Next Article in Journal
De  Novo Transcriptome Sequencing and Profiling of Ovarian Development of Argas persicus Along the Trophogonic Cycle
Previous Article in Journal
Mixture Deconvolution with Massively Parallel Sequencing Data: Microhaplotypes Versus Short Tandem Repeats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genetic Diversity and Selection of MHC I-UAA in Clariid Catfish from Thailand: Implications for Breeding and Conservation

by
Ton Huu Duc Nguyen
1,2,3,
Piangjai Chalermwong
1,
Chananya Patta
1,
Wattanawan Jaito
1,
Worapong Singchat
1,2,4,*,
Thitipong Panthum
1,4,
Trifan Budi
1,5,
Kednapat Sriphairoj
6,
Sittichai Hatachote
6,
Prapansak Srisapoome
1,7,
Narongrit Muangmai
1,8,
Darren K. Griffin
1,9,
Agostinho Antunes
1,10,11,
Prateep Duengkae
1,4 and
Kornsorn Srikulnath
1,2,4,12,13,*
1
Animal Genomics and Bioresource Research Unit (AGB Research Unit), Faculty of Science, Kasetsart University, 50 Ngamwongwan, Chatuchak, Bangkok 10900, Thailand
2
Interdisciplinary Graduate Program in Bioscience, Faculty of Science, Kasetsart University, 50 Ngamwongwan, Chatuchak, Bangkok 10900, Thailand
3
Faculty of Biology Education, School of Education, Can Tho University, 3/2 Street, Ninh Kieu Ward, Can Tho 900000, Vietnam
4
Special Research Unit for Wildlife Genomics (SRUWG), Department of Forest Biology, Faculty of Forestry, Kasetsart University, 50 Ngamwongwan, Chatuchak, Bangkok 10900, Thailand
5
School of Agricultural Technology, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
6
Faculty of Natural Resources and Agro-Industry, Kasetsart University Chalermphrakiat Sakon Nakhon Province Campus, Sakon Nakhon 47000, Thailand
7
Department of Aquaculture, Faculty of Fisheries, Kasetsart University, Bangkok 10900, Thailand
8
Department of Fishery Biology, Faculty of Fisheries, Kasetsart University, Bangkok 10900, Thailand
9
School of Natural Sciences, University of Kent, Canterbury CT2 7NJ, UK
10
Interdisciplinary Centre of Marine and Environmental Research, Terminal de Cruzeiros do Portode Leixões, University of Porto, 4450-208 Porto, Portugal
11
Department of Biology, Faculty of Sciences, University of Porto, 4169-007 Porto, Portugal
12
Department of Genetics, Faculty of Science, Kasetsart University, Chatuchak, Bangkok 10900, Thailand
13
Biodiversity Center, Kasetsart University (BDCKU), Bangkok 10900, Thailand
*
Authors to whom correspondence should be addressed.
Genes 2025, 16(9), 1106; https://doi.org/10.3390/genes16091106
Submission received: 1 September 2025 / Revised: 16 September 2025 / Accepted: 17 September 2025 / Published: 18 September 2025
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

Background/Objectives: Understanding variabilities in the Major Histocompatibility Complex class I (MHC I) gene is essential for evaluating immunogenetic diversity in clariid catfish. MHC I plays a critical role in immune defense by presenting endogenous antigens to cytotoxic T cells. Therefore, we aimed to investigate the genetic diversity, selection patterns, and phylogenetic relationships of MHC I alleles in three important clariid catfish species (Clarias gariepinus, Clarias macrocephalus, and Clarias batrachus) across wild and hatchery populations in Thailand. Methods: Targeted next-generation sequencing of a 174 bp fragment partial exon 6 of MHC I-UAA gene was performed, along with phylogenetic analyses, neutrality tests and dN/dS analyses. Results: Overall, 91 novel alleles were identified in 674 individuals, all of which were novel (100% novelty), with none matching existing reference sequences, thereby revealing extensive variation in population-specific variants. Phylogenetic analyses revealed allele sharing among species, which was consistent with balanced selection. Neutrality tests and dN/dS analyses provided evidence of both purifying and diversifying selection, with episodic positive selection detected at multiple codon sites associated with the antigen-binding α1 domain. Distinct selection patterns among populations, influenced by local environmental conditions and human pressures, along with high allele richness, are reflected in the diversity of immunogenetic variations. Conclusions: These findings provide critical insights into immune adaptation and highlight the potential of MHC I as a functional marker for genetic monitoring. Although a causal relationship between MHC I polymorphism and disease resistance is debated, studies suggest associations with pathogen survival, indicating future implications for aquaculture breeding and conservation, particularly in marker-assisted selection for broodstock management in Thailand.

1. Introduction

Aquaculture is recognized as the fastest-growing sector in global animal food production, supplying more than half of the fish consumed worldwide. Among the species of farmed fish, clariid catfish (Clarias spp.), favored for their rapid growth, environmental tolerance, and consumer demand, are especially important in Southeast Asia. African catfish (C. gariepinus) is the most widely farmed, bighead catfish (C. macrocephalus) is native to Southeast Asia, and walking catfish (C. batrachus) occurs mainly in South Asia. In Thailand, the total annual fish production, which is approximately 90,000 tons [1], is insufficient to fully supply local communities. Consequently, intensive expansion has relied on a narrow genetic base, raising concerns over reduced diversity and inbreeding in hatchery stocks [2,3]. Broodstock, often derived from few individuals and rarely replenished with wild genetics, show declining fitness and immune function over time [4,5,6]. Additionally, wild populations, are increasingly affected by pollution, salinity shifts, and habitat fragmentation, which erode local adaptations [7,8]. The rapid expansion of C. gariepinus farming and its hybridization with C. macrocephalus, combined with unregulated breeding practices, have intensified genetic mixing. As a consequence, production declined between 2013 and 2023, with reduced disease resistance likely linked to genetic outbreeding and unrecorded parental stocks [9,10,11]. Moreover, resistance traits are believed to be threatened by domestication, anthropogenic pressures, and diseases, highlighting the importance of understanding their underlying genetic resources [12].
The Major Histocompatibility Complex class I (MHC I) gene is a critical component of the vertebrate immune system that plays a central role in recognizing and eliminating intracellular pathogens, including viruses and certain bacteria [13,14]. As a key element in the adaptive immune response, MHC I encodes a transmembrane glycoprotein that presents endogenously derived peptide antigens to cytotoxic T lymphocytes, thereby initiating downstream immune reactions [15,16]. The MHC type I region in teleosts exhibits extensive polymorphism and enables populations to adapt to diverse and evolving pathogenic environments [17,18]. This polymorphism is maintained by balancing selection mechanisms such as heterozygote advantage, frequency-dependent selection, and fluctuating pathogen-driven pressures [19]. By contrast, variations in MHC I genes influence disease susceptibility in clariid catfish, affecting individual immune responses and important traits such as survival, reproduction, and resilience in high-density natural and aquaculture settings [20,21]. MHC I is also increasingly used as a functional genetic marker to evaluate the impact of environmental stressors, population bottlenecks, domestication, and habitat degradation [22,23]. However, unlike microsatellites, which track neutral variation, MHC loci directly reflect adaptive evolution [24]. MHC I is also considered a powerful marker in ecological and applied research, including conservation genetics and sustainable aquaculture, because of its ability to reflect adaptive variations. In hatcheries, immune genes, such as MHC, which are affected by reduced natural selection and increased artificial selection, can experience lowered adaptability and survival [25,26]. Furthermore, decreased MHC diversity, which has been linked to increased disease vulnerability and lower survival, has been observed in several freshwater fish groups such as salmonids and cyprinids [27,28].
Consequently, understanding MHC I variations, which are essential for designing effective breeding strategies for clariid catfish to enhance disease resistance without reducing genetic diversity, can also be used to compare wild and hatchery populations and identify population-specific alleles for genetic monitoring [17,25,29]. Despite its importance, the diversity of MHC I in clariid catfish and populations in Southeast Asia, including Thailand, remains poorly studied. To date, no prior comprehensive study has investigated MHC I diversity in Thai Clarias populations, representing a critical knowledge gap that impedes efficient hatchery management. This knowledge gap hampers effective hatchery management and highlights the urgent need to improve our understanding of advances in immunogenetics and resource management. Therefore, this study aimed to assess MHC I variations in C. gariepinus, C. macrocephalus, and C. batrachus in Thailand by (1) evaluating allelic diversity in wild and hatchery populations, (2) identifying signatures of natural selection, (3) reconstructing phylogenetic relationships, and (4) comparing amino acid variation and structural features.

2. Materials and Methods

2.1. Specimen Collection and DNA Extraction

A total of 674 individuals from three catfish species (246 individuals of C. gariepinus from five sites, 420 individuals of C. macrocephalus from 14 sites, and eight individuals of C. batrachus from one site) were collected from 20 sampling sites (wild and hatchery) across Thailand (Tables S1 and S2, Figure S1). Tissue samples (approximately 0.3 × 0.3 cm) were excised from the caudal fins of individual specimens and preserved in 1.5 mL microcentrifuge tubes containing 95% ethanol, followed by storage at 4 °C until subsequent analyses. Caudal fins were chosen because sampling causes minimal harm to the fish and these tissues have a natural ability to regenerate. Samples were collected with prior consent from fish farm owners or relevant authorities, and all individuals were immediately returned to their original habitats after sampling. Wild fish were sampled non-lethally by fin clipping. After recovery, individuals were released at their site of capture. No wild fish were retained in the laboratory. All experimental procedures, including animal care, were reviewed and approved by the Animal Experiment Committee of Kasetsart University (approval no. ACKU65-SCI-003, ACKU66-SCI-006, and ACKU66-SCI-014). These procedures were conducted in accordance with the Regulations on Animal Experiments of Kasetsart University ARRIVE guidelines (https://arriveguidelines.org/, accessed on 6 May 2025). Genomic DNA was extracted using the standard salting-out method, as described by Supikamolseni [30]. The quality and quantity of DNA were evaluated using 1% agarose gel electrophoresis and a NanoDrop 2000 Spectrophotometer (Thermo Fisher Scientific, Wilmington, DE, USA). DNA concentrations were further quantified using a Qubit 4.0 Fluorometer (Invitrogen, Carlsbad, CA, USA) to ensure accurate measurements.

2.2. PCR Amplification and IlluminaTM Short-Read Sequencing

A partial fragment of exon 6 of the MHC I-UAA gene, which contains single nucleotide polymorphisms previously associated with immune traits in teleost, was amplified using the primer pair 18MHCI_catfish_F (5′-GGTGTGGGTATACTGATGGGTC-3′) and 18MHCI_catfish_R (5′-TCAGGTAGTCCTCTGTTCCCTT-3′) [31], yielding an expected product size of 174 bp. Eight bp sample-specific barcode sequences were added to the 5′-end of each forward primer (Macrogen Inc., Seoul, Republic of Korea). PCR amplification was performed in a 15 µL reaction containing 50 ng of genomic DNA, 1× Apsalagen reaction buffer, 1.5 mM MgCl2, 0.2 mM dNTPs, 0.5 µM of each primer, and 0.5 U of Taq DNA polymerase (Apsalagen Co., Ltd., Bangkok, Thailand). The thermal cycling protocol included an initial denaturation at 94 °C for 10 min, followed by 35 cycles of denaturation at 94 °C for 30 s, annealing at 57 °C for 30 s, and extension at 72 °C for 1 min, with a final extension at 72 °C for 10 min. The PCR products were visualized by electrophoresis using a 1.5% agarose gel. Each sample was processed in triplicate to minimize allelic dropouts and ensure amplification reliability. A total of 92 samples per pool set with each barcode primer were pooled into nine sets and sent for paired-end short-read sequencing on the Illumina NovaSeqTM 6000 platform (Novogene Co., Ltd., Singapore, Singapore). This pooling strategy provided sufficient sequencing depth, yielding an average coverage of >100× per allele after quality filtering.

2.3. Sequence Quality Control and Read Processing

FastQC version 0.11.9 was used to evaluate the quality of the 174 bp paired-end reads [32]. Reads with Phred scores (q) > 20 were retained for downstream analyses, as recommended for high-throughput amplicon sequencing to ensure reliable base calls [33]. Libraries were prepared for each individual without pooling, and unique dual indices were used for multiplexing. After filtering, each individual had on average ~3500 reads for the target amplicon, with a range of 1200–5000 reads, ensuring sufficient coverage for reliable allele detection. All subsequent processing steps were performed using AmpliSAS version 1.0 [33], which performs quality filtering, read merging, and the assignment of sequences to individual samples. Amplicons corresponding to allelic variants of MHC I genes were clustered, filtered, and assigned to individuals based on the number of alleles per sample [33]. A minimum amplicon depth threshold of 100 reads was used to eliminate the background noise and low-read samples, in accordance with AmpliSAS guidelines for minimizing artifacts and false allele calls [33,34]. Variants occurring at <1% frequency within an amplicon were discarded to minimize sequencing artifacts. The number of true alleles was determined using the degree of change criterion based on the sequencing depth, followed by the default configuration for other parameters [34]. The maximum number of alleles per individual was set to 10 to account for potential duplications in MHC I of teleosts [16]. This threshold is consistent with reports of multiple duplicated MHC I loci in teleost fishes, including catfish, where individuals may carry more than two alleles per locus [16,35]. Sequences were subsequently aligned and mapped to the MHC I reference gene sequence of C. gariepinus (accession number: XM53477511) and translated into amino acid sequences using Geneious Prime version 2023.0.4 (https://www.geneious.com, accessed on 6 May 2025) to check for the presence of stop codons within the exons. The C. gariepinus reference was selected because it represents the most complete and annotated MHC I sequence currently available among Clarias species, whereas full-length references for C. macrocephalus and C. batrachus are not available [36].

2.4. Genetic Diversity and Phylogenetic Analysis

Nucleotide diversity (π) was estimated using DnaSP version 6.12 [37]. Analysis of molecular variance (AMOVA) was performed using R version 4.1.2 [38] with the “poppr” package to investigate the genetic structure of C. gariepinus, C. macrocephalus, and C. batrachus populations, and statistical significance was assessed using 1000 permutations [39]. Nucleotide sequences of the MHC I gene from C. gariepinus, C. macrocephalus, and C. batrachus were retrieved from the NCBI database using the BLASTn tool (https://blast.ncbi.nlm.nih.gov/Blast.cgi, accessed on 6 May 2025) with filtering thresholds set at a sequence similarity of >70% and a query coverage of >85%. The best-fit nucleotide substitution model was selected based on the lowest Bayesian Information Criterion (BIC), as determined by ModelFinder, and the GTR + G model was applied in the subsequent analyses [40]. All retrieved MHC I gene sequences were subsequently used to construct a Bayesian phylogenetic tree using MrBayes in Geneious Prime version 2023.0.4 (https://www.geneious.com, accessed on 6 May 2025) [41]. Five Markov Chain Monte Carlo (MCMC) chains were run for 2 million generations, with trees sampled every 5000 generations and a burn-in of 200,000 to illustrate the evolutionary relationships among MHC I alleles in catfish. The resulting phylogenetic tree was visualized using Interactive Tree of Life version 5 [42] (https://itol.embl.de/, accessed on 6 May 2025).

2.5. Selection Analysis

Neutrality tests, including Tajima’s D, Fu and Li’s F*, and D* were performed using DnaSP version 6.12 [37] to assess deviations from neutrality and identify potential signals of selection at the MHC I locus. Analyses were conducted for each population using the respective sample sizes (Table 1). The type of selective pressure acting on MHC I in clariid catfish was evaluated using the non-synonymous to synonymous substitution rate ratio (dN/dS or ω). The average number of synonymous (dS) and nonsynonymous (dN) substitutions per site was estimated using the Nei–Gojobori method [43], with Jukes–Cantor correction, implemented in Molecular Evolutionary Genetics Analysis (MEGA) version X [44]. A value of ω close to 1 indicates neutral selection, ω > 1 suggests positive selection, and ω < 1 indicates purifying selection. A suite of codon-based models implemented on the Datamonkey web server (https://www.datamonkey.org, accessed on 6 June 2025) was used to detect signals of natural selection acting on individual codon sites in a partial fragment of exon 6 of MHC I genes. A mixed-effects model of evolution (MEME) was used to identify codons under episodic diversifying selection, in which selection acts on a subset of branches at a given site [45]. The analysis was run with a significance threshold of p ≤ 0.01. Fixed Effects Likelihood (FEL) was applied to detect pervasive selection by estimating the non-synonymous and synonymous substitution rates at each site across the entire phylogeny. Sites with p ≤ 0.01 were considered significant [46]. Fast Unconstrained Bayesian AppRoximation (FUBAR) was used to infer pervasive selection using a Bayesian framework, with posterior probability ≥ 0.9 used as the threshold for significance [47].

2.6. Multiple Sequence Alignment of MHC I Amino Acid Residues

Reference sequences of the MHC I gene in C. gariepinus (XM53477511, MG545605 and EU714321), C. batrachus (KC750210), Silurus meridionalis (M46842361 and M46859545), and Tachysurus fulvidraco (KP881737), were retrieved from the NCBI database using the BLASTp tool (https://blast.ncbi.nlm.nih.gov/Blast.cgi, accessed on 6 June 2025). BLASTp analysis was performed using filtering thresholds of >60% sequence similarity and >85% query coverage, as the study sequences had relatively low similarity to available references, and stricter cutoffs might have excluded important functional variants [48,49]. Subsequently, MHC I amino acid sequences from this study and the reference teleost species were aligned using ClustalW in Geneious Prime v2023.0.4 (https://www.geneious.com, accessed on 6 June 2025). The resulting alignment was trimmed to 54 amino acids, and a Bayesian phylogenetic tree was constructed as mentioned above. Secondary protein structures were predicted using EMBOSS tool version 6.5.7 (http://emboss.sourceforge.net/, accessed on 6 June 2025).

3. Results

3.1. Genetic Diversity of Catfish Based on MHC I Gene

In total, 174 bp of the MHC I partial fragment were obtained. Nucleotide diversity of C. gariepinus was 0.095 ± 0.017; C. macrocephalus was 0.133 ± 0.033 and C. batrachus was 0.087 (Table 1). The number of alleles per individual ranged from 1 to 9, with over 80% of the samples possessing between 1 and 4 alleles. The number of alleles per population ranged from 2 to 38 in C. macrocephalus, 3–37 in C. gariepinus, and six in C. batrachus. The highest allele richness was observed for C. macrocephalus from Sakon Nakhon 1 with 38 alleles, followed by C. gariepinus from Kalasin 2 with 37 alleles, and C. batrachus from Ubon Ratchathani with six alleles. Compared to the reference MHC I gene sequence of C. gariepinus (accession number: XM53477511), 82 polymorphic sites were identified in all individuals, defining 91 newly identified alleles (Table S3). These alleles accounted for 100% novel sequences, as none overlapped with previously reported alleles [36]. Notably, Clarias_MHCI*TH2 and Clarias_MHCI*TH7 were the most prevalent alleles, detected in 15 of the 20 populations. Clarias_MHCI*TH15, Clarias_MHCI*TH19, Clarias_MHCI*TH23, Clarias_MHCI*TH33, Clarias_MHCI*TH37, Clarias_MHCI*TH39, Clarias_MHCI*TH42, and Clarias_MHCI*TH43 were unique to the KSN2 population. Clarias_MHCI*TH62, Clarias_MHCI*TH84—Clarias_MHCI*TH88 were specific to the SKN1 population, whereas Clarias_MHCI*TH64, Clarias_MHCI*TH66, Clarias_MHCI*TH67, Clarias_MHCI*TH69, and Clarias_MHCI*TH70 were unique to the STN population. Meanwhile, Clarias_MHCI*TH72 was exclusively found in the NST2 population. Alleles Clarias_MHCI*TH74, Clarias_MHCI*TH75, Clarias_MHCI*TH77, Clarias_MHCI*TH80, and Clarias_MHCI*TH81 were specific, in the SKN2 population, whereas Clarias_MHCI*TH90 and Clarias_MHCI*TH91 were unique to SKN3 and KSN1 populations, respectively. At the species level, eight alleles were observed only in C. gariepinus within this dataset, whereas 18 alleles were specifically found in C. macrocephalus. No species-specific alleles were detected in C. batrachus (Table S3).
The Bayesian phylogenetic tree, which revealed the clustering of newly characterized alleles (Clarias_MHCI_TH) into distinct clades, supported the divergence of MHC I alleles among the three clariid and reference species (Figure 1). AMOVA indicated that 43% of the genetic variation in MHC I was attributable to differences among individuals (p < 0.001), whereas 11% was due to differences among populations (p = 0.02) (Table S4).

3.2. Selection and Selective-Sweep Analyses of Clariid Catfish

Neutrality tests of the partial MHC type I gene fragment revealed differences in Tajima’s D, and Fu and Li’s D and F values across several populations of C. gariepinus, C. macrocephalus, and C. batrachus (Table 2). Tajima’s D values ranged from −1.700–0.604, which were not statistically significant. By contrast, Fu and Li’s D* values ranged from −3.082–1.356, with significant deviations from neutrality in C. macrocephalus populations SNK1-CM-W, SNK2-CM-W, and SNK4-CM-C (Table 2). Similarly, Fu and Li’s F* values ranged from −2.800–1.001, with significant results in SNK1-CM-W, SNK2-CM-W, and SNK4-CM-C (Table 2). The MHC I gene showed an average ω value of 1.563, ranging from 0.678 to 10.000. All populations of C. gariepinus had ω > 1 (ranging from 1.750 to 3.200). Meanwhile, SNK2-CM-W, SB-CM-C, and SNK4-CM-C populations in C. macrocephalus exhibited ω > 1, while others exhibited ω < 1. The ω = 1.103 for C. batrachus (Table 3). Codon-level selection analysis using MEME revealed that 39 of the 54 codon sites were subjected to episodic diversifying selection (p < 0.01). On average, 5.08 branches per selected site exhibited evidence of diversifying selection, highlighting lineage-specific adaptive evolution (Table S5). The FEL analysis revealed evidence of selection at two codon sites within the MHC I gene region (Table S6). Notably, codon 49 was found to be under diversified selection (p < 0.01). By contrast, codon 2 was under purifying selection (p < 0.01), suggesting functional constraints and evolutionary conservation. Additionally, FUBAR analysis identified two sites (codons 1 and 2) evolving under pervasive purifying selection and four sites (codons 10, 12, 48, and 49) evolving under pervasive diversifying selection based on posterior probability thresholds (>0.9) and supported by Bayes Factors (Table S7).

3.3. Multiple Sequence Alignment of MHC I Amino Acid Residues and Prediction

Overall, 54 amino acid residues of the partial fragment of the MHC I gene from 91 alleles were detected in the populations of C. gariepinus, C. macrocephalus, and C. batrachus. They showed low similarity with those of C. gariepinus (XM53477511, MG545605 and EU714321; range 6.45–38.71%), C. batrachus (KC750210; range 6.45–35.48%), S. meridionalis (M46842361 and M46859545; range 6.45–32.26%), and T. fulvidraco (KP881737; range 16.12–32.87%). The very low similarity values reflect the high polymorphism characteristic of MHC I genes, where strong diversifying selection maintains extensive amino acid diversity to enhance pathogen recognition. These alleles primarily encode amino acids located in the α1 domain of the MHC I molecule (Figure S2). In total, 258 mutations were identified, including 26 silent, 215 missense, and 17 nonsense mutations (Table S8). Bayesian phylogenetic analysis based on the amino acid sequences of MHC I alleles revealed no distinct clustering patterns among the clariid catfish (Figure 2).

4. Discussion

4.1. MHC I-UAA Diversity in Clariid Catfish

This study revealed remarkable genetic diversity in exon 6 of MHC I-UAA, with 47 polymorphic sites and 91 novel alleles identified from a short 174 bp fragment in clariid catfish, indicating high variability compared to Salmo salar (10–20 alleles), Cyprinus carpio (5–10 alleles), and Oreochromis niloticus (31 alleles), which have fewer alleles [16,27,50]. This is consistent with previous studies reporting that a large allelic diversity of MHC genes has been observed in clariid catfish, which are known for their exceptional environmental adaptability and broad pathogen resistance among teleosts [16,17,19,51,52,53]. Among the three clariid catfish evaluated in this study, C. macrocephalus, which exhibited the highest allele diversity with 59 alleles in 14 populations, is linked to its larger sampling sites, which are subject to varying selection pressures in hatchery and wild environments. However, from a phylogenetic perspective, no species-specific clade or clustering was observed in the MHC I-UAA gene sequences based on nucleotide and amino acid data. This allele sharing and lack of clear clustering may be influenced by recent divergence times, which are less than 20 million years, allowing for the retention of ancestral polymorphisms [54]. Additionally, shared alleles across species, such as Clarias_MHCITH12 and Clarias_MHCITH34, may result from historical introgression, which are associated with balanced selection at immune loci, such as MHC [55,56]. The persistence of identical or highly similar alleles across species boundaries is a hallmark of trans-species polymorphism, whereby balancing selection maintains ancient allelic lineages over extended evolutionary timescales despite speciation events [19]. Similar interspecific allele sharing, which is maintained across species due to long-term balancing selection, has been observed in other teleosts such as salmonids (S. salar, Oncorhynchus mykiss, and Oncorhynchus kisutch) and Etheostoma species (Etheostoma flabellare, Etheostoma caeruleum, and Etheostoma spectabile) [57,58]. Thus, evolutionary proximity supports the plausibility of shared MHC I alleles, despite their morphological and ecological differences. The newly identified MHC I-UAA alleles in this study demonstrated relatively low amino acid similarity to reference sequences and closely related species, such as C. batrachus, S. meridionalis, and T. fulvidraco. This suggests that they either follow independent evolutionary paths or persist as ancient, highly divergent lineages that have not been eliminated by selection [59]. Moreover, several amino acid substitutions, especially at residues forming β-strands and α-helices in the α1 domain, which are important for maintaining the structure of the peptide-binding groove, were detected in this study. This groove, which directly binds to endogenous peptides and presents them to cytotoxic T cells, is critical for immune function [15,18]. Thus, the substitutions, which may alter the local secondary structure of the peptide-binding groove and, in turn, affect folding stability and peptide-binding capacity of MHC I molecules, with possible consequences for pathogen recognition [60,61]. In addition to the high allelic diversity and trans-species polymorphisms observed, several experimental studies in C. gariepinus have demonstrated functional links between MHC I variation and disease resistance. It was shown by Azis et al. [20] and Alimuddin et al. [62] that offspring with specific MHC I markers had significantly higher survival rates after Aeromonas hydrophila infection. Imron et al. [31] identified a unique allele, Clg-UAA*07, amplified only in resistant individuals, suggesting a direct connection between MHC I polymorphism and pathogen survival. Additionally, Oyebola et al. reported differential expression of MHC I in challenged C. gariepinus, further supporting its role in immune response. Although the causal relationship remains debated [18], these findings imply that MHC I diversity could enhance resilience against pathogens and serve as a basis for marker-assisted selection in aquaculture.

4.2. Evidence of Selection at the MHC I-UAA Locus

Positive selection acting on the MHC I-UAA locus was supported by high ω values and codon-level evidence from MEME, FEL, and FUBAR analyses; however, allelic frequency patterns did not indicate dominance by one or two alleles. This suggests that positive selection may not act through selective sweeps favoring specific alleles but rather through mechanisms such as heterozygous advantage or fluctuating pathogen pressures that maintain diversity. Unlike a selective sweep, where one or a few alleles rapidly rise to dominance, balancing selection preserves multiple alleles within populations [19]. Additionally, results of Fu and Li’s neutrality tests further support the recent or ongoing selection with selective sweeps and population expansion, as evidenced by significantly negative Fu and Li’s D* and F* values in several C. macrocephalus populations. By contrast, Tajima’s D test did not reach statistical significance overall. Nevertheless, these ω values could reflect localized episodic selection without fixation, especially in hatchery populations, which are subjected to environmental instability and artificial selection [16,19]. Moreover, the conditions described may influence the selection regime acting on immune genes, where fluctuating or relaxed selection can temporarily increase ω without causing long-term fixation. For instance, population-specific alleles associated with localized environmental factors, limited gene flow, and demographic history suggest effects of domestication, artificial selection, or genetic drift from restricted broodstock diversity. These alleles may also indicate local adaptation, especially if linked to immune regions such as the MHC, shaped by pathogen communities, environmental pressures, and management practices in aquaculture. For practical breeding, such population-specific alleles can guide conservation of genetically diverse stocks (e.g., SNK2-CM-W) that serve as valuable reservoirs for future broodstock improvement and disease-resilient breeding programs. Such environments often exhibit directional and episodic selection owing to exposure to fluctuating or novel pathogens [3,5,23,63,64]. For example, strong diversifying selection for MHC I codons is observed in S. salar in pathogen-rich environments, whereas diversifying selection for MHC I codons is linked to environmental stressors, such as temperature and pollution in O. niloticus [17,50,65]. Furthermore, variation in selective pressures was observed among clariid catfish populations. For instance, the wild population SNK2-CM-W showed strong positive selection and significant neutrality test results, while NST1-CM-W and NPT2-CM-C showed ω < 1, indicating purifying selection. These patterns may be associated with environmental and management differences between populations. Potential factors include variation in pathogen exposure, water quality, and hatchery conditions (e.g., selective breeding and stocking density), although direct evidence for these effects is not yet available. Subsequently, using the MEME method, 39 of the 54 codons were identified as being under episodic diversifying selection, indicating that positive selection acted at different sites along various branches of the phylogeny. This proportion is unusually high compared with reports in other teleosts, where only a smaller subset of codons shows such signals, bringing attention to the strong and widespread selective pressures acting on clariid catfish MHC [50,65]. This pattern supports the occurrence of lineage-specific adaptations driven by population-specific immune challenges [45]. Additionally, strong selection signals were observed at codons 10, 12, 48, and 49 within the α1 domain of the MHC I molecule, which is involved in antigen recognition. Secondary structure prediction revealed that this domain is mainly composed of β-strands, α-helices, and coil regions, indicating a well-organized and stable structure typical of antigen-binding areas (Figure S3). Using FEL and FUBAR, which detect pervasive selection, codons 48 and 49 were identified, whereas only codons 10 and 12 were identified by MEME under episodic diversifying selection. This is consistent with the findings in teleosts, where MEME typically detects more positively selected sites than FEL and FUBAR [50,65]. These results, which reflect method-specific sensitivity, showed that MEME detects selection acting on a subset of branches (episodic), whereas FEL and FUBAR detect selection acting across the entire phylogeny (pervasive). These codons, which are present in multiple alleles such as Clarias_MHCI*TH12, Clarias_MHCI*TH34, Clarias_MHCI*TH48, and Clarias_MHCI*TH84, with nonsynonymous substitutions, are considered potential candidates for further functional studies [66,67]. Codons 10 and 12 are positioned at β-strand regions, while codons 48 and 49 are located within the α-helix of the antigen-binding α1 domain, suggesting potential influence on peptide-binding interactions [68].

4.3. Advances and Future Directions in MHC Genotyping

MHC I in clariid catfish, which is shaped by heterogeneous and largely episodic selection, particularly in hatcheries or environmentally unstable settings, reflects the influence of various selective pressures [17,19,31,69]. However, our findings provide novel insight into the evolutionary dynamics of MHC I in clariid catfish, which also suggest that high diversity and population-specific alleles reflect immune adaptation to local pathogens, supporting disease-resistant breeding strategies [19,24]. Functional markers, such as MHC I are used to enable more accurate assessments of immune potential than neutral markers in disease-prone high-density aquaculture systems [22]. A key application involves integrating beneficial MHC alleles into selective breeding programs to enhance disease resistance while maintaining genetic diversity. However, overemphasis on only a few alleles could inadvertently narrow overall diversity, increasing inbreeding risks and reducing long-term adaptability of farmed populations. Therefore, conservation of broad allele pools alongside targeted selection is essential [70]. Populations with unique alleles or strong positively selection signals (ω = 10.000), like SNK2-CM-W, may be prioritized as “immune variability reservoirs” [71]. SNK1-CM-W and SNK2-CM-W, which exhibited high allelic richness and private alleles, are valued for their immunogenetic diversity and are considered promising for broodstock conservation. Despite the advances in sequencing and selection models, this study focused on short MHC I segments. Consequently, further research, including expression assays and immune challenge experiments, is required to clarify the functional roles of these novel alleles [16,72]. Future studies could also apply RNA-seq to quantify allele-specific expression patterns under pathogen stress, providing direct evidence of functional relevance. For instance, long-read sequencing, such as PacBio or Nanopore, can be used to resolve full MHC loci, revealing gene copy variations and regulatory complexities that are often missed by short amplicon approaches [33]. Nevertheless, this study enhances our understanding of MHC class I evolution in clariid catfish and provides a genetic basis for disease-resistant breeding and conservation, thereby supporting the integration of MHC data into broodstock management for sustainable aquaculture.

5. Conclusions

To the best of our knowledge, this study provides the first comprehensive analysis of MHC I-UAA gene variation in three important clariid catfish species in Thailand using targeted next-generation sequencing. We identified 91 novel alleles across wild and hatchery populations, revealing substantial diversity in exon 6, which encodes the antigen-binding α1 domain. Purifying and diversifying selections, with population- and site-specific signals of adaptive evolution, were detected by molecular analyses, highlighting dynamic immune gene responses to environmental and pathogenic pressures. These findings support the use of population-specific alleles as MAS targets to improve disease resistance. Populations such as SNK2-CM-W and SNK1-CM-W, which exhibited high allelic richness and unique variants, are considered promising sources for marker-assisted selection in breeding programs. These genetic patterns provide a basis for defining populations that can be used for genetic monitoring and management. Future studies, including functional validation and full-length MHC genotyping through long-read sequencing, are expected to further enhance our understanding of adaptive immune variation and support the development of more effective breeding and conservation strategies for tropical freshwater fishes. In particular, disease-challenge experiments will be essential to validate whether specific alleles provide measurable resistance when exposed to pathogens.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/genes16091106/s1, Figure S1: Distribution of sampling sites in Thailand; Figure S2: MHC I protein sequence alignment of the catfish populations in Thailand and reference sequences; Figure S3: MHC I protein secondary structure production in catfish species in Thailand; Table S1: Summary of catfish individuals sampled in this study; Table S2: The catfish specimens used in this study; Table S3: Variable sites in the sequences of MHC I alleles of North African and bighead catfish populations in Thailand; Table S4:The results of Analysis of Variance (AMOVA) for catfish populations in Thailand; Table S5: Detailed site-by-site results from the MEME analysis; Table S6: Detailed site-by-site results from the FEL analysis; Table S7: Codon sites under selection identified by FUBAR analysis; Table S8: Types of single nucleotide polymorphism (SNPs) and their locations in the partial fragment of the exon 6 of MHC I gene of catfish catfish populations in Thailand compare with the reference sequence (Accession number XM53477511).

Author Contributions

T.H.D.N.: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Visualization, Writing—original draft, Writing—review and editing. P.C.: Data curation, Formal analysis, Methodology, Writing—review and editing. C.P.: Data curation, Formal analysis, Methodology, Writing—review and editing. W.J.: Data curation, Formal analysis, Methodology, Writing—review and editing. W.S.: Conceptualization, Data curation, Formal analysis, Visualization, Writing—review and editing. T.P.: Formal analysis, Visualization, Writing—review and editing. T.B.: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Visualization, Writing—original draft, Writing—review and editing. K.S. (Kednapat Sriphairoj): Resource, Methodology, Writing—review and editing. S.H.: Resource, Methodology, Writing—review and editing. P.S.: Resource, Methodology, Writing—review and editing. N.M.: Formal analysis, Methodology, Writing—review and editing. D.K.G.: Methodology, Investigation, Writing—review and editing. A.A.: Investigation, Methodology, Writing—review and editing. P.D.: Formal analysis, Methodology, Writing—review and editing. K.S. (Kornsorn Srikulnath): Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Visualization, Writing—original draft, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received funding support from The Program Management Unit for Human Resources and Institutional Development and Innovation (PMU-B) under the Program of National Postdoctoral and Postgraduate System (Contract No. B13F670053), which was awarded to W.S., T.P., N.M., P.D., and K.S. (Kornsorn Srikulnath); National Research Council of Thailand (NRCT): High-Potential Research Team Grant Program (contract number: N42A660605) awarded to W.S., T.P., K.S. (Kednapat Sriphairoj), S.H., N.M., P.D., and K.S. (Kornsorn Srikulnath); National Research Council of Thailand (NRCT) (contract number: N42A650233) awarded to K.S. (Kornsorn Srikulnath); a grants from Betagro Group (no. 6501.0901.1/68) awarded to K.S. (Kornsorn Srikulnath); a Postdoctoral Fellowship from Kasetsart University awarded to T.P.; a grant from Kasetsart University Research and Development Institute (FF(KU)25.64 and FF(S-KU)17.66) awarded to W.S. and K.S. (Kornsorn Srikulnath); the Visiting Research Scholar (VRC) Grant, Faculty of Science, Kasetsart University (Contract Number VRC 4/2024), awarded to D.K.G. and K.S. (Kornsorn Srikulnath); The Office of the Ministry of Higher Education, Science, Research and Innovation; and the Thailand Science Research and Innovation (TSRI) through the Kasetsart University Reinventing University Program 2025 (No. RUP_Climate68_No. 9) awarded to W.S. and A.A.; a support from the International SciKU Branding (ISB), Faculty of Science, Kasetsart University awarded to W.S. and K.S. (Kornsorn Srikulnath); and the Thailand Scholarship (Scholarship number: MHESI 0202.3/10651) awarded to T.H.D.N. and K.S. (Kornsorn Srikulnath). No funding source was involved in the study design, collection, analysis, and interpretation of the data, writing of the report, or decision to submit the article for publication.

Institutional Review Board Statement

All experimental procedures, including animal care, were reviewed and approved by the Animal Experiment Committee of Kasetsart University (approval no. ACKU65-SCI-003, ACKU66-SCI-006, and ACKU66-SCI-014) were conducted in accordance with the Regulations on Animal Experiments of Kasetsart University ARRIVE (https://arriveguidelines.org/, accessed on 8 September 2025).

Data Availability Statement

All sequences were deposited in the National Center for Biotechnology Information (NCBI) (https://www.ncbi.nlm.nih.gov/; accession number: PX243432–PX243522 (accessed on 8 September 2025)).

Acknowledgments

The authors would like to thank Phu Sing Research and Training Center at Kalasin University and Betagro Fish Breeding Farm for helping us with sample collection. The authors thank the Center for Agricultural Biotechnology (CAB) at Kasetsart University, Kamphaeng Saen Campus and the National Science and Technology Development Agency (NSTDA) Supercomputer Center (ThaiSC) for supporting us with server analysis services. The authors also thank the Faculty of Science of Kasetsart University (no. 6501.0901.1/336), the Faculty of Biology Education, School of Education, Can Tho University, and the Betagro Public Company Limited for providing research facilities.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AMOVAAnalysis of Molecular Variance
BICBayesian Information Criterion
ESUEvolutionarily Significant Unit
FELFixed Effects Likelihood
FUBARFast Unconstrained Bayesian Approximation
MCMCMarkov Chain Monte Carlo
MHC IMajor Histocompatibility Complex class I
NCBINational Center for Biotechnology Information
PCRPolymerase Chain Reaction

References

  1. FAO. Top 10 Species Groups in Global Aquaculture; Food and Agriculture Organization of The United Nations: Rome, Italy, 2023. [Google Scholar]
  2. Garcia de Leaniz, C.; Fleming, I.A.; Einum, S.; Verspoor, E.; Jordan, W.C.; Consuegra, S.; Aubin-Horth, N.; Lajus, D.; Letcher, B.H.; Youngson, A.F.; et al. A critical review of adaptive genetic variation in Atlantic salmon: Implications for conservation. Biol. Rev. 2007, 82, 173–211. [Google Scholar] [CrossRef]
  3. Neff, B.D.; Garner, S.R.; Pitcher, T.E. Conservation and enhancement of wild fish populations: Preserving genetic quality versus genetic diversity. Can. J. Fish. Aquat. Sci. 2011, 68, 1139–1154. [Google Scholar] [CrossRef]
  4. Wachirachaikarn, A.; Rungsin, W.; Srisapoome, P.; Na-Nakorn, U. Crossing of African catfish, Clarias gariepinus (Burchell, 1822), strains based on strain selection using genetic diversity data. Aquaculture 2009, 290, 53–60. [Google Scholar] [CrossRef]
  5. Gjedrem, T.; Robinson, N.; Rye, M. The importance of selective breeding in aquaculture to meet future demands for animal protein: A review. Aquaculture 2012, 350, 117–129. [Google Scholar] [CrossRef]
  6. Sonesson, A.K.; Hallerman, E.; Humphries, F.; Hilsdorf, A.W.S.; Leskien, D.; Rosendal, K.; Bartley, D.; Hu, X.; Gomez, R.G.; Mair, G.C. Sustainable management and improvement of genetic resources for aquaculture. J. World Aquac. Soc. 2023, 54, 364–396. [Google Scholar] [CrossRef]
  7. Bijlsma, R.; Loeschcke, V. Environmental stress, adaptation and evolution: An overview. J. Evol. Biol. 2005, 18, 744–749. [Google Scholar] [CrossRef]
  8. Hoffmann, A.A.; Sgrò, C.M. Climate change and evolutionary adaptation. Nature 2011, 470, 479–485. [Google Scholar] [CrossRef]
  9. Patta, C.; Sriphairoj, K.; Budi, T.; Quanoo, D.K.; Nguyen, T.H.D.; Jaito, W.; Chalermwong, P.; Pongsanarm, T.; Thatukan, C.; Wongloet, W.; et al. Genetic Structural Changes While Maintaining Effective Population Size of Bighead Catfish in Nong Han Lake: Implications of Metapopulation Dynamics or Release Activities. Fish. Manag. Ecol. 2024, 32, e12778. [Google Scholar] [CrossRef]
  10. Department of Fisheries. Marine Fisheries Management Plan of Thailand 2020–2022; Ministry of Agriculture and Cooperatives, Thailand: Bangkok, Thailand, 2023.
  11. Srikulnath, K.; Budi, T.; Panthum, T.; Sinthuvanich, C.; Haileselasie, T.H.; Kuldilok, K.; Photchanaprasert, N.; Sayruamyat, S.; Muangmai, N.; Duengkae, P.; et al. Sustainable Aquaculture in Thailand: Balancing Economic Growth and Ecological Integrity with North African Catfish Integration. Aquac. Res. 2025, in press. [CrossRef]
  12. Sciuto, S.; Colli, L.; Fabris, A.; Pastorino, P.; Stoppani, N.; Esposito, G.; Prearo, M.; Esposito, G.; Ajmone-Marsan, P.; Acutis, P.L.; et al. What can genetics do for the control of infectious diseases in aquaculture? Animals 2022, 12, 2176. [Google Scholar] [CrossRef]
  13. Wang, B.; Du, H.-h.; Huang, H.-q.; Xian, J.-a.; Xia, Z.-h.; Hu, Y.-h. Major histocompatibility complex class I (MHC Iα) of Japanese flounder (Paralichthys olivaceus) plays a critical role in defense against intracellular pathogen infection. Fish Shellfish. Immunol. 2019, 94, 122–131. [Google Scholar] [CrossRef]
  14. Olson, E.; Raghavan, M. Major histocompatibility complex class I assembly within endolysosomal pathways. Curr. Opin. Immunol. 2023, 84, 102356. [Google Scholar] [CrossRef]
  15. Janeway, C.A.; Travers, P.; Walport, M.; Shlomchik, M.J. The major histocompatibility complex and its functions. In Immunobiology: The Immune System in Health and Disease, 5th ed.; Garland Science: New York, NY, USA, 2001. [Google Scholar]
  16. Grimholt, U. MHC and Evolution in Teleosts. Biology 2016, 5, 6. [Google Scholar] [CrossRef]
  17. Piertney, S.B.; Oliver, M.K. The evolutionary ecology of the major histocompatibility complex. Heredity 2006, 96, 7–21. [Google Scholar] [CrossRef]
  18. Yamaguchi, T.; Dijkstra, J.M. Major histocompatibility complex (MHC) genes and disease resistance in fish. Cells 2019, 8, 378. [Google Scholar] [CrossRef] [PubMed]
  19. Spurgin, L.G.; Richardson, D.S. How pathogens drive genetic diversity: MHC, mechanisms and misunderstandings. Proc. R. Soc. B Biol. Sci. 2010, 277, 979–988. [Google Scholar] [CrossRef] [PubMed]
  20. Azis; Alimuddin; Sukenda; Zairin, M. MHC I molecular marker inheritance and first generation catfish (Clarias sp.) resistance against Aeromonas hydrophila infection. Pak. J. Biotechnol. 2015, 12, 131–137. [Google Scholar]
  21. Oyebola, O.O.; Babaoye, E.O.; Anifowose, O.R. Growth performance, survival rate and expression of MHC I gene in Clarias gariepinus at post Vagococcus carniphilus challenge period. Ecol. Genet. Genom. 2025, 34, 100319. [Google Scholar] [CrossRef]
  22. Castro-Prieto, A.; Wachter, B.; Sommer, S. Cheetah paradigm revisited: MHC diversity in the world’s largest free-ranging population. Mol. Biol. Evol. 2011, 28, 1455–1468. [Google Scholar] [CrossRef]
  23. Smallbone, W.; Ellison, A.; Poulton, S.; van Oosterhout, C.; Cable, J. Depletion of MHC supertype during domestication can compromise immunocompetence. Mol. Ecol. 2021, 30, 736–746. [Google Scholar] [CrossRef]
  24. Sommer, S. The importance of immune gene variability (MHC) in evolutionary ecology and conservation. Front. Zool. 2005, 2, 16. [Google Scholar] [CrossRef] [PubMed]
  25. O’Farrell, B.; Benzie, J.A.H.; McGinnity, P.; de Eyto, E.; Dillane, E.; Coughlan, J.; Cross, T.F.; Tramontano, A. Selection and phylogenetics of salmonid MHC class I: Wild brown trout (Salmo trutta) differ from a non-native introduced strain. PLoS ONE 2013, 8, e63035. [Google Scholar] [CrossRef] [PubMed]
  26. Schenekar, T.; Weiss, S. Selection and genetic drift in captive versus wild populations: An assessment of neutral and adaptive (MHC-linked) genetic variation in wild and hatchery brown trout (Salmo trutta) populations. Conserv. Genet. 2017, 18, 1011–1022. [Google Scholar] [CrossRef]
  27. Stet, R.M.; van Erp, S.H.; Hermsen, T.; Sültmann, H.A.; Egberts, E. Polymorphism and estimation of the number of MhcCyca class I and class I genes in laboratory strains of the common carp (Cyprinus carpio L.). Dev. Comp. Immunol. 1993, 17, 141–156. [Google Scholar] [CrossRef] [PubMed]
  28. Yáñez, J.M.; Houston, R.D.; Newman, S. Genetics and genomics of disease resistance in salmonid species. Front. Genet. 2014, 5, 415. [Google Scholar] [CrossRef]
  29. Liu, W.; Liu, Y.; Liu, P.; Zhao, W. Genetic diversity of major histocompatibility complex class I genes in Zootoca vivipara. Biosci. Rep. 2020, 40, BSR20193809. [Google Scholar] [CrossRef]
  30. Supikamolseni, A.; Ngaoburanawit, N.; Sumontha, M.; Chanhome, L.; Suntrarachun, S.; Peyachoknagul, S.; Srikulnath, K. Molecular barcoding of venomous snakes and species-specific multiplex PCR assay to identify snake groups for which antivenom is available in Thailand. Genet. Mol. Res. 2015, 14, 13981–13997. [Google Scholar] [CrossRef]
  31. Imron, I.; Marnis, H.; Iswanto, B.; Suprapto, R. Development of a PCR marker for the identification of resistance to Motile Aeromonad Septicemia disease in African catfish (Clarias gariepinus). Aquac. Aquar. Conserv. Legis. 2020, 13, 1255–1267. [Google Scholar]
  32. Simon, A. FastQC: A Quality Control Tool for High Throughput Sequence Data, Version 0.10; Babraham Bioinformatics: Cambridge, UK, 2010.
  33. Sebastian, A.; Herdegen, M.; Migalska, M.; Radwan, J. amplisas: A web server for multilocus genotyping using next-generation amplicon sequencing data. Mol. Ecol. Resour. 2016, 16, 498–510. [Google Scholar] [CrossRef]
  34. Lighten, J.; Van Oosterhout, C.; Bentzen, P. Critical review of NGS analyses for de novo genotyping multigene families. Mol. Ecol. 2014, 23, 3957–3972. [Google Scholar] [CrossRef]
  35. Antao, A.B.; Chinchar, V.G.; McConnell, T.J.; Miller, N.W.; Clem, L.W.; Wilson, M.R. MHC class I genes of the channel catfish: Sequence analysis and expression. Immunogenetics 1999, 49, 303–311. [Google Scholar] [CrossRef]
  36. FBN. BioProject: Clarias Gariepinus RefSeq Genome Sequencing and Assembly; National Center for Biotechnology Information: Bethesda, MD, USA, 2022. [Google Scholar]
  37. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef]
  38. R Core Team. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2023; Available online: https://www.R-project.org/ (accessed on 5 May 2025).
  39. Kamvar, Z.N.; Tabima, J.F.; Grünwald, N.J. Poppr: An R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ 2014, 2014, e281. [Google Scholar] [CrossRef]
  40. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.; Von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef]
  41. Ronquist, F.; Teslenko, M.; Van Der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef]
  42. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef]
  43. Nei, M.; Gojobori, T. Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions. Mol. Biol. Evol. 1986, 3, 418–426. [Google Scholar] [CrossRef] [PubMed]
  44. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  45. Murrell, B.; O Wertheim, J.; Moola, S.; Weighill, T.; Scheffler, K.; Pond, S.L.K. Detecting individual sites subject to episodic diversifying selection. PLoS Genet. 2012, 8, e1002764. [Google Scholar] [CrossRef]
  46. Kosakovsky Pond, S.L.; Frost, S.D. Not so different after all: A comparison of methods for detecting amino acid sites under selection. Mol. Biol. Evol. 2005, 22, 1208–1222. [Google Scholar] [CrossRef] [PubMed]
  47. Murrell, B.; Moola, S.; Mabona, A.; Weighill, T.; Sheward, D.; Pond, S.L.K.; Scheffler, K. FUBAR: A fast, unconstrained bayesian approximation for inferring selection. Mol. Biol. Evol. 2013, 30, 1196–1205. [Google Scholar] [CrossRef]
  48. Pearson, W.R. An introduction to sequence similarity (“homology”) searching. Curr. Protoc. Bioinform. 2013, 42, 3.1.1–3.1.8. [Google Scholar] [CrossRef] [PubMed]
  49. Meiklejohn, K.A.; Damaso, N.; Robertson, J.M.; Fugmann, S.D. Assessment of BOLD and GenBank–Their accuracy and reliability for the identification of biological materials. PLoS ONE 2019, 14, e0217084. [Google Scholar] [CrossRef]
  50. Sato, A.; Dongak, R.; Hao, L.; Takezaki, N.; Shintani, S.; Aoki, T.; Klein, J. MHC class I genes of the cichlid fish Oreochromis niloticus. Immunogenetics 2006, 58, 917–928. [Google Scholar] [CrossRef]
  51. Zhou, T.; Yuan, Z.; Tan, S.; Jin, Y.; Yang, Y.; Shi, H.; Wang, W.; Niu, D.; Gao, L.; Jiang, W.; et al. A review of molecular responses of catfish to bacterial diseases and abiotic stresses. Front. Physiol. 2018, 9, 1113. [Google Scholar] [CrossRef] [PubMed]
  52. Segaran, T.C.; Azra, M.N.; Piah, R.M.; Lananan, F.; Téllez-Isaías, G.; Gao, H.; Torsabo, D.; Kari, Z.A.; Noordin, N.M. Catfishes: A global review of the literature. Heliyon 2023, 9, e20081. [Google Scholar] [CrossRef]
  53. Zhou, X.; Tang, J.; Wang, S.; Zhang, Y.; Ye, H.; Zhang, Q.; Xiang, W.; Cai, T.; Zeng, C. Whole genome sequencing and transcriptomics-based characterization of a novel β-cypermethrin-degrading Gordonia alkanivorans GH-1 isolated from fermented foods. Chemosphere 2023, 320, 138017. [Google Scholar] [CrossRef]
  54. De Alwis, P.S.; Kundu, S.; Gietbong, F.Z.; Amin, M.H.F.; Lee, S.-R.; Kim, H.-W.; Kim, A.R. Mitochondriomics of Clarias fishes (Siluriformes: Clariidae) with a new assembly of Clarias camerunensis: Insights into the genetic characterization and diversification. Life 2023, 13, 482. [Google Scholar] [CrossRef]
  55. Klein, J.; Sato, A.; Nagl, S.; O’hUigín, C. Molecular trans-species polymorphism. Annu. Rev. Ecol. Syst. 1998, 29, 1–21. [Google Scholar] [CrossRef]
  56. Hedrick, P.W. Adaptive introgression in animals: Examples and comparison to new mutation and standing variation as sources of adaptive variation. Mol. Ecol. 2013, 22, 4606–4618. [Google Scholar] [CrossRef] [PubMed]
  57. Gómez, D.; Conejeros, P.; Consuegra, S.; Marshall, S.H. MHC mediated resistance to Piscirickettsia salmonis in salmonids farmed in Chile. Aquaculture 2011, 318, 15–19. [Google Scholar] [CrossRef]
  58. Million, K.M.; Lively, C.M. Trans-specific polymorphism and the convergent evolution of supertypes in major histocompatibility complex class II genes in darters (Etheostoma). Ecol. Evol. 2022, 12, e8485. [Google Scholar] [CrossRef]
  59. Charlesworth, B.; Jensen, J.D. Effects of selection at linked sites on patterns of genetic variability. Annu. Rev. Ecol. Evol. Syst. 2021, 52, 177–197. [Google Scholar] [CrossRef]
  60. Scheiblhofer, S.; Laimer, J.; Machado, Y.; Weiss, R.; Thalhamer, J. Influence of protein fold stability on immunogenicity and its implications for vaccine design. Expert Rev. Vaccines 2017, 16, 479–489. [Google Scholar] [CrossRef] [PubMed]
  61. Domnick, A.; Winter, C.; Sušac, L.; Hennecke, L.; Hensen, M.; Zitzmann, N.; Trowitzsch, S.; Thomas, C.; Tampé, R. Molecular basis of MHC I quality control in the peptide loading complex. Nat. Commun. 2022, 13, 4701. [Google Scholar] [CrossRef] [PubMed]
  62. Alimuddin, A.; Putri, F.M.; Wahjuningrum, D.; Hardiantho, D.; Sunarma, A.; Nuryati, S. Resistance against Aeromonas hydrophila infection and growth of second generation (F2) African catfish (Clarias gariepinus) using selected molecular markers. Biotropia 2018, 25, 95–102. [Google Scholar] [CrossRef]
  63. Lenz, T.L. Adaptive value of novel MHC immune gene variants. Proc. Natl. Acad. Sci. USA 2018, 115, 1414–1416. [Google Scholar] [CrossRef]
  64. López, M.E.; Benestan, L.; Moore, J.; Perrier, C.; Gilbey, J.; Di Genova, A.; Maass, A.; Diaz, D.; Lhorente, J.; Correa, K.; et al. Comparing genomic signatures of domestication in two Atlantic salmon (Salmo salar L.) populations with different geographical origins. Evol. Appl. 2019, 12, 137–156. [Google Scholar] [CrossRef] [PubMed]
  65. Dionne, M.; Miller, K.M.; Dodson, J.J.; Bernatchez, L. MHC standing genetic variation and pathogen resistance in wild Atlantic salmon. Philos. Trans. R. Soc. B Biol. Sci. 2009, 364, 1555–1565. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, J.; Sun, Y.; Xu, T. Identification of 48 full-length MHC-DAB functional alleles in miiuy croaker and evidence for positive selection. Fish Shellfish. Immunol. 2016, 54, 544–550. [Google Scholar] [CrossRef]
  67. Wegner, K. Historical and contemporary selection of teleost MHC genes: Did we leave the past behind? J. Fish Biol. 2008, 73, 2110–2132. [Google Scholar] [CrossRef]
  68. Loh, Z.; Huan, X.; Awate, S.; Schrittwieser, M.; Renia, L.; Ren, E.C. Molecular characterization of MHC class I alpha 1 and 2 domains in Asian seabass (Lates calcarifer). Int. J. Mol. Sci. 2022, 23, 10688. [Google Scholar] [CrossRef] [PubMed]
  69. Larson, W.A.; Seeb, J.E.; Dann, T.H.; Schindler, D.E.; Seeb, L.W. Signals of heterogeneous selection at an MHC locus in geographically proximate ecotypes of sockeye salmon. Mol. Ecol. 2014, 23, 5448–5461. [Google Scholar] [CrossRef] [PubMed]
  70. D’ambrosio, J.; Phocas, F.; Haffray, P.; Bestin, A.; Brard-Fudulea, S.; Poncet, C.; Quillet, E.; Dechamp, N.; Fraslin, C.; Charles, M.; et al. Genome-wide estimates of genetic diversity, inbreeding and effective size of experimental and commercial rainbow trout lines undergoing selective breeding. Genet. Sel. Evol. 2019, 51, 26. [Google Scholar] [CrossRef]
  71. Hedrick, P.W. Pathogen resistance and genetic variation at MHC loci. Evolution 2002, 56, 1902–1908. [Google Scholar] [CrossRef]
  72. Evans, K.M.; Wortley, A.H.; Mann, D.G. An assessment of potential diatom “barcode” genes (cox1, rbcL, 18S and ITS rDNA) and their effectiveness in determining relationships in Sellaphora (Bacillariophyta). Protist 2007, 158, 349–364. [Google Scholar] [CrossRef]
Figure 1. Bayesian phylogenetic tree based on MHC I gene in catfish species. Values above branches represent posterior probability. The scale bar indicates the number of substitutions per site. Alleles do not form species-specific clusters, reflecting allele sharing among C. gariepinus, C. macrocephalus, and C. batrachus. Reference sequences from T. fulvidraco was included as outgroups.
Figure 1. Bayesian phylogenetic tree based on MHC I gene in catfish species. Values above branches represent posterior probability. The scale bar indicates the number of substitutions per site. Alleles do not form species-specific clusters, reflecting allele sharing among C. gariepinus, C. macrocephalus, and C. batrachus. Reference sequences from T. fulvidraco was included as outgroups.
Genes 16 01106 g001
Figure 2. Bayesian phylogenetic tree based on MHC I gene amino acid residues in catfish species. Values above branches represent posterior probability. The scale bar indicates the number of substitutions per site. Alleles do not form species-specific clusters, reflecting allele sharing among C. gariepinus, C. macrocephalus, and C. batrachus. Reference sequences from T. fulvidraco was included as outgroups.
Figure 2. Bayesian phylogenetic tree based on MHC I gene amino acid residues in catfish species. Values above branches represent posterior probability. The scale bar indicates the number of substitutions per site. Alleles do not form species-specific clusters, reflecting allele sharing among C. gariepinus, C. macrocephalus, and C. batrachus. Reference sequences from T. fulvidraco was included as outgroups.
Genes 16 01106 g002
Table 1. Nucleotide sequence diversity in catfish populations based on MHC I sequences.
Table 1. Nucleotide sequence diversity in catfish populations based on MHC I sequences.
SpeciesPopulationCodeNNumber of Allele per PopulationNucleotide Diversity
C. gariepinusNakhon NayokNYK-CG-C 530.074
Kalasin 1KSN1-CG-C94200.111
Kalasin 2KSN2-CG-C134370.082
Ubon RatchathaniUBR-CG-C630.091
Sing BuriSB-CG-C770.118
C. gariepinus-246410.095
SD---0.017
C. macrocephalusSing BuriSB-CM-C440.240
Sakon Nakhon 1SNK1-CM-W182380.133
Sakon Nakhon 2SNK2-CM-W74340.121
Sakon Nakhon 3SNK3-CM-W82160.104
Sakon Nakhon 4SNK4-CM-C1450.132
Suphan Buri 1SPB1-CM-W6110.123
Suphan Buri 2SPB2-CM-W3100.127
Nakhon Pathom 1NPT1-CM-W280.131
Nakhon Pathom 2NPT2-CM-W270.144
Nakhon Si Thammarat 1NST1-CM-W360.145
Nakhon Si Thammarat 2NST2-CM-C10130.136
Surat ThaniSTN-CM-C25260.114
Nakhon PhanomNKPN-CM-C740.084
Ubon RatchathaniUBR-CM-C6100.128
C. macrocephalus-420590.133
SD---0.033
C. batrachusUbon RatchathaniUBR-CB-C860.087
Overall mean value-674910.121
SD---0.034
SD: standard deviation.
Table 2. Neutrality tests of MHC I gene sequences in catfish populations.
Table 2. Neutrality tests of MHC I gene sequences in catfish populations.
SpeciesPopulationCodeTajima’s DFu and Li’s DFu and Li’s F
C. gariepinusNakhon NayokNYK-CG-C 0.604 ns0.750 ns0.792 ns
Kalasin 1KSN1-CG-C0.228 ns1.356 ns1.001 ns
Kalasin 2KSN2-CG-C−0.699 ns0.045 ns−0.388 ns
Ubon RatchathaniUBR-CG-C−0.507 ns−0.683 ns−0.716 ns
Sing BuriSB-CG-C−0.869 ns−0.957 ns−1.064 ns
Mean-−0.2490.102−0.075
C. macrocephalusSing BuriSB-CM-C−1.292 ns−1.188 ns−1.330 ns
Sakon Nakhon 1SNK1-CM-W−0.936 ns−3.082 *−2.392 *
Sakon Nakhon 2SNK2-CM-W−1.142 ns−2.370 *−2.114 ns
Sakon Nakhon 3SNK3-CM-W−0.032 ns0.336 ns0.206 ns
Sakon Nakhon 4SNK4-CM-C−1.700 ns−2.704 *−2.800 *
Suphan Buri 1SPB1-CM-W−0.721 ns0.012 ns−0.247 ns
Suphan Buri 2SPB2-CM-W−0.521 ns0.607 ns0.326 ns
Nakhon Pathom 1NPT1-CM-W−0.614 ns0.112 ns−0.085 ns
Nakhon Pathom 2NPT2-CM-W−0.800 ns−0.347 ns−0.505 ns
Nakhon Si Thammarat 1NST1-CM-W−0.085 ns0.613 ns0.495 ns
Nakhon Si Thammarat 2NST2-CM-C−0.542 ns−0.727 ns−0.786 ns
Surat ThaniSTN-CM-C−0.935 ns1.288 ns0.404 ns
Nakhon PhanomNKPN-CM-C−0.879 ns−0.906 ns−1.010 ns
Ubon RatchathaniUBR-CM-C−1.478 ns−1.495 ns−1.713 ns
Mean-−0.834−0.704−0.825
C. batrachusUbon RatchathaniUBR-CB-C−0.670 ns−1.453 ns−1.422 ns
Overall mean value-−0.679−0.540−0.667
ns: not significant; *: p < 0.05.
Table 3. Rates of synonymous (dS) and nonsynonymous (dN) substitutions in nucleotide sequences of MHC I gene sequences in catfish populations.
Table 3. Rates of synonymous (dS) and nonsynonymous (dN) substitutions in nucleotide sequences of MHC I gene sequences in catfish populations.
SpeciesPopulationCodedNdSω
C. gariepinusNakhon NayokNYK-CG-C 0.049 ± 0.0180.023 ± 0.0162.130
Kalasin 1KSN1-CG-C0.094 ± 0.0200.037 ± 0.0172.541
Kalasin 2KSN2-CG-C0.064 ± 0.0120.020 ± 0.0083.200
Ubon RatchathaniUBR-CG-C0.035 ± 0.0130.011 ± 0.0083.182
Sing BuriSB-CG-C0.077 ± 0.0210.044 ± 0.0191.750
Mean-0.064 ± 0.0170.027 ± 0.0142.561
C. macrocephalusSing BuriSB-CM-C0.502 ± 0.1300.141 ± 0.0493.560
Sakon Nakhon 1SNK1-CM-W0.120 ± 0.0230.090 ± 0.0331.333
Sakon Nakhon 2SNK2-CM-W0.120 ± 0.0210.012 ± 0.03710.000
Sakon Nakhon 3SNK3-CM-W0.102 ± 0.0230.084 ± 0.0371.214
Sakon Nakhon 4SNK4-CM-C0.200 ± 0.0430.063 ± 0.0213.175
Suphan Buri 1SPB1-CM-W0.088 ± 0.0190.097 ± 0.0400.907
Suphan Buri 2SPB2-CM-W0.102 ± 0.0210.098 ± 0.0431.041
Nakhon Pathom 1NPT1-CM-W0.050 ± 0.0120.034 ± 0.0151.471
Nakhon Pathom 2NPT2-CM-W0.066 ± 0.0180.093 ± 0.0460.710
Nakhon Si Thammarat 1NST1-CM-W0.059 ± 0.0180.087 ± 0.0410.678
Nakhon Si Thammarat 2NST2-CM-C0.070 ± 0.0160.080 ± 0.0370.875
Surat ThaniSTN-CM-C0.122 ± 0.0220.117 ± 0.0431.043
Nakhon PhanomNKPN-CM-C0.062 ± 0.0170.052 ± 0.0271.192
Ubon RatchathaniUBR-CM-C0.139 ± 0.0220.132 ± 0.0491.053
Mean-0.129 ± 0.0290.084 ± 0.0372.018
C. batrachusUbon RatchathaniUBR-CB-C0.107 ± 0.0290.097 ± 0.0521.103
Overall mean value-0.111 ± 0.0970.071 ± 0.0391.563
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

Nguyen, T.H.D.; Chalermwong, P.; Patta, C.; Jaito, W.; Singchat, W.; Panthum, T.; Budi, T.; Sriphairoj, K.; Hatachote, S.; Srisapoome, P.; et al. Genetic Diversity and Selection of MHC I-UAA in Clariid Catfish from Thailand: Implications for Breeding and Conservation. Genes 2025, 16, 1106. https://doi.org/10.3390/genes16091106

AMA Style

Nguyen THD, Chalermwong P, Patta C, Jaito W, Singchat W, Panthum T, Budi T, Sriphairoj K, Hatachote S, Srisapoome P, et al. Genetic Diversity and Selection of MHC I-UAA in Clariid Catfish from Thailand: Implications for Breeding and Conservation. Genes. 2025; 16(9):1106. https://doi.org/10.3390/genes16091106

Chicago/Turabian Style

Nguyen, Ton Huu Duc, Piangjai Chalermwong, Chananya Patta, Wattanawan Jaito, Worapong Singchat, Thitipong Panthum, Trifan Budi, Kednapat Sriphairoj, Sittichai Hatachote, Prapansak Srisapoome, and et al. 2025. "Genetic Diversity and Selection of MHC I-UAA in Clariid Catfish from Thailand: Implications for Breeding and Conservation" Genes 16, no. 9: 1106. https://doi.org/10.3390/genes16091106

APA Style

Nguyen, T. H. D., Chalermwong, P., Patta, C., Jaito, W., Singchat, W., Panthum, T., Budi, T., Sriphairoj, K., Hatachote, S., Srisapoome, P., Muangmai, N., Griffin, D. K., Antunes, A., Duengkae, P., & Srikulnath, K. (2025). Genetic Diversity and Selection of MHC I-UAA in Clariid Catfish from Thailand: Implications for Breeding and Conservation. Genes, 16(9), 1106. https://doi.org/10.3390/genes16091106

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