Next Article in Journal
Finding Predictors of Leg Defects in Pigs Using CNV-GWAS
Previous Article in Journal
Evolution of the Critically Endangered Green Sawfish Pristis zijsron (Rhinopristiformes, Pristidae), Inferred from the Whole Mitochondrial Genome
Previous Article in Special Issue
Unraveling the Genetic Threads of History: mtDNA HVS-I Analysis Reveals the Ancient Past of the Aburra Valley
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Signatures of Positive Selection in Human Populations of the OXT, OXTR, AVP, AVPR1A and AVR1B Gene Variants Related to the Regulation of Psychoemotional Response

by
Siroj Yu. Bakoev
*,
Anna V. Korobeinikova
,
Arina I. Mishina
,
Shuanat Sh. Kabieva
,
Sergey I. Mitrofanov
,
Alexey A. Ivashechkin
,
Alexsandra I. Akinshina
,
Ekaterina A. Snigir
,
Sergey M. Yudin
,
Vladimir S. Yudin
,
Lyubov V. Getmantseva
and
Elmira A. Anderzhanova
Federal State Budgetary Institution “Centre for Strategic Planning and Management of Biomedical Health Risks” of the Federal Medical Biological Agency (Centre for Strategic Planning of FMBA of Russia), Pogodinskaya Street, 10, Bld. 1, 119121 Moscow, Russia
*
Author to whom correspondence should be addressed.
Genes 2023, 14(11), 2053; https://doi.org/10.3390/genes14112053
Submission received: 29 September 2023 / Revised: 3 November 2023 / Accepted: 4 November 2023 / Published: 8 November 2023
(This article belongs to the Special Issue Population Structure and Human Genetic Diversity)

Abstract

:
The neurobiological systems of maintenance and control of behavioral responses result from natural selection. We have analyzed the selection signatures for single nucleotide variants (SNV) of the genes of oxytocin (OXT, OXTR) and vasopressin (AVP, AVPR1A, AVPR1B) systems, which are associated with the regulation of social and emotional behavior in distinct populations. The analysis was performed using original WGS (whole genome sequencing) data on Eastern Slavs (SlEast), as well as publicly available data from the 1000 Genomes Project on GBR, FIN, IBR, PUR, BEB, CHB, and ACB populations (the latter were taken as reference). To identify selection signatures, we rated the integrated haplotype scores (iHS), the numbers of segregating sites by length (nSl), and the integrated haplotype homozygosity pooled (iHH12) measures; the fixation index Fst was implemented to assess genetic differentiation between populations. We revealed that the strongest genetic differentiation of populations was found with respect to the AVPR1B gene, with the greatest differentiation observed in GRB (Fst = 0.316) and CHB (Fst = 0.325) in comparison to ACB. Also, high Fst values were found for SNVs of the AVPR1B gene rs28499431, rs33940624, rs28477649, rs3883899, and rs28452187 in most of the populations. Selection signatures have also been identified in the AVP, AVPR1A, OXT, and OXTR genes. Our analysis shows that the OXT, OXTR, AVP, AVPR1A, and AVPR1B genes were subject to positive selection in a population-specific process, which was likely contributing to the diversity of adaptive emotional response types and social function realizations.

1. Introduction

Identification of positive selection genomic signals is a traditional population genetics approach based on the analysis of results of whole genome sequencing (WGS) from the perspective of comparison between species and populations. The signals indicate genetic variants corresponding to different adaptation models and, therefore, designate loci of phenotypic variability, thus appearing as objects of particular interest when mapping complex traits [1,2,3].
The driving forces and direction of selection during the evolution process are not always clear. Nevertheless, the development and human populations’ growth upon the settlement of Homo sapiens in various landscape and climatic zones [4] came to a realization through the accumulation of the variants that are most suitable for new environments and contributing to the reinforcement of specific patterns of cooperative behavior and offspring care [5]. Many of the selected variants can affect complex endophenotypes (characteristic sets of behavioral signs with a strong genetic component), which represent constellations of normal or altered emotional, social, and cognitive functions [6] and which, among others, can be determined by gene pleiotropy, hysteresis, and other phenomena that define the complex trait’s variability.
At present, there is growing interest in understanding the genetic basis for emotional and sociocognitive functions in s [7,8,9,10]. With regard to social and emotional behavior, the neuropeptides oxytocin (OXT), vasopressin (AVP) and their receptors (AVPR1A, AVPR1B, and OXTR) gain particular attention [11,12] because the role of OXT and AVP in the mechanisms of establishing and maintaining social interaction, social memory, empathy, prosociality, and anxiety is well recognized [11]. Mechanisms of psychoemotional adaptation are gaining attention due to the standing problem of acute and chronic stress and the aging of the general population. Therefore, both genetic factors and molecular-biological mechanisms of stress resilience and susceptibility to environmental changes evoke persistent interest. A review of the molecular mechanism of oxytocin and vasopressin action can be found in [13,14].
It is of note that the OXT and AVP genes show almost no variability in their sequences, while the genes encoding their receptors are characterized by a high degree of variability (the high number of SNVs associated with phenotypic peculiarities). Therefore, the study on variability of the AVPR1A, AVPR1B, and OXTR genes makes a great contribution to understanding the genetic architecture of behavioral traits associated with social behavior and emotions [15]. It has already been shown that genetic variants in the human AVPR1A, AVPR1B, and OXTR genes are associated with the diversity of perception of social signals (empathy, emotional intelligence, eye contact, face “reading,” autism), prosociality (trust, altruism, generosity, loyalty), social tolerance (aggressiveness, competition), and anxiety [7,8,16,17] that contributes to the variability of personality types appearing in human population. In turn, evaluation of the positively selected variants is of particular interest because it allows one to identify functionally significant SNVs contributing to the formation of psychopathological phenotypes.
The signaling mediated by oxytocin and vasopressin is essential not only in the regulation of physiological functions, among which labor in females and reabsorption of water in the kidney, respectively, are most prominent. Polymorphisms in genes for oxytocin and vasopressin and their specific receptors are associated with diversity in traits related to the behavioral functions of positive and negative valence and general arousal. They are significant in setting the cumulative risk for autism spectrum disorders (ASD), schizophrenia, stress susceptibility and panic disorder, major depression, and aggression [18,19,20,21,22,23,24]. The high biomedical significance of both signaling systems mediating psychoemotional functions and social interactions supports a persistent interest in the genetics of these two systems.
Identification of positive selection genomic signals in distinct populations would be helpful in revealing genetic mechanisms of behavioral traits. The study was based on the assumption that diversity in the climate/geographical zones is a factor of difference in the scenarios of positive gene selection during Homo sapiens migration. Therefore, the genetic diversity within the genes of the oxytocin and vasopressin systems across different human populations was our primary interest. In this study, we identified selection signatures in the OXT, OXTR, AVP, AVPR1A, and AVPR1B genes in different human populations with population genetic models.

2. Materials and Methods

2.1. Characteristics of Sample

An original sample of 50 individuals from the East Slavic subethnic group (SlEast) was stratified based on the criteria, which were set upon results of the principal component analysis (PCA) of the 1000 Genomes consortium data (Figure S1) [25]. In addition, we performed the cluster analysis by admixture to show the ancestry proportion of SIEast versus the 1000 Genome samples (Figure S2).
WGS data for Afro-Caribbeans in Barbados (ACB), British from England and Scotland (GBR), Finns in Finland (FIN), Iberian population in Spain (IBR), Puerto Ricans in Puerto Rico (PUR), Bengalis in Bangladesh (BEB), and Han Chinese in Beijing (CHB) (50 individuals in each ethnic group) were from the 1000 Genomes Consortium public database. Fifty individuals from each ethnic group were selected in a random order from the general pool of data.

2.2. DNA Isolation, Preparation of Sequencing Libraries and WGS of Whole Blood Samples

Isolation of genomic DNA (gDNA) from whole blood samples was performed manually using a DNA Blood Mini Kit (Qiagen, Hilden, Germany) in accordance with the manufacturer’s protocol. The yield and purity of the isolated gDNA were manually determined using Qubit 4.0 fluorimeter (Thermo Fisher Scientific, Waltham, MA, USA) and NanoDropTM One C Microvolume UV-Vis (Thermo Fisher Scientific, MA, Waltham, USA), respectively. Only gDNA samples with absorbance ratios A260/280 of 1.7–1.9 and A230/260 of 1.8–2.2, in accordance, were selected for further analysis.
The amount of 150–500 ng of gDNA was used to prepare Next-Generation Sequencing (NGS) libraries. Libraries were prepared using the Illumina DNA Prep kit (Illumina, Inc., San Diego, CA, USA) according to the manufacturer’s recommendations using the Tecan Freedom EVO robotic station (Tecan, Männedorf, Switzerland). The gDNA concentrations in library samples were measured using the Infinite F Nano Plus tablet reader (Tecan, Männedorf, Switzerland). The size of the resulting libraries was determined with an Agilent D1000 reagent kit using an Agilent 4200 TapeStation (Agilent Technologies, Inc., Santa Clara, CA, USA). Pooling was performed automatically using a Tecan Freedom EVO robotic station (Tecan, Männedorf, Switzerland). Each of the 50 samples of the library pool was diluted to the final gDNA concentration of 1.5 nM prior to sequencing. Pool quality control was performed with an Agilent High Sensitivity D1000 Screen Tape reagent kit using an Agilent 4200 TapeStation (Agilent Technologies, Inc., Santa Clara, CA, USA).
WGS was performed on an Illumina NovaSeq 6000 (Illumina, Inc., San Diego, CA, USA) with an S4 reagent kit (Illumina, Inc., San Diego, CA, USA) upon 300 cycles with 2 × 150 bp paired-end reads with at least 30× average depth of coverage (>350 million reads). The Illumina Dragen Bio-IT platform (Illumina, San Diego, CA, USA) was used to align reads to the reference genome (GRCh38).

2.3. Population Genetic Analysis

The choice of analytical approaches was based on their specificity to spot various characteristics of the selection signal. In general, selection signals are subdivided into two types: “hard” selection, in which de novo mutations are fixed and the frequency of specific haplotypes is significantly increased in a population, therefore eliminating genetic diversity in the vicinity of the adaptive locus [26]; and “soft” selection, when selection acts toward the previously neutral alleles or allele frequency increases, resulting in a higher frequency of a specific set of haplotypes [1,25]. Respectively, we used the following statistics to look for regions of recent and/or ongoing positive selection. (1) iHS (integrated haplotype score) identified loci in which a strong selection led to the fixation of new alleles and to the decrease in haplotype homozygosity (alleles that may be on their way to fixation or become a balanced polymorphism) [1]. (2) nSl (number of segregating sites by length), which generally serves the same purposes as iHS, allowing one to detect hard and soft selective sweeps, appears more stable, especially with regard to the difference in the rate of recombination and the influence of population factors such as population splits, population bottlenecks (when a sharp decrease in the size of the population results in a significant drop in the genetic variability), and population growth without requiring a genetic map [8]. (3) iHH12 (integrated haplotype homozygosity pooled), which was developed based on H12 statistics [27], is able to scan soft selective sweeps by adapting H12 into an integrated haplotype homozygosity structure [28]. Finally, (4) the fixation index Fst was used to assess the genetic differentiation of populations [4]. Because a local adaptation of a population can appear even at conditions of low genomic differentiation, a comparison of genetic variability between two populations increases the power of detection for positive selection, in contrast to what can be revealed at the level of a single population.
Quality control, cohort pooling, and removal of duplicate SNVs were performed in accordance [29]. Data for all 400 (eight populations, each represented by 50 individuals) samples were combined using the plink1.9 program [30]. bcftools software [31] was used to remove duplicate SNVs with identical positions. Duplicates were removed to provide effective processing on the next steps of analysis. After variant calling, variants where more than 5% or 10% of the genotype calls were not available, were removed with geno0.05 and mind0.1 tools, respectively. Hard-filtering of rare minor alleles was performed with maf0.05. Within-group data pooling was performed after removing multiallelic SNVs and correcting allele flipping. All data were normalized to the GRCh38 reference. The start and end positions for the OXT, OXTR, AVP, AVPR1A, and AVPR1B gene regions (GRCh 38 assembly) were obtained from the NCBI database (accessed on 15 August 2022) [32].
Genetic variability was assessed in every population to weight intrapopulation variability. The iHS, nSl, and iHH12 approaches for the detection of selection signals were implemented using the selscan program [33]. SNVs with a minor allele frequency (MAF) > 5% were kept for analysis, with each SNV sequentially treated as a major SNV. Non-standardized scores were normalized using the “norm” script of the selscan program. Windows with crit = 1 values were considered as candidate regions for selection.
We have compared a number of different populations with the ACB population aiming at the evaluation of changes in genomes, which were accompanying the migration of Homo sapiens from Africa and its accustoming to new regions of settling. The genetic differentiation of the ACB and SlEast, ACB and GBR, A and CB and FIN, ACB and IBR, ACB and PUR, ACB and BEB, and ACB and CHB populations for the OXT, AVP, AVPR1A, AVPR1B, OXTR genes was calculated based on the pairwise Fst method proposed by Weir & Cockerham, which was implemented in the VCFTools program [34]. A difference was considered significant at Fst ≥ 0.3, which corresponds to the quantile level of 0.99 [35].

3. Results

3.1. Overall Estimation of Population Differentiation Based on the Analysis of Genes of Oxytocin and Vasopressin Systems

The greatest genetic population differentiation was observed with respect to the AVPR1B gene. The average Fst value for all analyzed populations was of 0.277 with the differentiation of GRB to ACB and CHB to ACB of Fst = 0.316 and Fst = 0.325, respectively. Selection signatures have also been identified in the AVP, AVPR1A, OXT, and OXTR genes. The Fst values of population differentiation averaged among all variants of the OXT, OXTR, AVP, AVPR1A, and AVPR1B genes for each individual locus are presented in Table 1.

3.2. Population Differentiation with Respect to AVP, AVPR1A, and AVPR1B Genes

The strongest genetic population differentiation was found in the case of the AVPR1B gene. For the GRB to ACB and CHB to ACB populations, the mean Fst for all AVPR1B gene variants was 0.321. At the same time, rs28499431, rs33940624, rs28477649, rs3883899, and rs28452187 showed the highest Fst values in all populations under analysis (Table 2). As the frequency of the preferred allele rises rapidly, the classic strong directional selection signal tends to be localized to an unusually long invariable haplotype. According to the results of the iHS analysis, all SNVs of the AVPR1B gene were subject to positive selection in the GBR populations (Table 2). A similar phenomenon was observed in the FIN, IBS, and PUR populations, in which almost all SNVs of the AVPR1B gene showed a positive selection signature. In contrast, four positively selected SNVs were identified in the SlEast population, and only one SNV was identified in the BEB, CHB, and ACB populations.
Based on the results of nSl analysis, the selection was confirmed for all SNVs of the AVPR1B gene (with the exception of rs28588803 in GBR, which showed a positive signal detected by the iHS method). Additional SNVs have been identified in all populations except ACB.
Based on an assessment of iHH12, selection signals in the AVPR1B gene were identified only in the GBR population. In other populations, positive selection signals were not detected by the iHH12 method (Table 2).
Variants in the AVPR1A gene also showed significant multipopulational divergence relative to the ACB population. Two variants, rs10784339 and rs10747983, have been identified in the SlEast, GBR, IBS, and PUR populations; two variants, rs11829406 and rs11829452, in the SlEast, IBS, PUR, BEB, and CHB populations and in the SlEast, PUR, BEB, and CHB populations, respectively; and the rs10047514 variant in the GBR and FIN populations (Table 3). Selection traits in the AVPR1A gene were found only in the GBR population; two variants (rs11829406 and rs11829452) were identified by the iHS and nSL methods, and four variants (rs10747983, rs10784339, rs10877968, rs1565878685) by the nSL method (Table 3).
The analysis of Fst indicators revealed the absence of significant differentiation signals between ACB and other populations in relation to the AVP gene (Table 4). At the same time, a selection signal was found in the ACB population in the AVP gene (variant rs3787482) (Table 4).

3.3. Population Differentiation with Respect to OXT and OXTR Genes

Analysis of SNVs of the OXTR gene generally did not reveal differentiation; however, some variants showed rather high Fst values. Specifically, the levels of differentiation between ACB and CHB in relation to rs2324728l, rs237884, rs1042778, and rs237895, Fst were above 0.5, and between ACB and BEB in relation to rs2324728l and rs237884, Fst reached 0.5 (Table 5). In addition, rs59190448 has been identified in all populations except PUR.
The selection signals identified in the OXTR gene were found in all populations except CHB. The largest number of positively selected SNVs (13 SNVs) was found in the PUR population using the iHS method (Table 5). Three SNVs were identified in the SlEast and ACB populations using the iHS method. At the same time, the rs34992398 variant showed positive selection in the PUR, SlEast, and ACB populations. The rs918316 variant was detected in the GBR population using the iHS method. This variant was also present in the IBS population; however, here, it was discovered using the nSl method. Two SNVs were identified in the BEB population based on the nSl approach. Positive selection of the rs237888 variant was detected in the FIN population by the iHS and nSl methods.
Analysis of Fst scores revealed the absence of significant differentiation signals between ACB and other populations in relation to the OXT gene (Table 4). Of note, a selection signal was also found in the OXT gene; the rs111869749 variant was found in the GBR population (Table 4).

4. Discussion

The selection of specific genetic structures in human populations occurs under the pressure of various factors. In general, it aims at the fixation of adaptive changes and, in particular, at the consolidation of complex behavioral programs, which conduce the survival of the individual and the viability of an entire population. Genetic studies make a prodigious contribution not only to understanding the evolution of behavioral traits but also to identifying functional gene variants that influence behavior such as caregiving, tribal care, and cooperation, as well as competition, aggression, and stress response.
In our study, we showed that genes of the oxytocin and vasopressin systems are objects of positive selection in the population process. The selection evolves individually in each population and provides a variety of behavioral reactions. The significance of neuropeptides OXT and AVP and their receptors OXTR, AVPR1A, and AVPR1B is determined by the role they play in the regulation of social interaction and stress response, as well as in maintaining water homeostasis and carbohydrate metabolism. Respectively, the selection in the genes of the oxytocin and vasopressin systems can be forced by a combined driving force.
We found that variants of the OXT, OXTR, AVP, AVPR1A, and AVPR1B genes showed positive selection signals in populations belonging to Indo-European ethnic groups (SlEast, GBR, FIN, IBR, BEB), in the Han subpopulation (CHB), and in Puerto Ricans of Puerto Rico (PUR), who are of a mixed ethnic origin. Overall, the genetic differentiation of the scrutinized populations was stable, as evidenced by the high Fst fixation scores for the analyzed genes in all populations. It is noteworthy that the selection of SNVs of the AVPR1B gene was the most pronounced. SNVs of other genes showed selection signs that were characteristic only for selected populations.
The observed high differentiation of populations in relation to the AVPR1B gene confirms the results of the recent study [36], which indicated a positive selection for the AVPR1B gene both in the phylogenetic lineage of human primates and among populations of White Caucasians, African Americans, Yoruba Africans, and East Asians. Taking into account the role of the AVPR1B receptor in insulin secretion [37,38], the presence of significant positive selection for this gene in representatives of GBR, FIN, IBS, and PUR may be related to the peculiarities of the regulatory hormonal mechanisms of glucose utilization in African Americans and European Americans of European descent. The most notable one is the high secretion of insulin in African Americans [38,39,40,41].
The AVPR1B receptor expression has been shown both in peripheral tissues [42] and in the central nervous system [37,43]. AVPR1B receptor is mainly expressed in the pituitary gland in the brain. It mediates the effect of vasopressin on the corticoliberin-stimulated release of adrenocorticotropic hormone (ACTH) there [44] and, therefore, affects the activity of the hypothalamic-pituitary-adrenal (HPA) axis [45]. One of the hypotheses of the etiopathogenesis of stress-associated diseases, which considers activation of neuroinflammation and inhibition of neuroplasticity, is based on the fact of the HPA axis dysregulation and insufficient control over the basal and stimulated cortisol levels in patients [46,47]. Therefore, it is not surprising that SNVs of the AVPR1B gene are increasingly singled out in GWAS studies as candidates for assessing the risk of the development of major depression and other stress-related psychopathologies. In turn, a fixation of peculiarities of the HPA axis activity occurs at a population level. As had been shown, both basal and stress-stimulated cortisol levels were related to ethnic specificity. The flatter diurnal cortisol slopes were found among African Americans and Hispanics in comparison to Caucasians in the US [48]; in addition, stress-stimulated cortisol secretion during public speaking as well as circadian secretion pattern were reduced in African Americans compared to Caucasians [45]. Since the activity of AVPR1B receptors affects the level of ACTH release, differentiation of populations based on the selection of stress responses of different severity can be mediated by the positive selection of AVPR1B gene variants. Indeed, we found that the rs33933482 variant of the AVPR1B gene, which has been shown to be associated with increased cortisol secretion during public speaking stress in Europeans [45], was positively selected in Indo-European and CHB populations. Other positively selected variants of the AVPR1B gene that we found are known to be associated with side effects of vasopressin [49], panic disorders [18,19], aggression in children [20], and autism spectrum disorders (ASD) [21]. Thus, AVPR1B appears to be one of the genes, the variability of which is underlying the recent adaptation of the stress response.
AVPR1A receptor-mediated signaling is deeply involved in the regulation of physiological functions. It is determined by the effects of vasopressin on glycogenolysis in the liver, insulin secretion, contraction of vascular smooth muscle cells, water reabsorption in the renal tubules (antidiuretic activity), and platelet aggregation [50,51]. The importance of the AVPR1A gene in the control of social interaction, altruistic behavior, and pairing in humans has also been documented [36,52,53]. Variants of the AVPR1A gene, which were identified as positively selected in our study, are associated with heroin and general drug addiction (rs10784339) [19], mental and behavioral disorders associated with substance abuse (rs10747983) [54], diabetic status, increased glucose concentration and triglycerides in the blood, and increased body mass index [43,55]. Effects of AVPR1A gene variants related to the function of the nervous system can be interpreted as those which give preference to behaviors that lead to quick results and allow a more liberal risk assessment at any activity [56].
Our analysis of the selection signatures of the AVP, AVPR1A, and AVPR1B genes shows that population differentiation is accompanied by a selection for variants, which supports variability in stress reactivity among individuals and thus determines the diversity of stress-adaptation strategies. The evolutionary strategy of a positive modulation of the stress response, including positive selection in the genes of the vasopressin system, has certain benefits. The intensive response to stress has advantages from a short-term perspective since it supports a quick and balanced response to changing environmental conditions at any physiological level, including the metabolic and immune systems [36]. The directed selection of vasopressin system genes may result from the relatively recent spread of human populations with settlements in strictly different ecosystems.
Our analysis showed that, in contrast to the vasopressin system, the OXT and OXTR genes are main subjects of a neutral evolution, which occurs due to genetic drift. The relative stability of the OXT and OXTR genes in the population process may be due to the exclusive role of this neuropeptide hormone in childbirth (coordinating contraction of myometrial cells), lactation, and maternal behavior. However, we have shown a high level of selection in the OXTR gene in the PUR population; partially selective signals are also found in other populations but not in CHB. Selected OXTR gene variants have been shown to be associated with changes in sensitivity to exogenous oxytocin and contractility of myometrium, as well as (rs4686302) with preterm birth [57,58,59]. However, a significant interaction between the effects of rs4686302 and obesity and gestational diabetes cannot be overruled. Our analysis showed that the rs4686302 variant was identified only in the PUR population. This population is characterized by a preterm birth rate of 11.6%, which is higher than averages of 8.1% for other countries of Latin America and the Caribbean, but close to the average rate of 11.9% for African populations [60]. It should be noted that the actual phenotypic significance of this variant in the PUR population is difficult to ascertain due to the combination of the multifactorial nature of preterm birth and the presence of a large percentage of the West African ethnic component in the PUR population.
The neuropeptide OXT is also involved in the regulation of behavior. It determines the degree of affiliation in interpersonal interactions by increasing the efficiency and functional significance of neural networks involved in the control of social interactions [61]. We have identified several variants of the OXT and OXTR genes that were under directed selection. Among them, the rs53576 OXT gene variant showed an association with the level of empathy; OXT gene variants rs6133010 and rs53576 were shown to be associated with the level of emotional intelligence in Western Slavs [54]. Carriers of the G/G genotype in the OXTR rs1042778 gene variant are less sensitive to perceived family support received in order to interfere with the effect of low maternal emotional warmth and acceptance during early life [62]. The OXTR gene variant rs6770632 is associated with depressive symptoms [22]. In humans, the genetic variability in the OXTR rs2254298 variant determines the volume of the amygdala, a brain structure involved in the control of fear and anxiety. In turn, the OXTR rs53576 gene variant is associated with an increased level of stress-stimulated cortisol secretion [63]. According to the results of our analysis, other positively selected variants of the OXTR gene are associated with aggressive behavior and depressive states [22,23,63], alcohol abuse and associated aggressive behavior [24,64], severe schizophrenia, and autism spectrum disorders [23,24,65]. Although the selection for SNVs of the OXTR gene was presumably influenced by the same factors that determined the effective differentiation of populations with respect to the AVPR1A and AVPR1B genes, it should be noted that the exceptional importance of the oxytocin system in the physiological regulation of labor set apart the corresponding genes to a conservative pool that had determined the low the degree of populations’ differentiation in this gene.
Table 6 summarizes the publications on the functional significance of AVPR1A, AVPR1B, and OXTR gene SNVs.

5. Conclusions

An analysis of the selection signatures of isolated genomic regions of individuals, which was carried out using general population genetic models, allowed us to show positive selection signals in the AVP, AVPR1A, and AVPR1B genes, as well as, to a lesser degree, in the OXT and OXTR genes. Positive selection in these genes was actualized with the individual scenario in each case of population differentiation. High population differentiation was revealed with respect to genes of the vasopressin system and, most pronouncedly, to the AVPR1B gene, with Fst varied from 0.168 to 0.325. The significance of the OXT and OXTR genes in the control of labor (a complex of endocrine, paracrine, and immunological mechanisms providing parturition) and the maintenance of maternal behavior was, possibly, the factor determined their greater stability in population formation, which was shown in our study. In general, our results indicate that during the formation of populations, there was a positive selection towards variants that determine the diversity of emotional stress reactions. Despite the elevated probability of accumulation of individuals prone to increased stress-reactivity upon realization of the effects of these SVNs, such selection may be an advantage since it supports immediate behavioral and physiological changes, which result from a balanced reaction of the neuronal, metabolic, and immune systems upon acute environmental stimuli.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes14112053/s1, Figure S1: Sample stratification was based on the principal component analysis (PCA) of the first two components, which was performed with the plink software and visualized with the ggplot2 (R package); Figure S2: Ancestry analysis by admixture. The number of K ancestral clusters was varied from 2 to 8. Individuals were grouped according to the subpopulation assignments made by PCA (Figure S1).

Author Contributions

S.Y.B. and A.V.K.—Writing—Original Draft Preparation, Methodology, Investigation, Formal Analysis; S.S.K. and A.I.M.—Writing—Review & Editing; S.I.M. and A.I.A.—Data Curation; A.A.I., S.I.M., and E.A.S., Investigation, Formal Analysis; S.M.Y.—Writing—Review and Editing, Resources; V.S.Y.—Writing—Review and Editing, Supervision; L.V.G.—Conceptualization, Methodology, Visualization; E.A.A.—Conceptualization, Writing—Review and Editing, Project Administration, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by the Local Ethics Committee of the Federal State Budgetary Institution “Centre for Strategic Planning and Management of Biomedical Health Risks” of the Federal Medical Biological Agency (Protocol No. 5, 28 December 2020).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data are available on demand.

Acknowledgments

The authors thank Valentina Azaryan for valuable comments on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Voight, B.F.; Kudaravalli, S.; Wen, X.; Pritchard, J.K. A map of recent positive selection in the human genome. PLoS Biol. 2006, 4, e72, Erratum in PLoS Biol. 2006, 4, e154. Erratum in PLoS Biol. 2007, 5, e147. [Google Scholar] [CrossRef]
  2. Hernandez, M.; Perry, G.H. Scanning the human genome for “signatures” of positive selection: Transformative opportunities and ethical obligations. Evol. Anthropol. 2021, 30, 113–121. [Google Scholar] [CrossRef] [PubMed]
  3. da Silva Ribeiro, T.; Galván, J.A.; Pool, J.E. SNP-level FST outperforms window statistics for detecting soft sweeps in local adaptation. bioRxiv 2022. [Google Scholar] [CrossRef]
  4. Thornton, T.F.; Puri, R.K.; Bhagwat, S.; Howard, P. Human adaptation to biodiversity change: An adaptation process approach applied to a case study from southern India. Ambio 2019, 48, 1431–1446. [Google Scholar] [CrossRef]
  5. Abondio, P.; Cilli, E.; Luiselli, D. Inferring Signatures of Positive Selection in Whole-Genome Sequencing Data: An Overview of Haplotype-Based Methods. Genes 2022, 13, 926. [Google Scholar] [CrossRef]
  6. Pittman, Q.J. Vasopressin and central control of the cardiovascular system: A 40-year retrospective. J. Neuroendocrinol. 2021, 33, e13011. [Google Scholar] [CrossRef]
  7. Skuse, D.H.; Gallagher, L. Genetic influences on social cognition. Pediatr. Res. 2011, 69, 85R–91R. [Google Scholar] [CrossRef]
  8. Wilczyński, K.M.; Zasada, I.; Siwiec, A.; Janas-Kozik, M. Differences in oxytocin and vasopressin levels in individuals suffering from the autism spectrum disorders vs general population—A systematic review. Neuropsychiatr. Dis. Treat. 2019, 15, 2613–2620. [Google Scholar] [CrossRef]
  9. Kim, H.S.; Sasaki, J.Y. Emotion Regulation: The Interplay of Culture and Genes. Soc. Personal. Psychol. Compass 2012, 6, 865–877. [Google Scholar] [CrossRef]
  10. Barzman, D.; Geise, C.; Lin, P.I. Review of the genetic basis of emotion dysregulation in children and adolescents. World J. Psychiatry 2015, 5, 112–117. [Google Scholar] [CrossRef]
  11. Insel, T. Rethinking schizophrenia. Nature 2010, 468, 187–193. [Google Scholar] [CrossRef] [PubMed]
  12. Jurek, B.; Neumann, I.D. The Oxytocin Receptor: From Intracellular Signaling to Behavior. Physiol. Rev. 2018, 98, 1805–1908. [Google Scholar] [CrossRef]
  13. Chatterjee, O.; Patil, K.; Sahu, A.; Gopalakrishnan, L.; Mol, P.; Advani, J.; Mukherjee, S.; Christopher, R.; Prasad, T.S. An overview of the oxytocin-oxytocin receptor signaling network. J. Cell Commun. Signal. 2016, 10, 355–360. [Google Scholar] [CrossRef] [PubMed]
  14. Wasilewski, M.A.; Myers, V.D.; Recchia, F.A.; Feldman, A.M.; Tilley, D.G. Arginine vasopressin receptor signaling and functional outcomes in heart failure. Cell. Signal. 2016, 28, 224–233. [Google Scholar] [CrossRef] [PubMed]
  15. Staes, N.; Guevara, E.E.; Helsen, P.; Eens, M.; Stevens, J.M. The Pan social brain: An evolutionary history of neurochemical receptor genes and their potential impact on sociocognitive differences. J. Hum. Evol. 2021, 152, 2021. [Google Scholar] [CrossRef]
  16. Ebstein, R.P.; Israel, S.; Chew, S.H.; Zhong, S.; Knafo, A. Genetics of human social behavior. Neuron 2010, 65, 831–844. [Google Scholar] [CrossRef]
  17. Aspé-Sánchez, M.; Moreno, M.; Rivera, M.I.; Rossi, A.; Ewer, J. Oxytocin and Vasopressin Receptor Gene Polymorphisms: Role in Social and Psychiatric Traits. Front. Neurosci. 2016, 9, 510. [Google Scholar] [CrossRef]
  18. Anantasit, N.; Boyd, J.H.; Walley, K.R.; Russell, J.A. Serious adverse events associated with vasopressin and norepinephrine infusion in septic shock. Crit. Care Med. 2014, 42, 1812–1820. [Google Scholar] [CrossRef] [PubMed]
  19. Kreek, M.J.; Zhou, Y.; Levran, O. Functions of arginine vasopressin and its receptors: Importance of human molecular genetics studies in bidirectional translational research. Biol. Psychiatry 2011, 70, 502–503. [Google Scholar] [CrossRef]
  20. Keck, M.E.; Kern, N.; Erhardt, A.; Unschuld, P.G.; Ising, M.; Salyakina, D.; Müller, M.B.; Knorr, C.C.; Lieb, R.; Hohoff, C.; et al. Combined effects of exonic polymorphisms in CRHR1 and AVPR1B genes in a case/control study for panic disorder. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2008, 147B, 1196–1204. [Google Scholar] [CrossRef]
  21. Zai, C.C.; Muir, K.E.; Nowrouzi, B.; Shaikh, S.A.; Choi, E.; Berall, L.; Trépanier, M.O.; Beitchman, J.H.; Kennedy, J.L. Possible genetic association between vasopressin receptor 1B and child aggression. Psychiatry Res. 2012, 200, 784–788. [Google Scholar] [CrossRef] [PubMed]
  22. Keijser, R.; Åslund, C.; Nilsson, K.W.; Olofsdotter, S. Gene-environment interaction: Oxytocin receptor (OXTR) polymorphisms and parenting style as potential predictors for depressive symptoms. Psychiatry Res. 2021, 303, 114057. [Google Scholar] [CrossRef] [PubMed]
  23. Malik, A.I.; Zai, C.C.; Abu, Z.; Nowrouzi, B.; Beitchman, J.H. The role of oxytocin and oxytocin receptor gene variants in childhood-onset aggression. Genes Brain Behav. 2012, 11, 545–551. [Google Scholar] [CrossRef] [PubMed]
  24. Souza, R.P.; de Luca, V.; Meltzer, H.Y.; Lieberman, J.A.; Kennedy, J.L. Schizophrenia severity and clozapine treatment outcome association with oxytocinergic genes. Int. J. Neuropsychopharmacol. 2010, 13, 793–798. [Google Scholar] [CrossRef]
  25. The 1000 Genomes Project Consortium. Available online: https://www.ncbi.nlm.nih.gov/projects/faspftp/1000genomes/ (accessed on 20 March 2023).
  26. Hermisson, J.; Pennings, P.S. Soft sweeps and beyond: Understanding the patterns and probabilities of selection footprints under rapid adaptation. Methods Ecol. Evol. 2017, 8, 700–716. [Google Scholar] [CrossRef]
  27. Garud, N.R.; Messer, P.W.; Buzbas, E.O.; Petrov, D.A. Recent Selective Sweeps in North American Drosophila melanogaster Show Signatures of Soft Sweeps. PLoS Genet. 2015, 11, e1005004. [Google Scholar] [CrossRef]
  28. Furman, D.J.; Chen, M.C.; Gotlib, I.H. Variant in oxytocin receptor gene is associated with amygdala volume. Psychoneuroen-docrinology 2011, 36, 891–897. [Google Scholar] [CrossRef]
  29. Marees, A.T.; de Kluiver, H.; Stringer, S.; Vorspan, F.; Curis, E.; Marie-Claire, C.; Derks, E.M. A tutorial on conducting ge-nome-wide association studies: Quality control and statistical analysis. Int. J. Methods Psychiatr. Res. 2018, 27, e1608. [Google Scholar] [CrossRef]
  30. PLINK. Available online: http://pngu.mgh.harvard.edu/purcell/plink/ (accessed on 20 April 2023).
  31. Danecek, P.; Bonfield, J.K.; Liddle, J.; Marshall, J.; Ohan, V.; Pollard, M.O.; Whitwham, A.; Keane, T.; McCarthy, S.A.; Davies, R.M.; et al. Twelve years of SAMtools and BCFtools. Gigascience 2021, 10, giab008. [Google Scholar] [CrossRef]
  32. National Center for Biotechnology Information (NCBI)[Internet]. Bethesda (MD): National Library of Medicine (US), National Center for Biotechnology Information; 1988. Available online: https://www.ncbi.nlm.nih.gov/ (accessed on 20 April 2023).
  33. Szpiech, Z. selscan 2.0: Scanning for sweeps in unphased data. bioRxiv 2021. [Google Scholar] [CrossRef]
  34. Danecek, P.; Auton, A.; Abecasis, G.; Albers, C.A.; Banks, E.; DePristo, M.A.; Handsaker, R.E.; Lunter, G.; Marth, G.T.; Sherry, S.T.; et al. 1000 Genomes Project Analysis Group. Var. Call Format VCFtools Bioinform. 2011, 27, 2156–2158. [Google Scholar]
  35. González-Peñas, J.; de Hoyos, L.; Díaz-Caneja, C.M.; Andreu-Bernabeu, Á.; Stella, C.; Gurriarán, X.; Fañanás, L.; Bobes, J.; González-Pinto, A.; Martorell, L.; et al. Recent natural selection conferred protection against schizophrenia by non-antagonistic pleiotropy. Sci. Rep. 2023, 13, 15500. [Google Scholar] [CrossRef] [PubMed]
  36. Cagliani, R.; Fumagalli, M.; Pozzoli, U.; Riva, S.; Cereda, M.; Comi, G.P.; Pattini, L.; Bresolin, N.; Sironi, M. A complex selection signature at the human AVPR1B gene. BMC Evol. Biol. 2009, 9, 123. [Google Scholar] [CrossRef] [PubMed]
  37. Lolait, S.J.; O’Carroll, A.M.; Mahan, L.C.; Felder, C.C.; Button, D.C.; Young, W.S., 3rd; Mezey, E.; Brownstein, M.J. Extrapituitary expression of the rat V1b vasopressin receptor gene. Proc. Natl. Acad. Sci. USA 1995, 92, 6783–6787. [Google Scholar] [CrossRef] [PubMed]
  38. Oshikawa, S.; Tanoue, A.; Koshimizu, T.A.; Kitagawa, Y.; Tsujimoto, G. Vasopressin stimulates insulin release from islet cells through V1b receptors: A combined pharmacological/knockout approach. Mol. Pharmacol. 2004, 65, 623–629. [Google Scholar] [CrossRef]
  39. Yibchok-Anun, S.; Cheng, H.; Heine, P.A.; Hsu, W.H. Characterization of receptors mediating AVP- and OT-induced glucagon release from the rat pancreas. Am. J. Physiol. 1999, 277, E56–E62. [Google Scholar] [CrossRef]
  40. Abate, N.; Chandalia, M. The impact of ethnicity on type 2 diabetes. J. Diabetes Complicat. 2003, 17, 39–58. [Google Scholar] [CrossRef]
  41. Velasquez-Mieyer, P.A.; Cowan, P.A.; Umpierrez, G.E.; Lustig, R.H.; Cashion, A.K.; Burghen, G.A. Racial differences in gluca-gon-like peptide-1 (GLP-1) concentrations and insulin dynamics during oral glucose tolerance test in obese subjects. Int. J. Obes. Relat. Metab. Disord. 2003, 27, 1359–1364. [Google Scholar] [CrossRef]
  42. Vaccari, C.; Lolait, S.J.; Ostrowski, N.L. Comparative distribution of vasopressin V1b and oxytocin receptor messenger ribo-nucleic acids in brain. Endocrinology 1998, 139, 5015–5033. [Google Scholar] [CrossRef]
  43. Enhörning, S.; Leosdottir, M.; Wallström, P.; Gullberg, B.; Berglund, G.; Wirfält, E.; Melander, O. Relation between human vasopressin 1a gene variance, fat intake, and diabetes. Am. J. Clin. Nutr. 2009, 89, 400–406. [Google Scholar] [CrossRef]
  44. Doliba, N.M.; Roman, J.; Rozo, A.V.; Qin, W.; Liu, C.; Naji, A.; Rickels, M.R.; Atkinson, M.A.; Powers, A.C.; Brissova, M.; et al. 253-LB: Ethnic Differences in Pancreatic Hormone Secretion in Health and T2D. Diabetes 2022, 71 (Suppl. S1), 253-LB. [Google Scholar] [CrossRef]
  45. Van West, D.; Del-Favero, J.; Deboutte, D.; Van Broeckhoven, C.; Claes, S. Associations between common arginine vasopressin 1b receptor and glucocorticoid receptor gene variants and HPA axis responses to psychosocial stress in a child psychiatric popula-tion. Psychiatry Res. 2010, 179, 64–68. [Google Scholar] [CrossRef] [PubMed]
  46. Koo, J.W.; Wohleb, E.S. How Stress Shapes Neuroimmune Function: Implications for the Neurobiology of Psychiatric Disorders. Biol. Psychiatry 2021, 90, 74–84. [Google Scholar] [CrossRef] [PubMed]
  47. McEwen, B.S.; Bowles, N.P.; Gray, J.D.; Hill, M.N.; Hunter, R.G.; Karatsoreos, I.N.; Nasca, C. Mechanisms of stress in the brain. Nat. Neurosci. 2015, 18, 1353–1363. [Google Scholar] [CrossRef] [PubMed]
  48. DeSantis, A.S.; Adam, E.K.; Doane, L.D.; Mineka, S.; Zinbarg, R.E.; Craske, M.G. Racial/Ethnic Differences in Cortisol Diurnal Rhythms in a Community Sample of Adolescents. J. Adolesc. Health 2007, 41, 3–13. [Google Scholar] [CrossRef] [PubMed]
  49. Hostinar, C.E.; McQuillan, M.T.; Mirous, H.J.; Grant, K.E.; Adam, E.K. Cortisol responses to a group public speaking task for adolescents: Variations by age, gender, and race. Psychoneuroendocrinology 2014, 50, 155–166. [Google Scholar] [CrossRef]
  50. Fieder, M.; Huber, S. The Adapting Mind in the Genomic Era. Front. Psychol. 2016, 7, 78. [Google Scholar] [CrossRef]
  51. Birnbaumer, M. Vasopressin receptors. Trends Endocrinol. Metab. 2000, 11, 406–410. [Google Scholar] [CrossRef]
  52. Scattoni, M.L.; McFarlane, H.G.; Zhodzishsky, V.; Caldwell, H.K.; Young, W.S.; Ricceri, L.; Crawley, J.N. Reduced ultrasonic vocalizations in vasopressin 1b knockout mice. Behav. Brain Res. 2008, 187, 371–378. [Google Scholar] [CrossRef]
  53. Griebel, G.; Simiand, J.; Serradeil-Le Gal, C.; Wagnon, J.; Pascal, M.; Scatton, B.; Maffrand, J.P.; Soubrie, P. Anxiolytic- and anti-depressant-like effects of the non-peptide vasopressin V1b receptor antagonist, SSR14 suggest an innovative approach for the treatment of stress-related disorders. Proc. Natl. Acad. Sci. USA 2002, 99, 6370–6375. [Google Scholar] [CrossRef]
  54. Kucharska, K.; Kulakowska, D.; Starzomska, M.; Rybakowski, F.; Biernacka, K. The improvement in neurocognitive functioning in anorexia nervosa adolescents throughout the integrative model of psychotherapy including cognitive remediation therapy. BMC Psychiatry 2019, 19, 15. [Google Scholar] [CrossRef] [PubMed]
  55. Maher, B.S.; Vladimirov, V.I.; Latendresse, S.J.; Thiselton, D.L.; McNamee, R.; Kang, M.; Bigdeli, T.B.; Chen, X.; Riley, B.P.; Hettema, J.M.; et al. The AVPR1A gene and substance use disorders: Association, replication, and functional evidence. Biol. Psychiatry 2011, 70, 519–527. [Google Scholar] [CrossRef] [PubMed]
  56. Vassileva, J.; Conrod, P.J. Impulsivities and addictions: A multidimensional integrative framework informing assessment and interventions for substance use disorders. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2019, 374, 20180137. [Google Scholar] [CrossRef] [PubMed]
  57. Füeg, F.; Santos, S.; Haslinger, C.; Stoiber, B.; Schäffer, L.; Grünblatt, E.; Zimmermann, R.; Simões-Wüst, A.P. Influence of ox-ytocin receptor single nucleotide sequence variants on contractility of human myometrium: An in vitro functional study. BMC Med. Genet. 2019, 20, 178. [Google Scholar] [CrossRef]
  58. Kim, J.; Stirling, K.J.; Cooper, M.E.; Ascoli, M.; Momany, A.M.; McDonald, E.L.; Ryckman, K.K.; Rhea, L.; Schaa, K.L.; Cosentino, V.; et al. Sequence variants in oxytocin pathway genes and preterm birth: A candidate gene association study. BMC Med. Genet. 2013, 14, 77. [Google Scholar] [CrossRef]
  59. Reinl, E.L.; Goodwin, Z.A.; Raghuraman, N.; Lee, G.Y.; Jo, E.Y.; Gezahegn, B.M.; Pillai, M.K.; Cahill, A.G.; de Guzman Strong, C.; England, S.K. Novel oxytocin receptor variants in laboring women requiring high doses of oxytocin. Am. J. Obstet. Gynecol. 2017, 217, 214e1–214e214. [Google Scholar] [CrossRef]
  60. Beck, S.; Wojdyla, D.; Say, L.; Betran, A.P.; Merialdi, M.; Requejo, J.H.; Rubens, C.; Menon, R.; Van Look, P.F. The worldwide incidence of preterm birth: A systematic review of maternal mortality and morbidity. Bull. World Health Organ. 2010, 88, 31–38. [Google Scholar] [CrossRef]
  61. Toepfer, C.N.; Wakimoto, H.; Garfinkel, A.C.; McDonough, B.; Liao, D.; Jiang, J.; Tai, A.C.; Gorham, J.M.; Lunde, I.G.; Lun, M.; et al. Hypertrophic cardiomyopathy mutations in MYBPC3 dysregulate myosin. Sci. Transl. Med. 2019, 11, eaat1199. [Google Scholar] [CrossRef]
  62. Dobewall, H.; Tark, R.; Aavik, T. Health as a value and its association with health-related quality of life, mental health, physical health, and subjective well-being. Appl. Res. Qual. Life 2018, 13, 859–872. [Google Scholar] [CrossRef]
  63. Kirwa, K.; Feric, Z.; Manjourides, J.; Alshawabekeh, A.; Vega, C.M.V.; Cordero, J.F.; Meeker, J.D.; Suh, H.H. Preterm birth and PM2.5 in Puerto Rico: Evidence from the PROTECT birth cohort. Environ. Health 2021, 20, 69. [Google Scholar] [CrossRef]
  64. Kim, J.; Park, A. A systematic review: Candidate gene and environment interaction on alcohol use and misuse among ado-lescents and young adults. Am. J. Addict. 2018, 27, 345–363. [Google Scholar] [CrossRef] [PubMed]
  65. Wermter, A.K.; Kamp-Becker, I.; Hesse, P.; Schulte-Körne, G.; Strauch, K.; Remschmidt, H. Evidence for the involvement of genetic variation in the oxytocin receptor gene (OXTR) in the etiology of autistic disorders on high-functioning level. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2010, 153B, 629–639. [Google Scholar] [CrossRef] [PubMed]
  66. Levran, O.; Peles, E.; Randesi, M.; Li, Y.; Rotrosen, J.; Ott, J.; Adelson, M.; Kreek, M.J. Stress-related genes and heroin addiction: A role for a functional FKBP5 haplotype. Psychoneuroendocrinology 2014, 45, 67–76. [Google Scholar] [CrossRef] [PubMed]
  67. Francis, S.M.; Kim, S.J.; Kistner-Griffin, E.; Guter, S.; Cook, E.H.; Jacob, S. ASD and Genetic Associations with Receptors for Oxytocin and Vasopressin-AVPR1A, AVPR1B, and OXTR. Front. Neurosci. 2016, 10, 516. [Google Scholar] [CrossRef]
  68. Harrison, A.J.; Gamsiz, E.D.; Berkowitz, I.C.; Nagpal, S.; Jerskey, B.A. Genetic variation in the oxytocin receptor gene is associated with a social phenotype in autism spectrum disorders. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2015, 168, 720–729. [Google Scholar] [CrossRef]
  69. Grotegut, C.A.; Lewis, L.L.; Manuck, T.A.; Allen, T.K.; James, A.H.; Seco, A.; Deneux-Tharaux, C. The Oxytocin Product Cor-relates with Total Oxytocin Received during Labor: A Research Methods Study. Am. J. Perinatol. 2018, 35, 78–83. [Google Scholar] [CrossRef]
  70. Schaschl, H.; Huber, S.; Schaefer, K.; Windhager, S.; Wallner, B.; Fieder, M. Signatures of positive selection in the cis-regulatory sequences of the human oxytocin receptor (OXTR) and arginine vasopressin receptor 1a (AVPR1A) genes. BMC Evol. Biol. 2015, 15, 85. [Google Scholar] [CrossRef]
  71. Ulveling, D.; Le Clerc, S.; Cobat, A.; Labib, T.; Noirel, J.; Laville, V.; Coulonges, C.; Carpentier, W.; Nalpas, B.; Heim, M.H.; et al. A new 3p25 locus is associated with liver fibrosis progression in human immunodeficiency virus/hepatitis C virus-coinfected patients. Hepatology 2016, 64, 1462–1472. [Google Scholar] [CrossRef]
  72. Kalyoncu, T.; Özbaran, B.; Köse, S.; Onay, H. Variation in the Oxytocin Receptor Gene Is Associated With Social Cognition and ADHD. J. Atten. Disord. 2019, 23, 702–711. [Google Scholar] [CrossRef]
  73. Hovey, D.; Lindstedt, M.; Zettergren, A.; Jonsson, L.; Johansson, A.; Melke, J.; Kerekes, N.; Anckarsäter, H.; Lichtenstein, P.; Lundström, S.; et al. Antisocial behavior and polymorphisms in the oxytocin receptor gene: Findings in two independent samples. Mol. Psychiatry 2016, 21, 983–988. [Google Scholar] [CrossRef]
  74. Chen, B.; Vansteenkiste, M.; Beyers, W.; Boone, L.; Deci, E.L.; Van der Kaap-Deeder, J.; Duriez, B.; Lens, W.; Matos, L.; Mouratidis, A.; et al. Basic psychological need satisfaction, need frustration, and need strength across four cultures. Motiv. Emot. 2015, 39, 216–236. [Google Scholar] [CrossRef]
  75. Johansson, A.; Bergman, H.; Corander, J.; Waldman, I.D.; Karrani, N.; Salo, B.; Jern, P.; Algars, M.; Sandnabba, K.; Santtila, P.; et al. Alcohol and aggressive behavior in men--moderating effects of oxytocin receptor gene (OXTR) polymorphisms. Genes Brain Behav. 2012, 11, 214–221. [Google Scholar] [CrossRef] [PubMed]
Table 1. Fst estimates between ACB and other global populations. Each population group is represented by 50 individual data.
Table 1. Fst estimates between ACB and other global populations. Each population group is represented by 50 individual data.
GBRSlEastFINIBSPURBEBCHB
AVP0.12560.13230.13220.12410.09530.09580.0967
AVPR1A0.16770.20720.15890.15730.18070.08760.1222
AVPR1B0.31630.16770.19290.21490.1770.19510.3253
OXT0.00000.00000.00000.01000.01180.00000.0000
OXTR0.09080.08580.10600.0760.05270.11070.1503
Table 2. Result of selection mapping in the AVPR1B gene. Abbreviations: chr—chromosome, pos—position in the given chromosome, rs—reference number of a single nucleotide variant (SNV); ACB—Afro-Caribbeans in Barbados, SlEast—East Slavs, GBR—British from England and Scotland, FIN—Finns in Finland, IBR—Iberian population in Spain, PUR—Puerto Ricans in Puerto Rico, BEB—Bengalis in Bangladesh, CHB—Han Chinese in Beijing; color codes of cells indicate the cases of significance criterion fulfillment; nSl (orange color code)—the number of segregation sites along the length, iHS (blue color code)—the integral index of haplotypes, iHH12 (yellow color code)—the integral index of homozygosity of haplotypes, Fst (green color code)—fixation index; boxes with gray fill indicate cases of exclusion from the analysis (MAF < 0.05).
Table 2. Result of selection mapping in the AVPR1B gene. Abbreviations: chr—chromosome, pos—position in the given chromosome, rs—reference number of a single nucleotide variant (SNV); ACB—Afro-Caribbeans in Barbados, SlEast—East Slavs, GBR—British from England and Scotland, FIN—Finns in Finland, IBR—Iberian population in Spain, PUR—Puerto Ricans in Puerto Rico, BEB—Bengalis in Bangladesh, CHB—Han Chinese in Beijing; color codes of cells indicate the cases of significance criterion fulfillment; nSl (orange color code)—the number of segregation sites along the length, iHS (blue color code)—the integral index of haplotypes, iHH12 (yellow color code)—the integral index of homozygosity of haplotypes, Fst (green color code)—fixation index; boxes with gray fill indicate cases of exclusion from the analysis (MAF < 0.05).
chrposrsACBSlEastGBRFINIBSPURBEBCHB
nSliHSnSliHSFstnSliHSiHH2FstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFst
chr1206107041rs28480002
chr1206107356rs28405931
chr1206108125rs28570099
chr1206108197rs28517473
chr1206108533rs28418396
chr1206109902rs33935503
chr1206110067rs33933482
chr1206110072rs33985287
chr1206110128rs28415467
chr1206110179rs28529127
chr1206110345rs28676508
chr1206110373rs28632197
chr1206110634rs28380027
chr1206110702rs28733981
chr1206110783rs28607590
chr1206110952rs28425623
chr1206111170rs28483632
chr1206111431rs33971119
chr1206111732rs28575468
chr1206112039rs28499431
chr1206112053rs34792278
chr1206112099rs34327164
chr1206112821rs33940624
chr1206113095rs28477649
chr1206113467rs28588803
chr1206114223rs3883899
chr1206115006rs28452187
Table 3. Result of selection mapping in the AVPR1A gene. Abbreviations: see legend to Table 2.
Table 3. Result of selection mapping in the AVPR1A gene. Abbreviations: see legend to Table 2.
chrposrsACBSlEastGBRFINIBSPURBEBCHB
nSliHSnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFst
chr1263144342rs10047514
chr1263144493rs11829406
chr1263144669rs11829452
chr1263144678rs10747983
chr1263144866rs10784339
chr1263149405rs10877968
chr1263149775rs1565878685
Table 4. The result of selection mapping in the AVP and OXT genes. Abbreviations: see legend to Table 2.
Table 4. The result of selection mapping in the AVP and OXT genes. Abbreviations: see legend to Table 2.
gene/chrposrsACBSlEastGBRFINIBSPURBEBCHB
nSliHSnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFst
AVP/chr203083464rs3787482
OXT/chr203071940rs111869749
Table 5. The result of selection mapping in the OXTR gene. Abbreviations: see legend to Table 2.
Table 5. The result of selection mapping in the OXTR gene. Abbreviations: see legend to Table 2.
chrposrsACBSlEastGBRFINIBSPURBEBCHB
nSliHSnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFstnSliHSFst
chr38751042rs2324728
chr38751160rs4493422
chr38751899rs237884
chr38752038rs6770632
chr38752859rs1042778
chr38755409rs237888
chr38756495rs918316
chr38760793rs57329700
chr38761165rs11131149
chr38761315rs59190448
chr38762109rs34992398
chr38762685rs53576
chr38763680rs35498753
chr38764054rs7652281
chr38764628rs11711703
chr38764738rs73132848
chr38765737rs237895
chr38766123rs61183828
chr38766375rs6767512
chr38766599rs237897
chr38766747rs34880121
chr38767536rs4686302
chr38768322rs237911
chr38768722rs4564970
chr38768834rs35413809
chr38768922rs73132856
chr38769210rs2301261
chr38769543rs2301260
Table 6. Evidence for the functional significance of SNVs of genes of the oxytocin and vasopressin system. Abbreviations: rs, reference number of a single nucleotide variant (SNV); ACB—Afro–Caribbeans in Barbados, SlEast—Eastern Slavs, GBR—British from England and Scotland, FIN—Finns in Finland, IBR—Iberian population in Spain, PUR—Puerto Ricans in Puerto Rico, BEB—Bengalis in Bangladesh, CHB—Han Chinese in Beijing; nSl is the number of segregation sites along the length, iHS is the integrated index of haplotypes, iHH12 is the integrated index of homozygosity of haplotypes, and Fst is the index of fixation.
Table 6. Evidence for the functional significance of SNVs of genes of the oxytocin and vasopressin system. Abbreviations: rs, reference number of a single nucleotide variant (SNV); ACB—Afro–Caribbeans in Barbados, SlEast—Eastern Slavs, GBR—British from England and Scotland, FIN—Finns in Finland, IBR—Iberian population in Spain, PUR—Puerto Ricans in Puerto Rico, BEB—Bengalis in Bangladesh, CHB—Han Chinese in Beijing; nSl is the number of segregation sites along the length, iHS is the integrated index of haplotypes, iHH12 is the integrated index of homozygosity of haplotypes, and Fst is the index of fixation.
SNVs of a Positive Selection (Original Data)Proved Functional Associations (Literature Data)
Gene/rsMethod/PopulationPhenotype/EndophenotypeReference
1AVPR1A rs10747983Fst/ACB & SlEast
Fst/ACB & GBR
Fst/ACB & IBS
Fst/ACB & PUR
nSl/GBR
Diabetic status, elevated blood glucose and triglycerides, body mass index (BMI)Enhörning et al., 2008 [43]
2AVPR1A rs10784339Fst (ACB& SlEast)
Fst (ACB& GBR)
Fst (ACB& IBS)
Fst (ACB& PUR)
nSl GBR
Heroin addiction, drug abuse.Levran et al., 2014 [66]
Drug abuseMaher et al., 2011 [55]
Diabetic status, elevated blood glucose and triglycerides, BMIEnhörning et al., 2008 [43]
3AVPR1B rs28632197iHS—ACB
iHH12—GBR
nSl—FIN
nSl, iHS—IBS
Autism spectrum disorders (ASD)Francis et al., 2016 [67]
Panic disordersKreek et al., 2011 [19]
Panic disordersKeck et al., 2008 [20]
4AVPR1B rs28418396nSl, iHS, iHH12—GBR
nSl, iHS—FIN
nSl, iHS—IBS
nSl, iHS—PUR
nSl—BEB, nSl—HAN
Significant side effects of therapeutic doses of vasopressin and norepinephrineAnantasit et al., 2014 [18]
5AVPR1B rs33933482nSl, iHS, iHH12—GBR
Fst ACB & GBR
nSl, iHS—FIN
nSl, iHS—IBS
nSl, iHS—PUR
Fst ACB & HAN
Psychosocial stress test (public speaking)-evoked plasma cortisol levelsvan West et al., 2010 [45]
6AVPR1B rs28676508nSl, iHS, iHH12—GBR
Fst ACB & GBR
nSl, iHS—FIN
nSl, iHS—IBS
nSl, iHS—PUR
Fst ACB & HAN
High child aggression.Zai et al., 2012 [21]
7OXTR
rs237884
Fst ACB& HAN
Fst ACB& BEB
nSl BEB
Symptom severity and treatment outcomes in subjects with schizophreniaSouza et al., 2010 [24]
ASDWermter et al., 2010 [65]
Changes in social functioning in ASDHarrison et al., 2015 [68]
8OXTR
rs6770632
Fst ACB& HAN
Fst ACB& BEB
nSl BEB
Alcohol abuse in adolescents and young adultsKim et al., 2018 [64]
Child aggressionMalik et al., 2012 [23]
Extreme, persistent, and pervasive aggressive behaviors in females and malesMalik et al., 2012 [23]
9OXTR
rs1042778
Fst ACB& HANHigh scores of depressive symptomsKeijser et al., 2021 [22]
Aggressive behavior in boysMalik et al., 2012 [23]
10OXTR
rs237888
nSl FINIncreased oxytocin sensitivity of human myometrial cells in vitroFüeg et al., 2019 [57]
The value of the effective dose of oxytocin and the outcome of childbirthGrotegut et al., 2017 [69]
11OXTR
rs59190448
Fst ACB& SlEast
Fst ACB& GBR
Fst ACB& FIN
Fst ACB& IBS
Fst ACB& BEB
Fst ACB& HAN
Positive selection Schaschl et al., 2015 [70]
12OXTR
rs61183828
iHS/ACBLiver fibrosis in patients with human immunodeficiency virus/hepatitis C virus coinfectionUlveling et al., 2016 [71]
13OXTR
rs4686302
iHS/PURIncreased oxytocin sensitivity of human myometrial cells in vitroFüeg et al., 2019 [57]
Deficit in social cognition in individuals with Attention Deficit and Hyperactivity Disorder (ADHD)Kalyoncu et al., 2019 [72]
Alcohol abuse in adolescents and young adultsKim et al., 2018 [64]
Preterm birthKim et al., 2013 [58]
The need for high doses of oxytocin in parturientsReinl et al., 2017 [59]
14OXTR
rs4564970
iHS/PURAntisocial behavior of teenage boysHovey et al., 2015 [73]
High exogenous oxytocin sensitivityChen et al., 2015 [74]
Aggressive behavior under alcohol intoxicationJohansson et al., 2012 [75]
Panic disordersKreek et al., 2011 [19]
Panic disordersKeck et al., 2008 [20]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bakoev, S.Y.; Korobeinikova, A.V.; Mishina, A.I.; Kabieva, S.S.; Mitrofanov, S.I.; Ivashechkin, A.A.; Akinshina, A.I.; Snigir, E.A.; Yudin, S.M.; Yudin, V.S.; et al. Genomic Signatures of Positive Selection in Human Populations of the OXT, OXTR, AVP, AVPR1A and AVR1B Gene Variants Related to the Regulation of Psychoemotional Response. Genes 2023, 14, 2053. https://doi.org/10.3390/genes14112053

AMA Style

Bakoev SY, Korobeinikova AV, Mishina AI, Kabieva SS, Mitrofanov SI, Ivashechkin AA, Akinshina AI, Snigir EA, Yudin SM, Yudin VS, et al. Genomic Signatures of Positive Selection in Human Populations of the OXT, OXTR, AVP, AVPR1A and AVR1B Gene Variants Related to the Regulation of Psychoemotional Response. Genes. 2023; 14(11):2053. https://doi.org/10.3390/genes14112053

Chicago/Turabian Style

Bakoev, Siroj Yu., Anna V. Korobeinikova, Arina I. Mishina, Shuanat Sh. Kabieva, Sergey I. Mitrofanov, Alexey A. Ivashechkin, Alexsandra I. Akinshina, Ekaterina A. Snigir, Sergey M. Yudin, Vladimir S. Yudin, and et al. 2023. "Genomic Signatures of Positive Selection in Human Populations of the OXT, OXTR, AVP, AVPR1A and AVR1B Gene Variants Related to the Regulation of Psychoemotional Response" Genes 14, no. 11: 2053. https://doi.org/10.3390/genes14112053

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop