Acessibilidade / Reportar erro

Partial molecular characterization, expression pattern and polymorphism analysis of MHC I genes in Chinese domestic goose (Anser cygnoides)

Abstract

Major histocompatibility complex (MHC) allelic polymorphism is critically important for mediating antigen presentation in vertebrates. Presently, there are insufficient studies of MHC genetic diversity in domestic Anseriform birds. In this study, we analyzed the expression profile of MHC I genes and screened for MHC I exon 2 polymorphism in one domestic goose population from China using Illumina MiSeq sequencing. The results showed that four MHC I alleles (Ancy-IE2*09/*11/*13/*21) in one goose were identified based on cDNA cloning and sequencing using four primer combinations, and the varying number of cDNA clones implied that these four classical sequences showed differential expression patterns. Through next-generation sequencing, 27 alleles were obtained from 68 geese with 3-10 putative alleles per individual, indicating at least the existence of 5 MHC I loci in the goose. The marked excess of the non-synonymous over the synonymous substitution in the peptide-binding region (PBR) along 27 alleles and five positively selected sites (PSSs) detected around the PBR indicated that balancing selection might be the major force in shaping high MHC variation in the goose. Additionally, IA alleles displaying lower polymorphism were subject to less positive selection pressure than non-IA alleles with a higher level of polymorphism.

Keywords:
Domestic goose; MHC I; polymorphism; balancing selection

Introduction

The major histocompatibility complex (MHC) is an extremely variable multigene family relevant to the vertebrate immune response, encoding diverse transmembrane molecules in charge of presenting pathogenic peptides to T cells (Piertney and Oliver, 2006Piertney SB and Oliver MK (2006) The evolutionary ecology of the major histocompatibility complex. Hered 96:7-21.). MHC genes fall into at least two major groups: Class I and Class II. MHC I genes expressed in most cells are essential in the recognition and presentation of intracellular peptides (Hughes and Yeager, 1998Hughes AL and Yeager M (1998) Natural selection at major histocompatibility complex loci of vertebrates. Annu Rev Genet 32:415-435.; Sommer, 2005Sommer S (2005) The importance of immune gene variability (MHC) in evolutionary ecology and conservation. Front Zool 2:16. ). MHC I genes are mainly composed of eight exons. Among them, exon 2 and 3 (abbreviated as E2 and E3) encode the α1 and α2 domains, respectively, constituting the highly polymorphic peptide-binding region (PBR) (Bjorkman and Parham, 1990Bjorkman PJ and Parham P (1990) Structure, function, and diversity of class I major histocompatibility complex molecules. Annu Rev Biochem 59:253-288. ).

As noted above, the MHC comprises the most diverse region in the vertebrate genome with a high degree of heterozygosity (Davies and Antczak, 1991Davies CJ and Antczak DF (1991) Mixed lymphocyte culture studies reveal complexity in the bovine MHC not detected by class I serology. Anim Genet 22:31-44.; Zagalska-Neubauer et al., 2010Zagalska-Neubauer M, Babik W, Stuglik M, Gustafsson L, Cichoń M and Radwan J (2010) 454 sequencing reveals extreme complexity of the class II Major Histocompatibility Complex in the collared flycatcher. BMC Evol Biol 10:395.; Sallaberry-Pincheira et al., 2016Sallaberry-Pincheira N, González-Acuña D, Padilla P, Dantas GPM, Luna-Jorquera G, Frere E, Valdés-Velásquez A and Vianna JA (2016) Contrasting patterns of selection between MHC I and II across populations of Humboldt and Magellanic penguins. Ecol Evol 6:7498-7510. ). Recently, MHC I genes have been widely studied among different vertebrate clades, such as marine organisms (Klein et al., 2007Klein J, Sato A and Nikolaidis N (2007) MHC, TSP, and the origin of species: From immunogenetics to evolutionary genetics. Annu Rev Genet 41:281-304. ; Grimholt et al., 2015Grimholt U, Olsaker I, Lingaas F and Lie (2015) Sequence analysis of MHC class Ⅰ α2 domain exon variants in one diploid and two haploid atlantic salmon pedigrees. Anim Genet 28:385-390. ), mammals (Westerdahl et al., 1999Westerdahl H, von Schantz T and Wittzell H (1999) Polymorphism and transcription of Mhc class I genes in a passerine bird, the great reed warbler. Immunogenetics 49:158-170. ; Bartocillo et al., 2021Bartocillo F, Nishita Y, Abramov V and Masuda R (2021) Evolution of MHC class I genes in Japanese and Russian raccoon dogs, Nyctereutes procyonoides ( Carnivora : Canidae ). Mamm Res 66:371-383.), and birds (Biedrzycka et al., 2017Biedrzycka A, O’Connor E, Sebastian A, Migalska M, Radwan J, Zając T, Bielański W, Solarz W, Ćmiel A, and Westerdahl H (2017) Extreme MHC class i diversity in the sedge warbler (Acrocephalus schoenobaenus); Selection patterns and allelic divergence suggest that different genes have different functions. BMC Evol Biol 17:159.; Manjula et al., 2021Manjula P, Fulton JE, Seo D and Lee JH (2021) Comparison of major histocompatibility complex-B variability in Sri Lankan indigenous chickens with five global chicken populations using MHC-B SNP panel. Anim Genet 52:824-833.), because of their genetic variability (Bateson et al., 2015Bateson ZW, Whittingham LA, Johnson JA and Dunn PO (2015) Contrasting patterns of selection and drift between two categories of immune genes in prairie-chickens. Mol Ecol 24:6095-6106.). MHC polymorphism enables populations and individuals to produce corresponding immune responses to various pathogens and cope with changing environmental conditions (Sommer, 2005Sommer S (2005) The importance of immune gene variability (MHC) in evolutionary ecology and conservation. Front Zool 2:16. ). The genetic diversity among MHC genes is affected by parasite-related balancing selection (Potts and Slev, 1995Potts WK and Slev PR (1995) Pathogen‐based models favoring MHC genetic diversity. Immunol Rev 143:181-197. ; Bateson et al., 2015Bateson ZW, Whittingham LA, Johnson JA and Dunn PO (2015) Contrasting patterns of selection and drift between two categories of immune genes in prairie-chickens. Mol Ecol 24:6095-6106.). This balancing selection maintains genetic variation in populations based on two possible mechanisms: heterozygote advantage (Hughes and Yeager, 1998Hughes AL and Yeager M (1998) Natural selection at major histocompatibility complex loci of vertebrates. Annu Rev Genet 32:415-435.) and frequency-dependent selection (Spurgin and Richardson, 2010Spurgin LG and Richardson DS (2010) How pathogens drive genetic diversity: MHC, mechanisms and misunderstandings. Proc R Soc B: Biol Sci 277:979-988. ). Since evolving pathogens can easily escape presentation by common host alleles, rare alleles have higher fitness to survive in the host population under the prevalence of frequency-dependent selection. Hence, spatial-temporal dynamics in the pathogen community induce population variation at MHC. In the case of overdominance, individuals heterozygous in a particular gene can resist the pathogen infection better than homozygotes. In addition, sexual selection also acts to promote MHC variation (Knafler et al., 2012Knafler GJ, Clark JA, Boersma PD and Bouzat JL (2012) MHC diversity and mate choice in the magellanic penguin, Spheniscus magellanicus. J Hered 103:759-768. ).

Balancing selection was considered to be the main force for the generation and maintenance of MHC diversity (Hedrick, 1998Hedrick PW (1998) Balancing selection and MHC. Genet 104:207-214. ). Nucleotide positions under positive selective pressure were expected to produce excess non-synonymous substitutions, inducing amino acid changes and structural alterations in MHC molecules (Potts and Slev, 1995Potts WK and Slev PR (1995) Pathogen‐based models favoring MHC genetic diversity. Immunol Rev 143:181-197. ; Piertney and Oliver, 2006Piertney SB and Oliver MK (2006) The evolutionary ecology of the major histocompatibility complex. Hered 96:7-21.). Such parasite-medicated selection should be notable at the PBR (Hughes and Nei, 1989Hughes AL and Nei M (1989) Nucleotide substitution at major histocompatibility complex class II loci: Evidence for overdominant selection. Proc Natl Acad Sci U S A 86:958-962.; Bjorkman and Parham, 1990Bjorkman PJ and Parham P (1990) Structure, function, and diversity of class I major histocompatibility complex molecules. Annu Rev Biochem 59:253-288. ).

In birds, the number of MHC loci and MHC alleles variation exhibit enormous differences among different taxa. In passerines, they have a large number of MHC genes, and each locus produces a large number of alleles (Balasubramaniam et al., 2016Balasubramaniam S, Bray RD, Mulder RA, Sunnucks P, Pavlova A and Melville J (2016) New data from basal Australian songbird lineages show that complex structure of MHC class II β genes has early evolutionary origins within passerines. BMC Evol Biol 16:112.; Pardal et al., 2017Pardal S, Drews A, Alves JA, Ramos JA and Westerdahl H (2017) Characterization of MHC class I in a long distance migratory wader, the Icelandic black-tailed godwit. Immunogenetics 69:463-478. ). For example, a single sedge warbler (Acrocephalus schoenobaenus) could carry up to 65 different alleles, indicating at least 33 MHC I loci (O’Connor et al., 2016O’Connor EA, Strandh M, Hasselquist D, Nilsson J and Westerdahl H (2016) The evolution of highly variable immunity genes across a passerine bird radiation. Mol Ecol 25:977-989. ; Biedrzycka et al., 2017Biedrzycka A, O’Connor E, Sebastian A, Migalska M, Radwan J, Zając T, Bielański W, Solarz W, Ćmiel A, and Westerdahl H (2017) Extreme MHC class i diversity in the sedge warbler (Acrocephalus schoenobaenus); Selection patterns and allelic divergence suggest that different genes have different functions. BMC Evol Biol 17:159.). Conversely, Galliformes birds, such as chicken and turkey, has a limited number of MHC loci (Kaufman et al., 1999Kaufman J, Milne S, Göbel TWF, Walker BA, Jacob JP, Auffray C, Zoorob R and Beck S (1999) The chicken B locus is a minimal essential major histocompatibility complex. Nature 401:923-925. ; Chaves et al., 2009Chaves LD, Krueth SB and Reed KM (2009) Defining the Turkey MHC: Sequence and genes of the B Locus. J Immunol 183:6530-6537.). The chicken MHC B, called “minimal essential MHC”, has only two classical class I genes (BF1 and BF2). As representatives of domestic Anseriformes bird, duck and goose, few reports on MHC I polymorphism are available (Mesa et al., 2004Mesa CM, Thulien KJ, Moon DA, Veniamin SM and Magor KE (2004) The dominant MHC class I gene is adjacent to the polymorphic TAP2 gene in the duck, Anas platyrhynchos. Immunogenetics 56:192-203. ; Xia et al., 2004Xia C, Lin CY, Xu GX, Hu TJ and Yang TY (2004) cDNA cloning and genomic structure of the duck (Anas platyrhynchos) MHC class I gene. Immunogenetics 56:304-309. ; Moon et al., 2005Moon DA, Veniamin SM, Parks-Dely JA and Magor KE (2005) The MHC of the Duck ( Anas platyrhynchos ) contains five differentially expressed class I genes . J Immunol 175:6702-6712.). Estimating population genetic variation in MHC genes is hindered by extreme MHC diversity, and the acquisition of high-coverage sequence profiles through the amplicon-throughput sequencing make it convenient and efficient to assess MHC diversity (Babik et al., 2009Babik W, Taberlet P, Ejsmond MJ and Radwan J(2009) New generation sequencers as a tool for genotyping of highly polymorphic multilocus MHC system. Mol Ecol Resour 9:713-719.; Grogan et al., 2016Grogan KE, Mcginnis GJ, Sauther ML, Cuozzo FP and Drea CM (2016) Next-generation genotyping of hypervariable loci in many individuals of a non-model species: Technical and theoretical implications. BMC Genom 17:204. ).

The Yangzhou white goose (Anser cygnoides, Anatidae, Anseriformes) originated from Jiangsu Province in China, is one of the most important waterfowl breeds. Infectious diseases of domestic goose caused by pathogens, such as avian influenza viruses have broken out frequently, which severely restricts the development of the goose industry. MHC I alleles are associated with the resistance and susceptibility of pathogens (Chenani et al., 2021Chenani H, Nazari M, Beigi Nassiri MT, Roshanfekr H and Aghaie A (2021) Exonic SNP in MHC-DMB2 is associated with gene expression and humoral immunity in Japanese quails. Vet Immunol Immunop 239:110302.). Previous work published six goose MHC I full cDNA sequences and analyzed its genomic structure (Xia et al., 2005Xia C, Hu T, Yang T, Wang L, Xu G and Lin C (2005) cDNA cloning, genomic structure and expression analysis of the goose (Anser cygnoides) MHC class I gene. Vet Immunol Immunop 107:291-302.). Based on these findings, we further implement the expression profiles and the allelic polymorphism of MHC I E2 of one domestic goose population. The major goals in this study are to (i) characterize the genetic polymorphism of MHC I E2 alleles through Illumina MiSeq, (ii) determine the role of positive selection and recombination in shaping class I variation, and (iii) infer the evolutionary relationship of Anseriformes.

Material and Methods

Sample preparation

Sixty-eight blood samples of adult Yangzhou white geese (hereinafter referred to as YW-goose or goose) were collected by brachial venipuncture for DNA extraction. The animal studies were performed under the approval of the Animal Ethics Committee of Qilu University of Technology (Shandong Academy of Sciences). Genomic DNA was extracted by a standard phenol-chloroform protocol (Sambrook et al., 1990Sambrook J, Fritsch EF and Maniatis T (1990) Molecular cloning: A laboratory manual. Cold Spring Harbor Laboratories 186:182-183.). Additionally, the blood and three kinds of fresh organs (heart, lung, and intestine) from one YW-goose were preserved in liquid nitrogen as an mRNA resource. Total RNA was extracted by Trizol reagent (Vazyme Biotech Co. Ltd). The first strand cDNA was synthesized by reverse transcription according to the instruction of the HiScriptⅢRT SuperMix for qPCR kit (Vazyme Biotech Co. Ltd). cDNA samples were stored at -80 °C.

Primer design, molecular cloning and Sanger sequencing

The first set of new primers (E2-4IF/E2-4IR; Table S1 Table S1 - Primers used to amplify MHC I sequences in the domestic goose ) amplified approximately 1697 bp spanning from E2 to partial E4, on the basis of an alignment comprising MHC I sequences from domestic goose (AY387652; AY387658; AY387651; AY387650; AY387648; AY387699), domestic chicken (M31012), bar-headed goose (FJ606105; FJ606106; FJ606107; FJ606108; FJ606109) and duck (AB115244; AY294419). Then, three other primer pairs (E2AF/E2AR, E2IF/E2IR, E3I20F/E3I30R; Table S1 Table S1 - Primers used to amplify MHC I sequences in the domestic goose ) were designed to amplify E2 and E3, respectively. Both gDNA and cDNA samples were used as PCR templates. The 25 µL PCR reaction system contains the following components:5 µl of PCR enhancer, 12.5 µL of 2×phanta max master Mix, 20 ng of blood DNA template, 2 µM of each primer , add water to a total volume of 25 µL. The PCR program consisted of one cycle at 95 °C for 5 min, followed by 36 cycles of amplification at 95 °C for 15 s, Tm for 15 s and 72 °C for 0.5-1.5 min, and a final amplification at 72 °C for 5 min. The PCR products of the correct length were cloned and then the randomly-selected positive clones were sequenced using M13 primers. The combined sequences from different primer pairs gave satisfactory coverage of the hypervariable region E2 and E3, which encoded the peptide-binding region (Pardal et al., 2017Pardal S, Drews A, Alves JA, Ramos JA and Westerdahl H (2017) Characterization of MHC class I in a long distance migratory wader, the Icelandic black-tailed godwit. Immunogenetics 69:463-478. ).

Illumina MiSeq sequencing and data analysis

According to the above sequence alignments obtained through Sanger sequencing, the universal primer pair E2AF/E2AR (Table S1 Table S1 - Primers used to amplify MHC I sequences in the domestic goose ) was newly devised for the MiSeq sequencing, which amplified a 176bp fragment (excluding the primer sequences) of MHC I E2. After being purified by the KAPA Pure Beads (Roche), each PCR product labelled with a unique barcode was sequenced using a 2×250 bp paired-end sequencing strategy on the Illumina NovaSeq 6000 system. Raw FASTQ reads were demultiplexed using the barcode sequence with the exact barcode matching parameter (https://github.com/jameslz/div-utils/blob/master/div-utils, version: 0.0.1-r1-dirty). Quality-filtering utilized by the Trimmomatic version was performed as in Bolger et al. (2014Bolger AM, Lohse M and Usadel B (2014) Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30:2114-2120.). Paired reads were merged using the USEARCH command (http://drive5.com, version 11.2.64) (Edgar and Flyvbjerg, 2015Edgar RC and Flyvbjerg H (2015) Error filtering, pair assembly and error correction for next-generation sequencing reads. Bioinformatics 31:3476-3482.) with the default parameters. Reads that cannot be merged were discarded; the merged reads with more than two nucleotide mismatches in primer matching and the primer sequences from the merged reads were deleted.

Validation and identification of MHC I alleles

All alleles from cloning and Illumina sequencing were verified in BLASTN (NCBI; National Council for the Blind of Ireland) to confirm whether they were MHC I alleles. We performed repeated PCR verification for most of the suspicious alleles to exclude PCR artifacts. When a clone was found in more than two separate PCRs or an allele obtained by Illumina sequencing was found in more than two samples, it was considered an MHC I allele that met the criteria summarized by Kennedy et al. (2002Kennedy L, Ryvar R, Gaskell R, Addie D, Willoughby K, Carter S, Thomson W, Ollier W and Radford A (2002) Sequence analysis of MHC DRB alleles in domestic cats from the United Kingdom. Immunogenetics 54:348-352. ). According to the nomenclature of Klein et al. (1990Klein J, Bontrop RE, Dawkins RL, Erlich HA, Gyllensten UB, Heise ER, Jones PP, Parham P, Wakeland EK and Watkins DI (1990) Nomenclature for the major histocompatibility complexes of different species: A proposal. Immunogenetics 31:217-219.), we simply named our alleles Ancy-IE2*01-27. Unique alleles were uploaded to the GenBank database and specific codes were obtained (GenBank accession number: OK289674-OK289700). Specifically, the 10 E2 alleles (Ancy-IE2*01-10) shared the conserved sequence region, hence the specific primers (E2IF/E2QIR) were designed to determine whether there truly existed such an MHC I locus (hereinafter called as IA.I stands for MHC class I, and A refers to the locus ordinal number) in the goose.

Analysis of allelic diversity, positive selection and recombination

We analyzed MHC I E2 sequences in MEGA X (Kumar et al., 2018Kumar S, Stecher G, Li M, Knyaz C and Tamura K (2018) MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547-1549. ). The diversity analysis included the average nucleotide distance (d nt ), average amino acid distance (d aa ), and average nucleotide diversity (π). The d nt and d aa adopted the Kimura 2-parameter model (Kimura, 1980Kimura M (1980) A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J Mol Evol 16:111-120. ) and P-distance model, respectively, and both tests were run with 1000 bootstrap repeats.

For the selection effect of MHC I gene, firstly, we used the modified Nei-Gojobori method (Nei and Gojobori, 1986Nei M and Gojobori T (1986) Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions. Mol Biol Evol 3:418-426. ) in MEGA X to calculate the average rate of non-synonymous (d N ) and synonymous (d S ) substitution for all codons, the PBR and non-PBR region. Fourteen PBR codons (1V, 10V, 19M, 38D, 41S, 42S, 44S, 45N, 48I, 49Y, 51V, 52N, 55T, 56L) were verified among goose MHC I E2 alleles based on human and chicken MHC I molecules (Kaufman et al., 1992Kaufman J, Andersen R, Avila D, Engberg J, Lambris J, Salomonsen J, Welinder and Skjødt K (1992) Different features of the MHC class I heterodimer have evolved at different rates. Chicken BF and beta 2-microglobulin sequences reveal invariant surface residues. J Immunol 148:1532-1546. ; Wallny et al., 2006Wallny HJ, Avila D, Hunt LG, Powell TJ, Riegert P, Salomonsen J, Skjødt K, Vainio O, Vilbois F, Wiles MV et al. (2006) Peptide motifs of the single dominantly expressed class I molecule explain the striking MHC-determined response to Rous sarcoma virus in chickens. Proc Natl Acad Sci U S A 103:1434-1439. ). Secondly, we applied two software to detect the specific positive selection sites (PSSs). The maximum likelihood (ML) as a random site mode, could better identify the PSSs since it described the overall change among sites (Furlong and Yang 2008Furlong RF and Yang Z (2008) Diversifying and purifying selection in the peptide binding region of DRB in mammals. J Mol Evol 66:384-394.). The nested site models [M7 (0 <ω<1) and M8 (ω>1)] in PAML (Yang, 2007Yang Z (2007) PAML 4: Phylogenetic analysis by maximum likelihood. Mol Biol Evol 24:1586-1591. ), which assumed that all sites presented a beta distribution, were used for the comparison to determine PSSs. Two models, FEL (Fixed effects likelihood) and MEME (Mixed effects model of evolution), were performed through the Data-monkey web server (http://www.datamonkey.org/).

To test the recombination events occurring at MHC I genes of the goose, the genetic algorithm of the RDP (GARD) was also performed in the Data-monkey web server.

Phylogenetic relationship

In order to evaluate the phylogenetic relationship of MHC I genes among the Anseriform species, all the MHC I E2 alleles of goose identified in this article and the corresponding sequences of six other Anseriform species from GenBank were used for phylogenetic reconstruction. The human MHC I sequence (K02883) was used as an outgroup. Phylogenetic trees were constructed by the neighbor-joining (NJ) method in MEGA X with 1000 rapid bootstrap replicates. Nucleotide-based trees was built using Tamura-Kumar models (Tamura, 1992Tamura K (1992) Estimation of the number of nucleotide substitutions when there are strong transition-transversion and G+C-content biases. Mol Biol Evol 9:678-687. ). The network was built by Splits Tree v.4.14.4 based on the substitution model the Kimura 2-parameter (K2P) with 1000 bootstrap repetitions.

Results

cDNA analysis of goose MHC I genes

Sanger sequencing using four pairs of primers from both gDNA and cDNA of one single YW-goose confirmed MHC I sequences of different lengths (201bp-1697bp; Table S1 Table S1 - Primers used to amplify MHC I sequences in the domestic goose ), identifying a total of four different alleles (Ancy-IE2*09, Ancy-IE2*11, Ancy-IE2*13, and Ancy-IE2*21; Figure 1). 109 cDNA clones were obtained from the same individual. Among them, Ancy-IE2*13 and Ancy-IE2*21 were in the vast majority, accounting for 44% and 39.4% respectively, while Ancy-IE2*09 and Ancy-IE2*11 only accounted for 7.3% and 9.2%, respectively (Table S2 Table S2 - Numbers of cDNA clones amplified by four pairs of primers ). These four long transcripts covered the major part of E2-4: α1 (90 amino acids), α2 (90 amino acids), and partial α3 (70 out of 92 amino acids) domain. In non-mammalian vertebrates, the classical MHC I amino acid sequences comprised eight highly conserved sites “YYRTKWYY” (Kaufman et al., 1994Kaufman J, Salomonsen J and Flajnik M (1994) Evolutionary conservation of MHC class I and class II molecules--different yet the same. Semin Immunol 6:411-424. ; Shum et al., 1999Shum BP, Rajalingam R, Magor KE, Azumi K, Carr WH, Dixon B, Stet RJM, Adkison MA, Hedrick RP and Parham P (1999) A divergent non-classical class I gene conserved in salmonids. Immunogenetics 49:479-490. ), which were termed as peptide main-chain sites, and similarly, these eight sites were all embodied in the MHC I transcripts of goose and four other Anseriform species (Cygnus atratus; Cygnus olor; Anas platyrhynchos; Aythya fuligula). Besides, classical MHC molecules were extremely conservative in inter- and intra- domain contact residues (Shum et al., 1999). In the Anseriformes transcripts, 10 out of the 18 inter- and intra- domain contact residues were included: 9 positions remained unchanged, and only one position was highly variable in the alignment (Figure 1). cDNA sequence and expression analysis collectively illustrated that at least four classical alleles with distinct expression patterns existed in this single YW-goose individual.

Figure 1 -
Amino acid alignment of MHC I long transcripts in the domestic goose. The same sites as the Ancy-IE2*09 alleles were indicated by “·”; the transparent boxes represented the inter-domain and intra-domain contact residues and the light gray boxes represented the peptide main chain sites; c represented non-main chain peptide contacts, while a and f represented A and F pocket, respectively. The Anseriform sequences from NCBI included Cygnus atratus (XR004782210.1); Anas platyrhynchos (GU245865.1); Cygnus olor (XM040542813.1); Anas platyrhynchos (KX118687.1); Aythya fuligula (XM032205162.1).

Exon 2 sequences from Illumina MiSeq

383 MHC I sequences (176 bp) in 68 individuals were detected using next-generation sequencing, and eventually, a total of 27 MHC I alleles were determined after sequence alignment (Table S3 Table S3 - Statistics of MHC I exon 2 alleles from 68 domestic geese ). 27 MHC I alleles from Illumina MiSeq were translated into 21 unique amino acid sequences, which covered most of the PBR of the α1 domain (14/18). The number of alleles for each individual varied greatly from 3 to 10. Thus, it was speculated that there were at least 5 MHC I loci in the goose. Six alleles were present in more than half of the individuals (Ancy-IE2*01(50/68), Ancy-IE2*09(45/68), Ancy-IE2*11(47/68), Ancy-IE2*13(48/68), Ancy-IE2*014(37/68), Ancy-IE2*16 (56/68)), while fifteen alleles only appeared in 2-7 individuals (Table S3 Table S3 - Statistics of MHC I exon 2 alleles from 68 domestic geese ).

Among the 176 nucleotide positions of the 27 MHC I alleles, 54% were variable. There was no premature stop codon in the putative amino acid sequences, implying that these alleles might be functional. The average nucleotide diversity (π), nucleotide distance (dnt) and amino acid distance (daa) of the PBR region in goose MHC I E2 were significantly higher than the non-PBR region (Table 1).

Table 1 -
Measures of average nucleotide diversity (π) and nucleotide distance (dnt) and amino acid distance (daa) for goose MHC I exon 2 alleles.

Positive selection and recombination of goose MHC I exon 2 alleles

Calculations of d N (non-synonymous substitution) and d S (synonymous substitution) were listed in Table 2. All MHC E2 alleles yielded higher d N than d S values in all codons, PBR and non-PBR; moreover, the d N /ds of PBR was higher than that of non-PBR (2.08, 1.657, respectively), although its d N and ds values were slightly less than that of non-PBR codons.

Table 2 -
Calculations of non-synonymous (dN) and synonymous (dS) substitutions for the goose MHC I exon 2 alleles.

Five PSSs within non-IA alleles were detected using three methods based on the detection of PSSs for 58 codons of E2, while only one PSS within IA alleles (Table S4 Table S4 - Inference of positively selected amino acid sites for domestic goose MHC I sequences ). Five PSSs were identified through the likelihood ratio test of comparing the nested models M7 and M8. Two PSSs were detected in FEL and MEME. However, there were few sites within E2 under positive selection using four methods, all the PSSs located in or around PBR, which were directly related to peptide binding.

Recombination studies were also carried out using GARD among 27 alleles; the results showed that no recombination point was found, mainly because of the short sequence length.

Locus determination and polymorphism analysis of IA

By aligning 27 amino acid sequences, it was impossible to assign all the sequences to separate loci by visual inspection. However, it was found that the specific amino acid motif composed of multiple conserved amino acid residues in an α helix and its vicinity was shared by ten alleles (Ancy-IE2*01-10, Figure 2), resembling that of chicken BF1 (Livant et al., 2004Livant EJ, Brigati JR and Ewald SJ (2004) Diversity and locus specificity of chicken MHC B class I sequences. Anim Genet 35:18-27. ), golden pheasant IA1 (Zeng et al., 2016Zeng QQ, He K, Sun DD, Ma MY, Ge YF, Fang SG and Wan QH (2016) Balancing selection and recombination as evolutionary forces caused population genetic variations in golden pheasant MHC class I genes. BMC Evol Biol 16:42. ), and human HLA-C (Zemmour and Parham, 1992Zemmour J and Parham P (1992) Distinctive polymorphism at the HLA-C locus: Implications for the expression of HLA-C. J Exp Med 176:937-950. ) alleles, which could be used to distinguish individual loci. Based on this remarkable sequence feature, we preliminarily determined that these ten sequences belonged to one gene locus-locus IA. For the IA locus, no more than two alleles were identified per YW-goose by sequence-specific PCR and Sanger sequencing, further elucidating that such a single locus truly existed.

Figure 2 -
Amino acid alignment of 27 MHC I exon 2 alleles of domestic goose. “+” indicated the predicted PBR site. “★” indicated the positively selected sites (PSSs) predicted by M7 and M8 site model, while “○” and “▲” were the PSSs detected by FEL and MEME models, respectively. The locus-specific residues of IA alleles were indicated by open boxes.

The polymorphisms of ten alleles of predicted locus IA were analyzed (Table 1), which had few variant sites, and the amino acid diversity, nucleotide distance and amino acid distance were extremely low, indicating that the locus IA had relatively low genetic diversity. Since they were highly conserved around the PBR, the π, d nt and d aa values of the PBR region were lower than other codons. Nevertheless, a higher d N than d S value was observed in all codons and PBR (Table 2), suggesting locus IA might be subjected to positive selection.

Phylogenetic analysis

The goose MHC I sequences clustered with the previously reported MHC I sequences from other Anseriformes species, such as Anser anser, Anser indicus, Cygnus olor, Anas platyrhynchos, Cygnus atratus, Aythya fuligula and Anas laysanensis. At the same time, they diverged from those of non- Anseriformes birds both in the nucleic acid (Figure 3) phylogenetic trees. Within the Anseriformes clade, sequences from different species were inclined to gather together. For the IA locus alleles of geese, all alleles formed an independent cluster with a supporting rate of more than 75%, indicative of the independent evolution of this locus; for non-IA alleles, mostly fell into two major clusters due to sequence divergence while some mingled with MHC I sequences from other Anseriform species (Ancy-IE2*11/21/23/27). Besides, the network relationship among MHC I sequences of Anseriform species resembled the situation described above but was not shown here.

Figure 3 -
Phylogenetic tree of nucleic acid (a) and amino acid (b) sequences for domestic goose MHC I exon 2. The sequences used to generate the trees were as follows: Chrysolophus pictus (KM005713, KM005703); Larus scopulinus (HM025966, HM025979); Grus grus (MK034099); Grus carunculatus (MK034103); Gallus (KF032308, AY234770); Anser anser (AM114924, AY387649, AY387648); Anser indicus (FJ606105, FJ606106, FJ606107); Cygnus olor (XM040542809); Anas platyrhynchos (AB115240, AB115244, AB115246); Cygnus atratus (XR004782210, XM035570471); Aythya fuligula (XM032205203); Anas laysanensis (KF612482); Homo sapiens (KO2883). Only bootstrap values >50% were shown at the node. The red triangle represented non-IA alleles of domestic geese, the green triangle represented IA alleles of domestic geese, and the yellow triangle represented class I sequences from other Anseriform species. Homo sapiens (KO2883) as the outgroup was shown in blue and sequences from non-Anseriform bird were shown in red.

Discussion

The determination of goose classical MHC I sequences

In this study, we isolated four MHC I alleles from the cDNA of a single YW-goose, spanning the region from E2 to partial E4. Four alleles that could express were detected by different primer combinations and displayed sequence characteristics of the classical MHC I genes, i.e., conserved peptide main-chain sites and inter- and intra- domain contact residues (Kaufman et al., 1994Kaufman J, Salomonsen J and Flajnik M (1994) Evolutionary conservation of MHC class I and class II molecules--different yet the same. Semin Immunol 6:411-424. ). Previous studies showed that the relative cDNA clone number of each sequence was a good indicator of the expression levels of these genes (Mesa et al., 2004Mesa CM, Thulien KJ, Moon DA, Veniamin SM and Magor KE (2004) The dominant MHC class I gene is adjacent to the polymorphic TAP2 gene in the duck, Anas platyrhynchos. Immunogenetics 56:192-203. ). Hence, through the number of cDNA clones of the four alleles, we speculated that the expression levels of the four alleles might be different: two highly expressed alleles (Ancy-IE2*13 and Ancy-IE2*21) and two relatively weakly expressed alleles (Ancy-IE2*09 and Ancy-IE2*11). The differential expression pattern between different MHC classical loci implied differences in their capacity for defense against viruses, which was commonly found in other species (Zemmour and Parham, 1992Zemmour J and Parham P (1992) Distinctive polymorphism at the HLA-C locus: Implications for the expression of HLA-C. J Exp Med 176:937-950. ; Biedrzycka et al., 2017Biedrzycka A, O’Connor E, Sebastian A, Migalska M, Radwan J, Zając T, Bielański W, Solarz W, Ćmiel A, and Westerdahl H (2017) Extreme MHC class i diversity in the sedge warbler (Acrocephalus schoenobaenus); Selection patterns and allelic divergence suggest that different genes have different functions. BMC Evol Biol 17:159.): BF2 was dominantly expressed compared with BF1 in the chicken (Kaufman, 2020Kaufman J (2020) From Chickens to Humans: The importance of peptide repertoires for MHC Class I Alleles. Front Immunol 11:601089.).

Extensive MHC allelic variation in the YW-goose population

It was the first time we characterized the E2 of MHC I genes in one population of YW-geese using the Illumina MiSeq. We founded that each individual contained 3-10 alleles, indicating that there were at least 5 class I loci in the goose. Similarly, the MHC of the Duck, as a closely related species of goose, reportedly contained 5 class I loci (UAA, UBA, UCA, UDA, and UEA) (Moon et al., 2005Moon DA, Veniamin SM, Parks-Dely JA and Magor KE (2005) The MHC of the Duck ( Anas platyrhynchos ) contains five differentially expressed class I genes . J Immunol 175:6702-6712.). In total, 27 distinct MHC I E2 alleles were identified from 68 individuals, revealing a high genetic variation of ɑ1 domain in the goose class I genes. Other avian species where they also reported to exhibit the extensive polymorphism at the peptide-binding domains of MHC I genes (Pardal et al., 2017Pardal S, Drews A, Alves JA, Ramos JA and Westerdahl H (2017) Characterization of MHC class I in a long distance migratory wader, the Icelandic black-tailed godwit. Immunogenetics 69:463-478. ). Especially Passerine species showed extreme polymorphism at MHC genes, e.g., 88 MHC I alleles obtained from 18 siskins (Spinus spinus) (Drews and Westerdahl, 2019Drews A and Westerdahl H (2019) Not all birds have a single dominantly expressed MHC-I gene: Transcription suggests that siskins have many highly expressed MHC-I genes. Sci Rep 9:19506. ). The abundant sequence variation of YW-goose MHC I loci might enable this species to recognize and present a large number of pathogenic antigens (Potts and Slev, 1995Potts WK and Slev PR (1995) Pathogen‐based models favoring MHC genetic diversity. Immunol Rev 143:181-197. ).

According to the allele distribution of 68 individuals, Ancy-IE2*01 (50/68) and Ancy-IE2*09 (56/68) were the most common alleles, indicating that these two alleles might be favoured by selection in this YW-goose population (Oliver et al., 2009Oliver MK, Lambin X, Cornulier T and Piertney SB (2009) Spatio-temporal variation in the strength and mode of selection acting on major histocompatibility complex diversity in water vole (Arvicola terrestris) metapopulations. Mol Ecol 18:80-92. ). There were also 14 alleles only appeared in 2-3 individuals, suggesting many rare alleles had evolved in this population under frequency-dependent selection, the fundamental process contributing to the generation and maintenance of MHC variability (Hughes and Nei, 1989Hughes AL and Nei M (1989) Nucleotide substitution at major histocompatibility complex class II loci: Evidence for overdominant selection. Proc Natl Acad Sci U S A 86:958-962.; Takahata and Nei, 1990Takahata N and Nei M (1990) Allelic genealogy under overdominant and frequency-dependent selection and polymorphism of major histocompatibility complex loci. Genetics 124:967-978. ). Furthermore, five alleles (Ancy-IE2*01, Ancy-IE2*09, Ancy-IE2*11, Ancy-IE2*13, Ancy-IE2*19) were also detected in the geese from Beijing (Xia et al., 2005Xia C, Hu T, Yang T, Wang L, Xu G and Lin C (2005) cDNA cloning, genomic structure and expression analysis of the goose (Anser cygnoides) MHC class I gene. Vet Immunol Immunop 107:291-302.), indicating that these alleles were shared among different areas. This situation implied that two different populations of domestic geese experienced very similar pathogen-mediated selective regimes, which was commonly reported in birds, such as owls (Burri et al., 2008Burri R, Hirzel HN, Salamin N, Roulin A and Fumagalli L (2008) Evolutionary patterns of MHC class II B in owls and their implications for the understanding of avian MHC evolution. Mol Biol Evol 25:1180-1191.), penguins (Sallaberry-Pincheira et al., 2016Sallaberry-Pincheira N, González-Acuña D, Padilla P, Dantas GPM, Luna-Jorquera G, Frere E, Valdés-Velásquez A and Vianna JA (2016) Contrasting patterns of selection between MHC I and II across populations of Humboldt and Magellanic penguins. Ecol Evol 6:7498-7510. ), and passerines (Eimes et al., 2016Eimes JA, Lee SI, Townsend AK, Jablonski P, Nishiumi I and Satta Y (2016) Early duplication of a single MHC IIB locus prior to the passerine radiations. PLoS One 11:e0163456. ).The discovery of these alleles laid the foundation for the breeding and disease resistance research of goose.

Locus identification of IA

Previous studies verified that E2 sequences from domestic chicken BF1 (Piertney and Oliver, 2006Piertney SB and Oliver MK (2006) The evolutionary ecology of the major histocompatibility complex. Hered 96:7-21.), golden pheasant IA1 (Zeng et al., 2016Zeng QQ, He K, Sun DD, Ma MY, Ge YF, Fang SG and Wan QH (2016) Balancing selection and recombination as evolutionary forces caused population genetic variations in golden pheasant MHC class I genes. BMC Evol Biol 16:42. ), and human HLA-C (Zemmour and Parham, 1992Zemmour J and Parham P (1992) Distinctive polymorphism at the HLA-C locus: Implications for the expression of HLA-C. J Exp Med 176:937-950. ) shared a locus-specific amino acid motif, and such a phenomenon also appeared among ten MHC I alleles of the goose (Ancy-IE2*01-10). Based on this striking finding, these ten alleles most likely have classified into the IA locus of the goose. Besides, the locus-specific PCR amplification results proved the existence of the IA locus. Undoubtedly, the determination of locus IA in geese would contribute to further elucidation of the evolutionary processes underlying MHC variation at such a particular locus.

Differential polymorphism caused by different selection pressure

Based on the defined criteria, d N /d S >1 was usually considered as evidence of positive selection for MHC diversification. For all the 27 E2 alleles of goose MHC I, positive selection exerted an influence on not only PBR codons but also non-PBR codons, which both exhibited an obvious excess of non-synonymous over synonymous substitutions. Furthermore, E2 alleles showed a higher d N /d S ratio at PBR codons than that at non-PBR codons, which conformed to the hallmark of functionally important residues under balancing selection. However, compared with 17 non-IA alleles, ten alleles of the IA locus exhibited less allele polymorphisms and fewer frequent nucleotide mutations, implying less selection pressure acting on the functional variation at IA. Previous studies demonstrated that the difference of allele diversity at MHC genes would result from the difference of selection intensities acting on different loci. In humans, HLA-A and HLA-B was more polymorphic due to stronger selection pressure than HLA-C loci (Zemmour and Parham, 1992Zemmour J and Parham P (1992) Distinctive polymorphism at the HLA-C locus: Implications for the expression of HLA-C. J Exp Med 176:937-950. ). In the alpine newt, the DAB gene produced more mutations due to strong positive selection, while DBB gene produced only a small number of mutations due to lack of selection pressure (Babik et al., 2008Babik W, Pabijan M and Radwan J (2008) Contrasting patterns of variation in MHC loci in the alpine newt. Mol Ecol 17:2339-2355.).

According to the predictive results of 4 codon-based models, we got some evidence of balancing evidence acting on specific codons, as well as divergent selective pressure between IA alleles and non-IA alleles. Five PSSs were determined in non-IA alleles; all fell into the PBR and its vicinity, which was consistent with the signs of functionally important areas as targets for balancing selection (Zhang et al., 1998Zhang C, Anderson A and DeLisi C (1998) Structural principles that govern the peptide-binding motifs of class I MHC molecules. J Mol Biol 281:929-947.). A larger substitution happened at those specific codons of non-synonymous than synonymous under positive selection (Figure 2), demonstrating that these variations within the α1 domains might be functionally important (Kuduk et al., 2012Kuduk K, Johanet A, Allaine D, Cohas A and Radwan J (2012) Contrasting patterns of selection acting on MHC class I and class II DRB genes in the Alpine marmot (Marmota marmota). J Evol Biol 25:1686-1693. ). Contrarily, fewer PSSs were detected at IA. Therefore, selective pressure on IA alleles and non-IA alleles differed strikingly, and IA locus was less affected by selection pressure. In conclusion, differing levels of diversity and differential selection might indicate different functional roles for the goose loci. Moreover, although we did not detect recombination events in E2 alleles, we could not completely exclude the possibilities of inter-locus recombination.

The phylogenetic relationship in Anseriformes

In order to evaluate the goose phylogenetic relationship, we constructed a phylogenetic tree including domestic geese and some other birds. In Anseriformes, goose alleles did not cluster together, but scattered on the tree with other species of Anseriformes. The higher similarity of sequences between rather than within species (trans-species similarity) was usually explained with trans-species polymorphism (TSP). Under the effect of balanced selection and others, TSP was the maintenance and sharing of favorable functionally important alleles of immune-related genes between species. Although trans-species similarity could be also explained with convergent evolution, most avian MHC studies indicated that balanced TSP was a predominant mechanism responsible for trans-species alleles (Sallaberry-Pincheira et al., 2016Sallaberry-Pincheira N, González-Acuña D, Padilla P, Dantas GPM, Luna-Jorquera G, Frere E, Valdés-Velásquez A and Vianna JA (2016) Contrasting patterns of selection between MHC I and II across populations of Humboldt and Magellanic penguins. Ecol Evol 6:7498-7510. ), such as Accipitriform (Minias et al., 2019Minias P, Pikus E and Anderwald D (2019) Allelic diversity and selection at the MHC class i and class II in a bottlenecked bird of prey, the White-tailed Eagle. BMC Evol Biol 19:2. ), Passerines (Chen et al., 2015Chen W, Bei Y and Li H (2015) Genetic variation of the major histocompatibility complex (MHC class II B gene) in the threatened Hume’s pheasant, Syrmaticus humiae. PLoS One 10:e0116499. ; Balasubramaniam et al., 2016Balasubramaniam S, Bray RD, Mulder RA, Sunnucks P, Pavlova A and Melville J (2016) New data from basal Australian songbird lineages show that complex structure of MHC class II β genes has early evolutionary origins within passerines. BMC Evol Biol 16:112.; Gillingham et al., 2016Gillingham MAF, Courtiol A, Teixeira M, Galan M, Bechet A and Cezilly F (2016) Evidence of gene orthology and trans-species polymorphism, but not of parallel evolution, despite high levels of concerted evolution in the major histocompatibility complex of flamingo species. J Evol Biol 29:438-454.) and Spheniscidae (Kuduk et al., 2012Kuduk K, Johanet A, Allaine D, Cohas A and Radwan J (2012) Contrasting patterns of selection acting on MHC class I and class II DRB genes in the Alpine marmot (Marmota marmota). J Evol Biol 25:1686-1693. ).

Conclusions

In this study, we first examined the tissue expression levels of MHC I genes in one goose, conducing that in this goose there existed at least four classical sequences displaying differential expression patterns. It was the first time that we genotyped E2 of MHC I genes of 68 individuals from one YW-goose population by Illumina MiSeq sequencing, revealing that: 1) balancing selection might be the main force to shape the high MHC allelic variation in this goose population; 2) IA alleles and non-IA alleles exhibited differential genetic polymorphisms, and weaker positive selection detected at IA than non-IA alleles might account for much less variation of IA than non-IA alleles; 3) YW-goose alleles might have a trans-species polymorphism in Anseriformes.

Acknowledgments

This work was supported by the Doctoral Research Startup Funds Project of Qilu University of Technology (No. 81110260), and the Key Research and Development Program of Shandong Province (No. 2019YYSP024, 2019GNC106154) of China.

References

  • Babik W, Pabijan M and Radwan J (2008) Contrasting patterns of variation in MHC loci in the alpine newt. Mol Ecol 17:2339-2355.
  • Babik W, Taberlet P, Ejsmond MJ and Radwan J(2009) New generation sequencers as a tool for genotyping of highly polymorphic multilocus MHC system. Mol Ecol Resour 9:713-719.
  • Balasubramaniam S, Bray RD, Mulder RA, Sunnucks P, Pavlova A and Melville J (2016) New data from basal Australian songbird lineages show that complex structure of MHC class II β genes has early evolutionary origins within passerines. BMC Evol Biol 16:112.
  • Bartocillo F, Nishita Y, Abramov V and Masuda R (2021) Evolution of MHC class I genes in Japanese and Russian raccoon dogs, Nyctereutes procyonoides ( Carnivora : Canidae ). Mamm Res 66:371-383.
  • Bateson ZW, Whittingham LA, Johnson JA and Dunn PO (2015) Contrasting patterns of selection and drift between two categories of immune genes in prairie-chickens. Mol Ecol 24:6095-6106.
  • Biedrzycka A, O’Connor E, Sebastian A, Migalska M, Radwan J, Zając T, Bielański W, Solarz W, Ćmiel A, and Westerdahl H (2017) Extreme MHC class i diversity in the sedge warbler (Acrocephalus schoenobaenus); Selection patterns and allelic divergence suggest that different genes have different functions. BMC Evol Biol 17:159.
  • Bjorkman PJ and Parham P (1990) Structure, function, and diversity of class I major histocompatibility complex molecules. Annu Rev Biochem 59:253-288.
  • Bolger AM, Lohse M and Usadel B (2014) Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30:2114-2120.
  • Burri R, Hirzel HN, Salamin N, Roulin A and Fumagalli L (2008) Evolutionary patterns of MHC class II B in owls and their implications for the understanding of avian MHC evolution. Mol Biol Evol 25:1180-1191.
  • Chaves LD, Krueth SB and Reed KM (2009) Defining the Turkey MHC: Sequence and genes of the B Locus. J Immunol 183:6530-6537.
  • Chen W, Bei Y and Li H (2015) Genetic variation of the major histocompatibility complex (MHC class II B gene) in the threatened Hume’s pheasant, Syrmaticus humiae PLoS One 10:e0116499.
  • Chenani H, Nazari M, Beigi Nassiri MT, Roshanfekr H and Aghaie A (2021) Exonic SNP in MHC-DMB2 is associated with gene expression and humoral immunity in Japanese quails. Vet Immunol Immunop 239:110302.
  • Davies CJ and Antczak DF (1991) Mixed lymphocyte culture studies reveal complexity in the bovine MHC not detected by class I serology. Anim Genet 22:31-44.
  • Drews A and Westerdahl H (2019) Not all birds have a single dominantly expressed MHC-I gene: Transcription suggests that siskins have many highly expressed MHC-I genes. Sci Rep 9:19506.
  • Edgar RC and Flyvbjerg H (2015) Error filtering, pair assembly and error correction for next-generation sequencing reads. Bioinformatics 31:3476-3482.
  • Eimes JA, Lee SI, Townsend AK, Jablonski P, Nishiumi I and Satta Y (2016) Early duplication of a single MHC IIB locus prior to the passerine radiations. PLoS One 11:e0163456.
  • Furlong RF and Yang Z (2008) Diversifying and purifying selection in the peptide binding region of DRB in mammals. J Mol Evol 66:384-394.
  • Gillingham MAF, Courtiol A, Teixeira M, Galan M, Bechet A and Cezilly F (2016) Evidence of gene orthology and trans-species polymorphism, but not of parallel evolution, despite high levels of concerted evolution in the major histocompatibility complex of flamingo species. J Evol Biol 29:438-454.
  • Grimholt U, Olsaker I, Lingaas F and Lie (2015) Sequence analysis of MHC class Ⅰ α2 domain exon variants in one diploid and two haploid atlantic salmon pedigrees. Anim Genet 28:385-390.
  • Grogan KE, Mcginnis GJ, Sauther ML, Cuozzo FP and Drea CM (2016) Next-generation genotyping of hypervariable loci in many individuals of a non-model species: Technical and theoretical implications. BMC Genom 17:204.
  • Hedrick PW (1998) Balancing selection and MHC. Genet 104:207-214.
  • Hughes AL and Nei M (1989) Nucleotide substitution at major histocompatibility complex class II loci: Evidence for overdominant selection. Proc Natl Acad Sci U S A 86:958-962.
  • Hughes AL and Yeager M (1998) Natural selection at major histocompatibility complex loci of vertebrates. Annu Rev Genet 32:415-435.
  • Kaufman J (2020) From Chickens to Humans: The importance of peptide repertoires for MHC Class I Alleles. Front Immunol 11:601089.
  • Kaufman J, Andersen R, Avila D, Engberg J, Lambris J, Salomonsen J, Welinder and Skjødt K (1992) Different features of the MHC class I heterodimer have evolved at different rates. Chicken BF and beta 2-microglobulin sequences reveal invariant surface residues. J Immunol 148:1532-1546.
  • Kaufman J, Milne S, Göbel TWF, Walker BA, Jacob JP, Auffray C, Zoorob R and Beck S (1999) The chicken B locus is a minimal essential major histocompatibility complex. Nature 401:923-925.
  • Kaufman J, Salomonsen J and Flajnik M (1994) Evolutionary conservation of MHC class I and class II molecules--different yet the same. Semin Immunol 6:411-424.
  • Kennedy L, Ryvar R, Gaskell R, Addie D, Willoughby K, Carter S, Thomson W, Ollier W and Radford A (2002) Sequence analysis of MHC DRB alleles in domestic cats from the United Kingdom. Immunogenetics 54:348-352.
  • Kimura M (1980) A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J Mol Evol 16:111-120.
  • Klein J, Bontrop RE, Dawkins RL, Erlich HA, Gyllensten UB, Heise ER, Jones PP, Parham P, Wakeland EK and Watkins DI (1990) Nomenclature for the major histocompatibility complexes of different species: A proposal. Immunogenetics 31:217-219.
  • Klein J, Sato A and Nikolaidis N (2007) MHC, TSP, and the origin of species: From immunogenetics to evolutionary genetics. Annu Rev Genet 41:281-304.
  • Knafler GJ, Clark JA, Boersma PD and Bouzat JL (2012) MHC diversity and mate choice in the magellanic penguin, Spheniscus magellanicus J Hered 103:759-768.
  • Kuduk K, Johanet A, Allaine D, Cohas A and Radwan J (2012) Contrasting patterns of selection acting on MHC class I and class II DRB genes in the Alpine marmot (Marmota marmota). J Evol Biol 25:1686-1693.
  • Kumar S, Stecher G, Li M, Knyaz C and Tamura K (2018) MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547-1549.
  • Livant EJ, Brigati JR and Ewald SJ (2004) Diversity and locus specificity of chicken MHC B class I sequences. Anim Genet 35:18-27.
  • Manjula P, Fulton JE, Seo D and Lee JH (2021) Comparison of major histocompatibility complex-B variability in Sri Lankan indigenous chickens with five global chicken populations using MHC-B SNP panel. Anim Genet 52:824-833.
  • Mesa CM, Thulien KJ, Moon DA, Veniamin SM and Magor KE (2004) The dominant MHC class I gene is adjacent to the polymorphic TAP2 gene in the duck, Anas platyrhynchos Immunogenetics 56:192-203.
  • Minias P, Pikus E and Anderwald D (2019) Allelic diversity and selection at the MHC class i and class II in a bottlenecked bird of prey, the White-tailed Eagle. BMC Evol Biol 19:2.
  • Moon DA, Veniamin SM, Parks-Dely JA and Magor KE (2005) The MHC of the Duck ( Anas platyrhynchos ) contains five differentially expressed class I genes . J Immunol 175:6702-6712.
  • Nei M and Gojobori T (1986) Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions. Mol Biol Evol 3:418-426.
  • O’Connor EA, Strandh M, Hasselquist D, Nilsson J and Westerdahl H (2016) The evolution of highly variable immunity genes across a passerine bird radiation. Mol Ecol 25:977-989.
  • Oliver MK, Lambin X, Cornulier T and Piertney SB (2009) Spatio-temporal variation in the strength and mode of selection acting on major histocompatibility complex diversity in water vole (Arvicola terrestris) metapopulations. Mol Ecol 18:80-92.
  • Pardal S, Drews A, Alves JA, Ramos JA and Westerdahl H (2017) Characterization of MHC class I in a long distance migratory wader, the Icelandic black-tailed godwit. Immunogenetics 69:463-478.
  • Piertney SB and Oliver MK (2006) The evolutionary ecology of the major histocompatibility complex. Hered 96:7-21.
  • Potts WK and Slev PR (1995) Pathogen‐based models favoring MHC genetic diversity. Immunol Rev 143:181-197.
  • Sallaberry-Pincheira N, González-Acuña D, Padilla P, Dantas GPM, Luna-Jorquera G, Frere E, Valdés-Velásquez A and Vianna JA (2016) Contrasting patterns of selection between MHC I and II across populations of Humboldt and Magellanic penguins. Ecol Evol 6:7498-7510.
  • Sambrook J, Fritsch EF and Maniatis T (1990) Molecular cloning: A laboratory manual. Cold Spring Harbor Laboratories 186:182-183.
  • Shum BP, Rajalingam R, Magor KE, Azumi K, Carr WH, Dixon B, Stet RJM, Adkison MA, Hedrick RP and Parham P (1999) A divergent non-classical class I gene conserved in salmonids. Immunogenetics 49:479-490.
  • Sommer S (2005) The importance of immune gene variability (MHC) in evolutionary ecology and conservation. Front Zool 2:16.
  • Spurgin LG and Richardson DS (2010) How pathogens drive genetic diversity: MHC, mechanisms and misunderstandings. Proc R Soc B: Biol Sci 277:979-988.
  • Takahata N and Nei M (1990) Allelic genealogy under overdominant and frequency-dependent selection and polymorphism of major histocompatibility complex loci. Genetics 124:967-978.
  • Tamura K (1992) Estimation of the number of nucleotide substitutions when there are strong transition-transversion and G+C-content biases. Mol Biol Evol 9:678-687.
  • Wallny HJ, Avila D, Hunt LG, Powell TJ, Riegert P, Salomonsen J, Skjødt K, Vainio O, Vilbois F, Wiles MV et al (2006) Peptide motifs of the single dominantly expressed class I molecule explain the striking MHC-determined response to Rous sarcoma virus in chickens. Proc Natl Acad Sci U S A 103:1434-1439.
  • Westerdahl H, von Schantz T and Wittzell H (1999) Polymorphism and transcription of Mhc class I genes in a passerine bird, the great reed warbler. Immunogenetics 49:158-170.
  • Xia C, Hu T, Yang T, Wang L, Xu G and Lin C (2005) cDNA cloning, genomic structure and expression analysis of the goose (Anser cygnoides) MHC class I gene. Vet Immunol Immunop 107:291-302.
  • Xia C, Lin CY, Xu GX, Hu TJ and Yang TY (2004) cDNA cloning and genomic structure of the duck (Anas platyrhynchos) MHC class I gene. Immunogenetics 56:304-309.
  • Yang Z (2007) PAML 4: Phylogenetic analysis by maximum likelihood. Mol Biol Evol 24:1586-1591.
  • Zagalska-Neubauer M, Babik W, Stuglik M, Gustafsson L, Cichoń M and Radwan J (2010) 454 sequencing reveals extreme complexity of the class II Major Histocompatibility Complex in the collared flycatcher. BMC Evol Biol 10:395.
  • Zemmour J and Parham P (1992) Distinctive polymorphism at the HLA-C locus: Implications for the expression of HLA-C. J Exp Med 176:937-950.
  • Zeng QQ, He K, Sun DD, Ma MY, Ge YF, Fang SG and Wan QH (2016) Balancing selection and recombination as evolutionary forces caused population genetic variations in golden pheasant MHC class I genes. BMC Evol Biol 16:42.
  • Zhang C, Anderson A and DeLisi C (1998) Structural principles that govern the peptide-binding motifs of class I MHC molecules. J Mol Biol 281:929-947.

Edited by

Associate Editor:

Antonio Matteo Solé-Cava

Publication Dates

  • Publication in this collection
    15 July 2024
  • Date of issue
    2024

History

  • Received
    06 Oct 2022
  • Accepted
    09 May 2024
Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
E-mail: editor@gmb.org.br