b.30215]

  69. Lombroso, C. Crime: Its Causes and Remedies. Boston: Little, Brown 1911.

  70. Lopez Leon, S., Croes, E. A., Sayed-Tabatabaei, F. A., Claes, S., Van Broeckhoven, C., van Duijn, C. M. The dopamine D4 receptor gene 48-base-pair-repeat polymorphism and mood disorders: a meta-analysis. Biol. Psychiat. 57: 999-1003, 2005. [PubMed: 15860340] [Full Text: https://doi.org/10.1016/j.biopsych.2005.01.030]

  71. Maeng, S., Hunsberger, J. G., Pearson, B., Yuan, P., Wang, Y., Wei, Y., McCammon, J., Schloesser, R. J., Zhou, R., Du, J., Chen, G., McEwen, B., Reed, J. C., Manji, H. K. BAG1 plays a critical role in regulating recovery from both manic-like and depression-like behavioral impairments. Proc. Nat. Acad. Sci. 105: 8766-8771, 2008. [PubMed: 18562287] [Full Text: https://doi.org/10.1073/pnas.0803736105]

  72. Maziade, M., Roy, M.-A., Chagnon, Y. C., Cliche, D., Fournier, J.-P., Montgrain, N., Dion, C., Lavallee, J.-C., Garneau, Y., Gingras, N., Nicole, L., Pires, A., Ponton, A.-M., Potvin, A., Wallot, H., Merette, C. Shared and specific susceptibility loci for schizophrenia and bipolar disorder: a dense genome scan in Eastern Quebec families. Molec. Psychiat. 10: 486-499, 2005. [PubMed: 15534619] [Full Text: https://doi.org/10.1038/sj.mp.4001594]

  73. McInnes, L. A., Escamilla, M. A., Service, S. K., Reus, V. I., Leon, P., Silva, S., Rojas, E., Spesny, M., Baharloo, S., Blankenship, K., Peterson, A., Tyler, D., and 11 others. A complete genome screen for genes predisposing to severe bipolar disorder in two Costa Rican pedigrees. Proc. Nat. Acad. Sci. 93: 13060-13065, 1996. [PubMed: 8917544] [Full Text: https://doi.org/10.1073/pnas.93.23.13060]

  74. McInnes, L. A., Service, S. K., Reus, V. I., Barnes, G., Charlat, O., Jawahar, S., Lewitzky, S., Yang, Q., Duong, Q., Spesny, M., Araya, C., Araya, X., and 23 others. Fine-scale mapping of a locus for severe bipolar mood disorder on chromosome 18p11.3 in the Costa Rican population. Proc. Nat. Acad. Sci. 98: 11485-11490, 2001. [PubMed: 11572994] [Full Text: https://doi.org/10.1073/pnas.191519098]

  75. McInnis, M. G., McMahon, F. J., Chase, G. A., Simpson, S. G., Ross, C. A., DePaulo, J. R., Jr. Anticipation in bipolar affective disorder. Am. J. Hum. Genet. 53: 385-390, 1993. [PubMed: 8328456]

  76. McInnis, M. G. Anticipation: an old idea in new genes. (Editorial) Am. J. Hum. Genet. 59: 973-979, 1996. [PubMed: 8900222]

  77. McMahon, F. J., Simpson, S. G., McInnis, M. G., Badner, J. A., MacKinnon, D. F., DePaulo, J. R. Linkage of bipolar disorder to chromosome 18q and the validity of bipolar II disorder. Arch. Gen. Psychiat. 58: 1025-1031, 2001. [PubMed: 11695948] [Full Text: https://doi.org/10.1001/archpsyc.58.11.1025]

  78. McMahon, F. J., Stine, O. C., Meyers, D. A., Simpson, S. G., DePaulo, J. R. Patterns of maternal transmission in bipolar affective disorder. Am. J. Hum. Genet. 56: 1277-1286, 1995. [PubMed: 7762550]

  79. McQueen, M. B., Devlin, B., Faraone, S. V., Nimgaonkar, V. L., Sklar, P., Smoller, J. W., Jamra, R. A., Albus, M., Bacanu, S.-A., Baron, M., Barrett, T. B., Berrettini, W., and 41 others. Combined analysis from eleven linkage studies of bipolar disorder provides strong evidence of susceptibility loci on chromosomes 6q and 8q. Am. J. Hum. Genet. 77: 582-595, 2005. [PubMed: 16175504] [Full Text: https://doi.org/10.1086/491603]

  80. McQuillin, A., Bass, N. J., Kalsi, G., Lawrence, J., Puri, V., Choudhury, K., Detera-Wadleigh, S. D., Curtis, D., Gurling, H. M. D. Fine mapping of a susceptibility locus for bipolar and genetically related unipolar affective disorders, to a region containing the C21ORF29 and TRPM2 genes on chromosome 21q22.3. Molec. Psychiat. 11: 134-142, 2006. [PubMed: 16205735] [Full Text: https://doi.org/10.1038/sj.mp.4001759]

  81. Middleton, F. A., Pato, M. T., Gentile, K. L., Morley, C. P., Zhao, X., Eisener, A. F., Brown, A., Petryshen, T. L., Kirby, A. N., Medeiros, H., Carvalho, C., Macedo, A., and 11 others. Genomewide linkage analysis of bipolar disorder by use of a high-density single-nucleotide-polymorphism (SNP) genotype assay: a comparison with microsatellite marker assays and finding of significant linkage to chromosome 6q22. Am. J. Hum. Genet. 74: 886-897, 2004. [PubMed: 15060841] [Full Text: https://doi.org/10.1086/420775]

  82. Mill, J., Tang, T., Kaminsky, Z., Khare, T., Yazdanpanah, S., Bouchard, L., Jia, P., Assadzadeh, A., Flanagan, J., Schumacher, A., Wang, S.-C., Petronis, A. Epigenetic profiling reveals DNA-methylation changes associated with major psychosis. Am. J. Hum. Genet. 82: 696-711, 2008. [PubMed: 18319075] [Full Text: https://doi.org/10.1016/j.ajhg.2008.01.008]

  83. Morel, B. A. Traite des Degenerescences. Paris: JB Bailliere 1857.

  84. Morissette, J., Villeneuve, A., Bordeleau, L., Rochette, D., Laberge, C., Gagne, B. et al. Genome-wide search for linkage of bipolar affective disorders in a very large pedigree derived from a homogeneous population in Quebec points to a locus of major effect on chromosome 12q23-q24. Am. J. Med. Genet. 88: 567-587, 1999. [PubMed: 10490718] [Full Text: https://doi.org/10.1002/(sici)1096-8628(19991015)88:5<567::aid-ajmg24>3.0.co;2-8]

  85. Munakata, K., Tanaka, M., Mori, K., Washizuka, S., Yoneda, M., Tajima, O., Akiyama, T., Nanko, S., Kunugi, H., Tadokoro, K., Ozaki, N., Inada, T., Sakamoto, K., Fukunaga, T., Iijima, Y., Iwata, N., Tatsumi, M., Yamada, K., Yoshikawa, T., Kato, T. Mitochondrial DNA 3644T-C mutation associated with bipolar disorder. Genomics 84: 1041-1050, 2004. [PubMed: 15533721] [Full Text: https://doi.org/10.1016/j.ygeno.2004.08.015]

  86. Neiswanger, K., Slaugenhaupt, S. A., Hughes, H. B., Frank, E., Frankel, D. R., McCarty, M. J., Chakravarti, A., Zubenko, G. S., Kupfer, D. J., Kaplan, B. B. Evidence against close linkage of unipolar affective illness to human chromosome 11p markers HRAS1 and INS and chromosome Xq marker DXS52. Biol. Psychiat. 28: 63-72, 1990. [PubMed: 1973904] [Full Text: https://doi.org/10.1016/0006-3223(90)90433-3]

  87. Neves-Pereira, M., Mundo, E., Muglia, P., King, N., Macciardi, F., Kennedy, J. L. The brain-derived neurotrophic factor gene confers susceptibility to bipolar disorder: evidence from a family-based association study. Am. J. Hum. Genet. 71: 651-655, 2002. [PubMed: 12161822] [Full Text: https://doi.org/10.1086/342288]

  88. O'Reilly, R. L., Bogue, L., Singh, S. M. Pharmacogenetic response to antidepressants in a multicase family with affective disorder. Biol. Psychiat. 36: 467-471, 1994. [PubMed: 7811843] [Full Text: https://doi.org/10.1016/0006-3223(94)90642-4]

  89. Ohtsuki, T., Ishiguro, H., Detera-Wadleigh, S. D., Toyota, T., Shimizu, H., Yamada, K., Yoshitsugu, K., Hattori, E., Yoshikawa, T., Arinami, T. Association between serotonin 4 receptor gene polymorphisms and bipolar disorder in Japanese case-control samples and the NIMH Genetics Initiative bipolar pedigrees. Molec. Psychiat. 7: 954-961, 2002. [PubMed: 12399948] [Full Text: https://doi.org/10.1038/sj.mp.4001133]

  90. Ophoff, R. A., Escamilla, M. A., Service, S. K., Spesny, M., Meshi, D. B., Poon, W., Molina, J., Fournier, E., Gallegos, A., Mathews, C., Neylan, T., Batki, S. L., Roche, E., Ramirez, M., Silva, S., De Mille, M. C., Dong, P., Leon, P. E., Reus, V. I., Sandkuijl, L. A., Freimer, N. B. Genomewide linkage disequilibrium mapping of severe bipolar disorder in a population isolate. Am. J. Hum. Genet. 71: 565-574, 2002. [PubMed: 12119601] [Full Text: https://doi.org/10.1086/342291]

  91. Pakstis, A. J., Kidd, J. R., Castiglione, C. M., Kidd, K. K. Status of the search for a major genetic locus for affective disorder in the Old Order Amish. Hum. Genet. 87: 475-483, 1991. [PubMed: 1879834] [Full Text: https://doi.org/10.1007/BF00197172]

  92. Pardue, L. H. Familial unipolar depressive illness: a pedigree study. Am. J. Psychiat. 132: 970-972, 1975. [PubMed: 1155636] [Full Text: https://doi.org/10.1176/ajp.132.9.970]

  93. Park, N., Juo, S. H., Cheng, R., Liu, J., Loth, J. E., Lilliston, B., Nee, J., Grunn, A., Kanyas, K., Lerer, B., Endicott, J., Gilliam, T. C., Baron, M. Linkage analysis of psychosis in bipolar pedigrees suggests novel putative loci for bipolar disorder and shared susceptibility with schizophrenia. Molec. Psychiat. 9: 1091-1099, 2004. [PubMed: 15241432] [Full Text: https://doi.org/10.1038/sj.mp.4001541]

  94. Pato, C. N., Pato, M. T., Kirby, A., Petryshen, T. L., Medeiros, H., Carvalho, C., Macedo, A., Dourado, A., Coelho, I., Valente, J., Soares, M. J., Ferreira, C. P., Lei, M., Verner, A., Hudson, T. J., Morley, C. P., Kennedy, J. L., Azevedo, M. H., Daly, M. J., Sklar, P. Genome-wide scan in Portuguese Island families implicates multiple loci in bipolar disorder: fine mapping adds support on chromosomes 6 and 11. Am. J. Med. Genet. 127B: 30-34, 2004. [PubMed: 15108176] [Full Text: https://doi.org/10.1002/ajmg.b.30001]

  95. Pauls, D. L., Gerhard, D. S., Lacy, L. G., Hostetter, A. M., Allen, C. R., Bland, S. D., LaBuda, M. C., Egeland, J. A. Linkage of bipolar affective disorders to markers on chromosome 11p is excluded in a second lateral extension of Amish pedigree 110. Genomics 11: 730-736, 1991. [PubMed: 1774072] [Full Text: https://doi.org/10.1016/0888-7543(91)90081-o]

  96. Pauls, D. L., Ott, J., Paul, S. M., Allen, C. R., Fann, C. S. J., Carulli, J. P., Falls, K. M., Bouthillier, C. A., Gravius, T. C., Keith, T. P., Egeland, J. A., Ginns, E. I. Linkage analysis of chromosome 18 markers do not identify a major susceptibility locus for bipolar affective disorder in the Old Order Amish. Am. J. Hum. Genet. 57: 636-643, 1995. [PubMed: 7668292]

  97. Radhakrishna, U., Senol, S., Herken, H., Gucuyener, K., Gehrig, C., Blouin, J.-L., Akarsu, N. A., Antonarakis, S. E. An apparently dominant bipolar affective disorder (BPAD) locus on chromosome 20p11.2-q11.2 in a large Turkish pedigree. Europ. J. Hum. Genet. 9: 39-44, 2001. [PubMed: 11175298] [Full Text: https://doi.org/10.1038/sj.ejhg.5200584]

  98. Risch, N., Botstein, D. A manic depressive history. Nature Genet. 12: 351-352, 1996. [PubMed: 8630482] [Full Text: https://doi.org/10.1038/ng0496-351]

  99. Rotondo, A., Mazzanti, C., Dell'Osso, L., Rucci, P., Sullivan, P., Bouanani, S., Gonnelli, C., Goldman, D., Cassano, G. B. Catechol O-methyltransferase, serotonin transporter, and tryptophan hydroxylase gene polymorphisms in bipolar disorder patients with and without comorbid panic disorder. Am. J. Psychiat. 159: 23-29, 2002. [PubMed: 11772685] [Full Text: https://doi.org/10.1176/appi.ajp.159.1.23]

  100. Rybakowski, J., Borkowska, A., Skibinska, M., Hauser, J. Illness-specific association of val66met BDNF polymorphism with performance on Wisconsin Card Sorting Test in bipolar disorder. (Letter) Molec. Psychiat. 11: 122-124, 2006. [PubMed: 16222333] [Full Text: https://doi.org/10.1038/sj.mp.4001765]

  101. Schulze, T. G., Buervenich, S., Badner, J. A., Steele, C. J. M., Detera-Wadleigh, S. D., Dick, D., Foroud, T., Cox, N. J., MacKinnon, D. F., Potash, J. B., Berrettini, W. H., Byerley, W., Coryell, W., DePaulo, J. R., Jr., Gershon, E. S., Kelsoe, J. R., McInnis, M. G., Murphy, D. L., Reich, T., Scheftner, W., Nurnberger, J. I., Jr., McMahon, F. J. Loci on chromosomes 6q and 6p interact to increase susceptibility to bipolar affective disorder in the National Institute of Mental Health Genetics Initiative pedigrees. Biol. Psychiat. 56: 18-23, 2004. [PubMed: 15219468] [Full Text: https://doi.org/10.1016/j.biopsych.2004.04.004]

  102. Segurado, R., Detera-Wadleigh, S. D., Levinson, D. F., Lewis, C. M., Gill, M., Nurnberger, J. I., Jr., Craddock, N., DePaulo, J. R., Baron, M., Gershon, E. S., Ekholm, J., Cichon, S., and 66 others. Genome scan meta-analysis of schizophrenia and bipolar disorder, Part III: bipolar disorder. Am. J. Hum. Genet. 73: 49-62, 2003. [PubMed: 12802785] [Full Text: https://doi.org/10.1086/376547]

  103. Sherrington, R., Curtis, D., Brynjolfsson, J., Moloney, E., Rifkin, L., Petursson, H., Gurling, H. A linkage study of affective disorder with DNA markers for the ABO-AK1-ORM linkage group near the dopamine beta hydroxylase gene. Biol. Psychiat. 36: 434-442, 1994. [PubMed: 7811839] [Full Text: https://doi.org/10.1016/0006-3223(94)90638-6]

  104. Sklar, P., Gabriel, S. B., McInnis, M. G., Bennett, P., Lim, Y.-M., Tsan, G., Schaffner, S., Kirov, G., Jones, I., Owen, M., Craddock, N., DePaulo, J. R., Lander, E. S. Family-based association study of 76 candidate genes in bipolar disorder: BDNF is a potential risk locus. Molec. Psychiat. 7: 579-593, 2002. [PubMed: 12140781] [Full Text: https://doi.org/10.1038/sj.mp.4001058]

  105. Smeraldi, E., Negri, F., Melica, A. M., Scorza-Smeraldi, R. HLA system and affective disorders: a sibship genetic study. Tissue Antigens 12: 270-274, 1978. [PubMed: 725908] [Full Text: https://doi.org/10.1111/j.1399-0039.1978.tb01334.x]

  106. Smyth, C., Kalsi, G., Curtis, D., Brynjolfsson, J., O'Neill, J., Rifkin, L., Moloney, E., Murphy, P., Petursson, H., Gurling, H. Two-locus admixture linkage analysis of bipolar and unipolar affective disorder supports the presence of susceptibility loci on chromosomes 11p15 and 21q22. Genomics 39: 271-278, 1997. [PubMed: 9119364] [Full Text: https://doi.org/10.1006/geno.1996.4486]

  107. Stancer, H. C., Weitkamp, L. R., Persad, E., Flood, C., Jorna, T., Guttormsen, S. A., Yagnow, R. L. Confirmation of the relationship of HLA (chromosome 6) genes to depression and manic depression. II. The Ontario follow-up and analysis of 117 kindreds. Ann. Hum. Genet. 52: 279-298, 1988. [PubMed: 3268040] [Full Text: https://doi.org/10.1111/j.1469-1809.1988.tb01108.x]

  108. Stine, O. C., Xu, J., Koskela, R., McMahon, F. J., Gschwend, M., Friddle, C., Clark, C. D., McInnis, M. G., Simpson, S. G., Breschel, T. S., Vishio, E., Riskin, K., Feilotter, H., Chen, E., Shen, S., Folstein, S., Meyers, D. A., Botstein, D., Marr, T. G., DePaulo, J. R. Evidence for linkage of bipolar disorder to chromosome 18 with a parent-of-origin effect. Am. J. Hum. Genet. 57: 1384-1394, 1995. [PubMed: 8533768]

  109. Stratakis, C. A., Chrousos, G. P. Neuroendocrinology and pathophysiology of the stress system. Ann. N.Y. Acad. Sci. 771: 1-18, 1995. [PubMed: 8597390] [Full Text: https://doi.org/10.1111/j.1749-6632.1995.tb44666.x]

  110. Stratakis, C. A., Sarlis, N. J., Berrettini, W. H., Badner, J. A., Chrousos, G. P., Gershon, E. S., Detera-Wadleigh, S. D. Lack of linkage between the corticotropin-releasing hormone (CRH) gene and bipolar affective disorder. Molec. Psychiat. 2: 483-485, 1997. [PubMed: 9399692] [Full Text: https://doi.org/10.1038/sj.mp.4000268]

  111. Straub, R. E., Lehner, T., Luo, Y., Loth, J. E., Shao, W., Sharpe, L., Alexander, J. R., Das, K., Simon, R., Fieve, R. R., Lerer, B., Endicott, J., Ott, J., Gilliam, T. C., Baron, M. A possible vulnerability locus for bipolar affective disorder on chromosome 21q22.3. Nature Genet. 8: 291-296, 1994. [PubMed: 7874172] [Full Text: https://doi.org/10.1038/ng1194-291]

  112. Tanna, V. L., Wilson, A. F., Winokur, G., Elston, R. C. Linkage analysis of pure depressive disease. J. Psychiat. Res. 23: 99-107, 1989. [PubMed: 2585349] [Full Text: https://doi.org/10.1016/0022-3956(89)90001-0]

  113. Tsutsumi, T., Holmes, S. E., McInnis, M. G., Sawa, A., Callahan, C., DePaulo, J. R., Ross, C. A., DeLisi, L. E., Margolis, R. L. Novel CAG/CTG repeat expansion mutations do not contribute to the genetic risk for most cases of bipolar disorder or schizophrenia. Am. J. Med. Genet. 124B: 15-19, 2004. [PubMed: 14681907] [Full Text: https://doi.org/10.1002/ajmg.b.20058]

  114. Venken, T., Claes, S., Sluijs, S., Paterson, A. D., van Duijn, C., Adolfsson, R., Del-Favero, J., Van Broeckhoven, C. Genomewide scan for affective disorder susceptibility loci in families of a northern Swedish isolated population. Am. J. Hum. Genet. 76: 237-248, 2005. [PubMed: 15614721] [Full Text: https://doi.org/10.1086/427836]

  115. Vlachadis, N., Vrachnis, N., Economou, E. Variant GADL1 and response to lithium in bipolar I disorder. (Letter) New Eng. J. Med. 370: 1856 only, 2014. [PubMed: 24806174] [Full Text: https://doi.org/10.1056/NEJMc1401817]

  116. Weitkamp, L. R., Stancer, H. C., Persad, E., Flood, C., Guttormsen, S. Depressive disorders and HLA: a gene on chromosome 6 that can affect behavior. New Eng. J. Med. 305: 1301-1306, 1981. [PubMed: 6945479] [Full Text: https://doi.org/10.1056/NEJM198111263052201]

  117. Weitkamp, L. R., Stancer, H. C. Analysis of the Toronto-Rochester depression study follow-up data confirms an HLA-region gene contribution to susceptibility to affective disorder. Genet. Epidemiol. 6: 305-310, 1989. [PubMed: 2499504] [Full Text: https://doi.org/10.1002/gepi.1370060153]

  118. Weitkamp, L. R. HLA and disease: predictions for HLA haplotype sharing in families. Am. J. Hum. Genet. 33: 776-784, 1981. [PubMed: 7294026]

  119. Weitkamp, L. R. Genetic analysis of multifactorial diseases. (Letter) Am. J. Hum. Genet. 35: 130-133, 1983. [PubMed: 6823971]

  120. Williams, R. S. B., Cheng, L., Mudge, A. W., Harwood, A. J. A common mechanism of action for three mood-stabilizing drugs. Nature 417: 292-295, 2002. [PubMed: 12015604] [Full Text: https://doi.org/10.1038/417292a]

  121. Wilson, A. F., Elston, R. C., Mallott, D. B., Tran, L. D., Winokur, G. The current status of genetic linkage studies of alcoholism and unipolar depression. Psychiat. Genet. 2: 107-124, 1991.

  122. Wilson, A. F., Tanna, V. L., Winokur, G., Elston, R. C., Hill, E. M. Linkage analysis of depression spectrum disease. Biol. Psychiat. 26: 163-175, 1989. [PubMed: 2736265] [Full Text: https://doi.org/10.1016/0006-3223(89)90020-6]

  123. Wingerson, L. Searching for depression genes. Discover 3(2): 60-64, 1982.

  124. Wright, A. F., Crichton, D. N., Loudon, J. B., Morten, J. E. N., Steel, C. M. Beta-adrenoceptor binding defects in cell lines from families with manic-depressive disorder. Ann. Hum. Genet. 48: 201-214, 1984. [PubMed: 6087716] [Full Text: https://doi.org/10.1111/j.1469-1809.1984.tb01016.x]


Contributors:
Cassandra L. Kniffin - updated : 1/28/2015
Ada Hamosh - updated : 6/10/2014
Ada Hamosh - updated : 2/6/2014
Ada Hamosh - updated : 3/16/2010
Cassandra L. Kniffin - updated : 8/25/2009
Cassandra L. Kniffin - updated : 7/29/2009
Ada Hamosh - updated : 10/22/2008
Carol A. Bocchini - updated : 9/8/2008
Victor A. McKusick - updated : 4/14/2008
Victor A. McKusick - updated : 10/10/2007
Ada Hamosh - updated : 7/19/2007
Victor A. McKusick - updated : 2/26/2007
John Logan Black, III - updated : 1/23/2007
John Logan Black, III - updated : 11/16/2006
John Logan Black, III - updated : 11/15/2006
John Logan Black, III - updated : 6/7/2006
John Logan Black, III - updated : 6/2/2006
John Logan Black, III - updated : 5/25/2006
John Logan Black, III - updated : 5/16/2006
John Logan Black, III - updated : 4/21/2006
John Logan Black, III - updated : 4/6/2006
John Logan Black, III - updated : 4/6/2006
John Logan Black, III - updated : 12/6/2005
Victor A. McKusick - updated : 10/5/2005
John Logan Black, III - updated : 8/8/2005
John Logan Black, III - updated : 7/26/2005
John Logan Black, III - updated : 7/22/2005
John Logan Black, III - updated : 7/21/2005
John Logan Black, III - updated : 7/20/2005
George E. Tiller - updated : 5/5/2005
John Logan Black, III - updated : 4/8/2005
John Logan Black, III - updated : 3/2/2005
Cassandra L. Kniffin - updated : 10/28/2004
Victor A. McKusick - updated : 8/12/2004
John Logan Black, III - updated : 10/30/2003
John Logan Black, III - updated : 8/19/2003
Victor A. McKusick - updated : 7/9/2003
John Logan Black, III - updated : 11/8/2002
Michael B. Petersen - updated : 11/7/2002
Victor A. McKusick - updated : 10/9/2002
Dawn Watkins-Chow - updated : 8/23/2002
George E. Tiller - updated : 8/15/2002
Victor A. McKusick - updated : 7/31/2002
Ada Hamosh - updated : 5/29/2002
Victor A. McKusick - updated : 1/30/2002
Victor A. McKusick - updated : 11/1/2001
Victor A. McKusick - updated : 9/12/2001
Michael B. Petersen - updated : 7/17/2001
Victor A. McKusick - updated : 2/26/2001
Victor A. McKusick - updated : 1/12/2000
Victor A. McKusick - updated : 3/1/1999
Victor A. McKusick - updated : 2/8/1999
Victor A. McKusick - updated : 5/13/1998
Victor A. McKusick - updated : 4/23/1997
Victor A. McKusick - updated : 2/20/1997

Creation Date:
Victor A. McKusick : 6/4/1986

Edit History:
alopez : 09/30/2022
carol : 09/25/2020
carol : 01/22/2020
carol : 07/18/2017
alopez : 07/17/2017
carol : 07/21/2015
alopez : 1/30/2015
ckniffin : 1/28/2015
alopez : 6/10/2014
carol : 4/11/2014
alopez : 2/6/2014
carol : 9/23/2011
alopez : 11/10/2010
wwang : 4/6/2010
alopez : 3/18/2010
alopez : 3/18/2010
terry : 3/16/2010
wwang : 9/8/2009
wwang : 9/4/2009
ckniffin : 8/25/2009
ckniffin : 7/29/2009
mgross : 7/17/2009
carol : 5/1/2009
carol : 12/2/2008
carol : 10/28/2008
terry : 10/22/2008
carol : 9/8/2008
carol : 5/7/2008
carol : 5/7/2008
alopez : 4/30/2008
terry : 4/14/2008
alopez : 10/15/2007
terry : 10/10/2007
alopez : 7/24/2007
terry : 7/19/2007
carol : 5/10/2007
terry : 5/9/2007
alopez : 3/8/2007
terry : 2/26/2007
carol : 1/23/2007
carol : 1/23/2007
carol : 11/16/2006
terry : 11/16/2006
carol : 11/16/2006
terry : 11/15/2006
carol : 11/13/2006
wwang : 6/7/2006
wwang : 6/7/2006
terry : 6/2/2006
wwang : 5/25/2006
wwang : 5/16/2006
carol : 4/25/2006
terry : 4/21/2006
carol : 4/11/2006
wwang : 4/10/2006
terry : 4/6/2006
terry : 4/6/2006
terry : 12/13/2005
carol : 12/6/2005
wwang : 11/7/2005
alopez : 10/6/2005
terry : 10/5/2005
carol : 10/4/2005
terry : 8/8/2005
carol : 7/28/2005
terry : 7/26/2005
carol : 7/25/2005
terry : 7/22/2005
carol : 7/21/2005
terry : 7/21/2005
carol : 7/21/2005
terry : 7/20/2005
tkritzer : 5/5/2005
mgross : 4/8/2005
mgross : 4/5/2005
tkritzer : 3/2/2005
tkritzer : 11/1/2004
ckniffin : 10/28/2004
terry : 8/12/2004
alopez : 3/17/2004
carol : 10/30/2003
carol : 10/13/2003
carol : 9/9/2003
carol : 8/20/2003
terry : 8/19/2003
terry : 8/15/2003
tkritzer : 7/23/2003
tkritzer : 7/15/2003
terry : 7/9/2003
cwells : 5/21/2003
carol : 11/21/2002
carol : 11/8/2002
cwells : 11/7/2002
carol : 10/11/2002
tkritzer : 10/10/2002
terry : 10/9/2002
mgross : 10/9/2002
tkritzer : 8/23/2002
cwells : 8/15/2002
mgross : 8/5/2002
terry : 7/31/2002
alopez : 5/30/2002
terry : 5/29/2002
terry : 3/8/2002
alopez : 2/5/2002
terry : 1/30/2002
carol : 11/5/2001
terry : 11/1/2001
mcapotos : 9/18/2001
mcapotos : 9/12/2001
carol : 7/17/2001
mgross : 7/17/2001
mgross : 7/17/2001
mgross : 2/26/2001
carol : 5/4/2000
terry : 4/24/2000
mgross : 2/7/2000
terry : 1/12/2000
carol : 8/13/1999
carol : 7/16/1999
carol : 6/25/1999
carol : 6/25/1999
terry : 4/30/1999
carol : 3/22/1999
terry : 3/1/1999
carol : 2/14/1999
terry : 2/8/1999
dkim : 12/10/1998
alopez : 5/19/1998
terry : 5/13/1998
terry : 11/11/1997
terry : 11/11/1997
terry : 11/11/1997
terry : 11/10/1997
alopez : 7/29/1997
terry : 7/28/1997
alopez : 7/7/1997
carol : 6/23/1997
alopez : 4/23/1997
terry : 4/23/1997
alopez : 3/20/1997
alopez : 3/19/1997
mark : 2/20/1997
terry : 2/12/1997
terry : 1/16/1997
jamie : 1/15/1997
terry : 1/8/1997
terry : 12/30/1996
terry : 12/20/1996
terry : 12/10/1996
terry : 12/5/1996
mark : 4/3/1996
terry : 4/3/1996
mark : 3/21/1996
terry : 3/19/1996
mark : 3/15/1996
mark : 3/15/1996
mark : 3/14/1996
mark : 3/1/1996
terry : 2/29/1996
mark : 12/15/1995
terry : 12/14/1995
mark : 9/10/1995
terry : 3/3/1995
carol : 3/1/1995
jason : 7/12/1994
mimadm : 6/25/1994
warfield : 4/8/1994