Next Article in Journal
Occurrence and Characteristics of Cut Injuries in Hedgehogs in Germany: A Collection of Individual Cases
Next Article in Special Issue
Analysis of Genetic Diversity in Romanian Carpatina Goats Using SNP Genotyping Data
Previous Article in Journal
Exploring the Microbial Community Structure in the Chicken House Environment by Metagenomic Analysis
Previous Article in Special Issue
Investigating How Genetic Merit and Country of Origin Impact the Profitability of Grass-Based Sheep Production Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Variations in Casein Genes Are Associated with Milk Protein and Fat Contents in Sarda Goats (Capra hircus), with an Important Role of CSN1S2 for Milk Yield

1
Dipartimento Medicina Veterinaria, Università degli Studi di Sassari, 07100 Sassari, Italy
2
Centre de Recerca Agrigenòmica (CRAG), Campus Universitat Autònoma de Barcelona, Universitat Autònoma de Barcelona, 08193 Bellaterra, Spain
3
Dipartimento Medicina Veterinaria, Università degli Studi di Bari Aldo Moro, 70121 Bari, Italy
*
Author to whom correspondence should be addressed.
Animals 2024, 14(1), 56; https://doi.org/10.3390/ani14010056
Submission received: 3 October 2023 / Revised: 18 November 2023 / Accepted: 19 December 2023 / Published: 22 December 2023
(This article belongs to the Collection Genetic Diversity in Livestock and Companion Animals)

Abstract

:

Simple Summary

Extensive goat breeding follows principles of sustainable agriculture, and it can be achieved with local breeds because their genome has been shaped by centuries of interaction with the environment in which goats were raised. The Sarda goat is well adapted to the arid environment of Sardinia (Italy). The aim of the present work was to study the variability of the casein genes, the main proteins of milk. We measured milk traits such as protein, fat, lactose, total solids, and milk yield, and performed association analyses with casein gene variants. The variability of the four casein genes was associated with milk protein, fat, total solids, and milk energy. The main finding was that intronic variants of the CSN1S2 gene (encoding the αs2-casein protein) were associated with milk yield and protein and fat content. This information might be used in selection schemes, and in future investigations aiming to disclose the direct link between genotype and phenotype.

Abstract

This work aimed to assess the variability of casein genes in a population of 153 bucks and 825 lactating does of the Sarda breed, and to perform association analysis between polymorphic sites and milk yield and composition traits. To genotype the casein genes, we chose an SNP panel including 44 SNPs mapping to the four casein genes CSN1S1, CSN2, CSN1S2, and CSN3. Genotyping (made by KASP™ genotyping assay, based on competitive allele-specific PCR) revealed the high variability of the Sarda goat, and haplotype analysis revealed linkage disequilibrium (LD) between CSN1S1 and CSN2 genes, in addition to two LD blocks within the CSN1S2 and two LD blocks within the CSN3 gene, in bucks and does. Association analysis revealed that variability at all four casein genes was associated with milk protein content, total solids, and milk energy. The three Ca-sensitive casein genes were associated with lipid content, and CSN1S2 showed a unique pattern, with intron variants associated with milk yield, in addition to milk pH, NaCl, and SCS (Somatic Cell Score). This information might prove useful in selection schemes and in future investigations aiming to better understand the biology of lactation, and the direct link between genotype and phenotype.

1. Introduction

Dairy goats provide a reliable source of milk and other dairy products to communities around the world, as they can adapt to different environments, allowing the use of harsh territories that are otherwise unused or abandoned [1,2]. In Europe, the dairy goat sector has evolved into a thriving industry, with dedicated breeds reared under intensive farming systems, and local breeds reared under extensive systems, especially in mountainous areas, drylands, and islands [3]. While cosmopolitan breeds usually have higher milk production capacities when reared under high-input farming systems, the local breeds produce less milk, and they might have good milk quality and aptitude for coagulation [2]. With these premises, the dairy goat sector has the potential to respond to the growing demand for high-quality dairy products, and to meet the consumer preference for sustainable and locally sourced products; goat farming aligns well with the principles of sustainable agriculture [4].
Sardinia (Italy) is an island located in the center of the Mediterranean Sea, where dairy goat farming is traditionally run under extensive conditions, especially in the central, eastern, and south-western regions [5]. The estimated number of heads of Sarda goats, the local caprine breed reared in Sardinia, is 165,000, of which 25,429 heads are registered in the herdbook of the Sarda breed and make up the official population [6] (Figure 1). The Sarda goat is well adapted to the arid and inaccessible soils of Sardinia, displays high morphological and genetic variability, and its milk is mainly destined for cheese making. The milk yield was reported to be highly variable, about 200 L per lactation, and 0.95 Kg/d for Sarda goats [7].
Caseins make up about 80% of milk proteins, contribute to the nutritional value of milk, and have a functional and structural role, due to their involvement in the formation of casein micelles and milk transformation into cheese. Four different casein proteins are distinguished in goat milk: αs1-casein (encoded by the CSN1S1 gene), β-casein (encoded by the CSN2 gene), αs2-casein (encoded by the CSN1S2 gene), and k-casein (encoded by the CSN3 gene). The four casein genes form a cluster spanning 250 kbp on chromosome 6 [8]. Casein genes exhibit variability in their genomic DNA sequence, mRNA alternative splicing, transcript-level variation, and post-transcriptional modifications [9,10]. Polymorphisms in casein genes have been shown to affect milk protein and fat contents, milk yield, the organoleptic traits of milk and cheese, coagulation properties, and cheese yield [11]. Leroux et al. (1992) [11] revealed that the CSN1S1 gene encoding the as1-casein expressed nine different mRNAs, most lacking the three exons 9, 10, and 11. Jansa Perez et al. (1994) [12] demonstrated that the “intermediate” E variant was due to the insertion of a LINE sequence in the 3’UTR region of the CSN1S1 gene, which destabilized the mRNA, with consequent instability and reduced protein secretion. The molecular basis of casein variants was explained in detail, and numerous studies were conducted to evaluate allelic variants’ occurrence in the different breeds and populations and their association with milk traits. The E variant prevailed in the Saanen and Alpine breeds, followed by F, and the strong allele was CSN1S1 A [13]. In the Sarda, a local goat breed not subjected to extreme selection, the strong alleles of the CSN1S1 gene prevailed, with a high frequency of A and B alleles. The occurrence of the B allele associated with a higher expression level than A (both strong alleles) was also highlighted in Spanish goats [14].
In this context, the use of SNPs with a major effect on phenotype highlights the direct effect of the single variation. In contrast, neutral markers serve as reference points distributed in the region of interest, and allow a comparison to identify DNA regions that might be associated with a trait. In the present study, we used a set of neutral markers that ultimately allowed us to highlight the effect of as2-casein on milk yield, which had not been demonstrated so far in the literature. This work aimed to assess the variability of a set of DNA variants in casein genes in a population including bucks and does of Sarda breed, and to perform association analysis between polymorphic sites and milk yield and composition traits. To genotype the casein genes, we chose an SNP panel already tested in Murciano Granadina goats [15].

2. Materials and Methods

2.1. Animals and Sampling

The population analyzed included 825 lactating goats, sampled from 19 farms (about 40 goats per farm), and 153 bucks from 38 farms (1 to 5 per farm). The farms were located in Sardinia’s central, eastern, and south-western areas (Italy) (Figure 2).
Animals were officially registered in the flock books of the Sarda breed and enrolled in the milk recording system of provincial associations of goat breeders. A description of the extensive farming system adopted in Sardinia, and of the most recent historical events that influenced the current composition of the Sardinian goat population, has been given by Usai et al. (2006) [16]. At the same time, a more accurate description of the animals analyzed in this article is reported by Paschino et al. (2020) [5]. Individual milk samples (200 mL/goat) were collected from each female goat in sterile containers during the morning milking and immediately stored at 4 °C; one sampling day was completed on each farm. All the sampled goats were healthy, as certified by a veterinarian. When the milking system was mechanical, milk was taken from the recorder jar of each goat, and the same recorder jar was also used to measure milk yield (MY). When animals were hand-milked, milk was collected from the stainless-steel graduated pail, which was also used to record milk yield. For DNA extraction, a blood sample was taken from each goat in K3EDTA vacuum tubes (BD Vacutainer, Becton Dickinson, Franklin Lakes, NJ, USA).

2.2. Phenotyping

Milk samples were analyzed within 24 h after collection. Daily milk yield was calculated as the total record of morning plus evening milkings, measured on the same day of sample collection. Phenotyping included measurements of milk fat, protein, total solids, lactose, pH, and NaCl by using a MilkoScan FT6000 milk analyzer (Foss Electric A/S, Hillerød, Denmark), calibrated according to the following FIL-IDF references: ISO-IDF, 2013 [17], for fat, protein, lactose, pH, and NaCl; ISO-IDF, 2010, for total solids [18]. Milk energy was calculated as the sum of the values of the nutrients, according to the NRC (2001) [19] energy values: fat = 38.89 kJ/g; protein = 23.90 kJ/g; lactose = 16.53 kJ/g. Somatic cell count was evaluated using a Fossomatic 5000 somatic cell counter (Foss Electric) and then transformed into the logarithmic [log2(SCC × 10−5) + 3] SCS, according to Shook (1993) [20]. The total bacterial count was evaluated using a BactoScan FC150 analyzer (Foss Electric) and transformed into the logarithmic bacterial count [LBC = log10(total bacterial count/1000)] based on ISO-IDF (2004) [21].

2.3. Genotyping

Genomic DNA was obtained from leukocytes using the Gentra Puregene Blood Kit (Qiagen, Hilden, Germany), and purity and concentration were measured with an Eppendorf BioPhotometer (Eppendorf, Hamburg, Germany). Forty-four SNPs (Single Nucleotide Polymorphism) spanning the goat casein gene cluster were genotyped in all 978 genomic DNA samples, including 825 does and 153 bucks (Table 1). The 44-SNP panel was designed by Pizarro et al. (2020) [15]. Genotyping was performed with the KASP™ genotyping technology (Kompetitive Allele-Specific PCR assay, LGC Limited, Fordham, UK), and raw allele calls were analyzed with KlusterCaller software (LGC Limited, Fordham, UK).

2.4. Bioinformatic and Statistical Analysis

Minor allele frequencies (MAF), observed and expected heterozygosity and the Hardy–Weinberg equilibrium were measured at each polymorphic marker with Haploview v4.2 [22]. Pairwise linkage disequilibrium (LD) measures, including the normalized coefficient of linkage disequilibrium (D′) and the squared correlation (r2), as well as haplotype frequencies and the possible occurrence of LD blocks, were estimated with the Haploview v4.2 software package [23].
The association analysis between 44 polymorphic marker genotypes and milk traits of 825 does of the Sarda breed was based on the model:
Y = µ + G + F + P + DIM + eijklmn,
where Y is the observed trait; µ is the general mean; G is the fixed effect of the SNP genotype, one at a time (2 to 3 levels: 2 homozygous and the heterozygous); F is the random effect of the farm, which also includes animal management and feeding (1 to 19 levels, representing the different farms included in the experiment); P is the fixed effect of the parity of the goats (1 to 4 levels; first to fourth or more parities); DIM is the fixed effect of the days in milking (4 levels; 1 = ≤100 d; 2 = 101–140 d; 3 = 141–160 d; 4 = ≥161 d); and eijklmn is the random residual effect.
The same model as in Equation (1) was used to evaluate the association between milk traits and each of the 6 haplotype blocks, one at a time. In the single SNP and LD block analyses, we only considered SNP with an MAF > 0.05 to make sure that genotypic means were correctly estimated. The MIXED procedure of SAS (version 9.4, SAS Institute Inc., Cary, NC, USA) was used to carry out the association analysis. Multiple testing was performed using the Bonferroni method (one milk trait for each SNP or LD block at a time). Model effects were significant at p < 0.05.

3. Results

3.1. Genotyping

All 44 genotyped SNPs were polymorphic in the does population (Table 1), while 39 SNPs were polymorphic in the bucks population (Table S1). The SNPs’ distribution is displayed in Figure S1.
Table 1. List of marker mapping to the casein genes, with population parameters, for 825 Sarda goats (Capra hircus) reared in Sardinia (Italy).
Table 1. List of marker mapping to the casein genes, with population parameters, for 825 Sarda goats (Capra hircus) reared in Sardinia (Italy).
Gene#SNP ID 1Chr_6 Position 2Gene RegionObsHET 3PredHET 4Hwpval 5MAF 6Alleles
CSN1S101rs15550552485977292upstream0.330.310.170.19A:(G)
02CNS1S1_41985977582upstream0.010.011.000.01T:(G)
03rs66193053385977593upstream0.070.070.160.04C:(G)
04rs66805987785977607upstream0.020.021.000.01A:(G)
05rs15550552685977912upstream0.230.315.0 × 10−70.19T:(C)
06rs15550552785977916upstream0.330.393.5 × 10−50.27A:(G)
07rs66471903385977985upstream0.460.480.180.39C:(T)
08rs15550552885978133upstream0.450.500.010.47G:(A)
09rs15550552985978197upstream0.340.391.0 × 10−40.27A:(G)
10rs64418935385982583intron_30.060.070.150.04A:(G)
11rs66305075585982740intron_40.340.398.8 × 10−50.27G:(A)
CSN212rs65825366486006394exon_90.450.470.210.39C:(T)
13rs15550553986008016exon_70.440.450.680.33T:(C)
14rs15550554086015338upstream0.070.070.670.04A:(G)
15rs15550554186015756upstream0.180.180.850.10G:(A)
16rs63986877386015830upstream0.010.011.000.01G:(A)
17rs15550554486016649upstream0.470.500.200.46C:(T)
CSN1S218CNS1S2_45186075209upstream0.200.210.590.12C:(T)
19CNS1S2_43786075223upstream0.470.500.160.48A:G
20CNS1S2_67086075442upstream0.460.490.050.45G:(T)
21CNS1S2_67986075451upstream0.460.500.050.45T:(C)
22CNS1S2_34786075961upstream0.460.490.030.45G:(A)
23rs66027898786088523intron_130.450.460.430.36A:(C)
24rs26829309886088596intron_140.350.350.880.22T:(C)
25rs15550554786088687intron_140.350.350.730.22C:(G)
26rs65486780386088716Intron_140.450.460.520.36G:(A)
27rs65085923886088978Intron_140.440.460.340.36C:(T)
28rs66657848286089078Intron_140.450.460.460.36A:(G)
29rs15550554886089096Intron_140.170.170.620.10C:(T)
30rs64965318486089260intron_140.030.031.000.01A:(G)
31rs26829310086093348exon_170.440.460.360.36T:(C)
32rs63825988686093378exon_170.170.504.4 × 10−870.50C:(G)
CSN333rs15550555386195919upstream0.450.450.770.35A:(T)
34rs66916208286195968upstream0.380.390.630.27C:(T)
35rs15550555586196117upstream0.450.450.790.35G:(C)
36rs15550555786196315upstream0.300.344.0 × 10−40.22T:(G)
37rs65875707086196318upstream0.490.470.280.38C:(T)
38rs63674770186196358upstream0.470.440.080.33T:(C)
39rs66687211286196545upstream0.400.380.440.26C:(T)
40rs63570633886196623upstream0.460.480.240.39A:(T)
41rs65145712386196658upstream0.500.480.370.40AATC:(_)
42rs64468388686196737upstream0.390.360.780.24G:(A)
43rs15550556086196912upstream0.390.360.040.24G:(T)
44rs26829311486209184Exon_40.200.230.010.13A:(G)
1 SNP ID = dbSNP reference records (https://www.ncbi.nlm.nih.gov/projects/SNP/, last accessed on 16 November 2023); 2 Chr_6 position = position on chromosome 6, ASM170441v1 Sequence ID—NC_030813.1; 3 ObsHET = observed heterozygosity; 4 PredHET = predicted heterozygosity; 5 HWpval = Hardy–Weinberg test p-value. 6 MAF = minor allele frequency (minor allele in brackets).
In the does population, at the CSN1S1 gene, four out of eleven SNPs had minor allele frequency (MAF) lower than 0.05, and four SNPs were out of Hardy–Weinberg Equilibrium (HWE). At CSN2, all SNPs were in HW equilibrium, in the CSN1S2 gene only 1 SNP out of 15 was highly significant for HWE, while in the CSN3 gene, 3 SNPs were out of HWE (Table 1). In the bucks population, five SNPs overall were monomorphic and only one SNP of the CSN1S2 gene did not follow Hardy–Weinberg Equilibrium (rs638259886) (Supplementary Table S1). The DNA polymorphic markers used in the present investigation fell mostly in the upstream DNA region of the four casein genes, although some were in intronic regions and 3’UTR (CSN1S2 gene). All markers were SNPs (Single Nucleotide Polymorphism) except one indel (insertion-deletion), rs651457123 at the CSN3 gene, and two missense-variants, rs155505539 at CSN2 exon 7, and rs268293114 at CSN3 exon 4. Both missense variants were predicted to be tolerated by the SIFT prediction algorithm (http://sift.jcvi.org/, last accessed on 10 November 2023).

3.2. Descriptive Statistics and Association Analysis

The descriptive statistics of daily milk yield and quality traits of 825 Sarda does are shown in Supplementary Table S2. The most variable phenotypes were LBC and milk yield (MY), with coefficients of variation of 48% and 45%, respectively.
The results from the ANOVA (F-values) of daily milk yield and quality traits versus 44 SNP genotypes mapped to the casein genes are reported in Table S3. The SNPs with significant effects on phenotype variance are detailed in Table 2, and the effects of haplotype blocks are reported in Table 3.
At CSN1S1, two out of eleven SNPs were significantly associated with milk protein content (p < 0.001), rs664719033 and rs155505528, and these were also associated with fat content (p < 0.05 and p < 0.01, respectively). The same SNPs showed association with milk NaCl content (p < 0.05), and their impact on milk traits extended to total solids and milk energy traits (p < 0.001) (Table 2). At the CSN2 gene, the SNPs rs658253664 and rs155505544 were associated with milk protein content at p < 0.001 and with the fat content at p < 0.01; the same SNPs were associated with total solids and milk energy (Table 2). In addition, SNPs rs1555505539 and rs155505541 were associated with milk protein content at p < 0.01. All the above-mentioned SNPs were associated with NaCl content. SNP rs155505539 at the CSN2 gene was associated with SCS at p < 0.01. At the CSN1S2 gene, 10 out of 15 genotyped SNPs were highly associated with milk protein content (p < 0.001), and 4 SNPs were associated with milk fat content (p < 0.01); this association was also reflected in milk total solids and milk energy (Table 2). In total, 6 SNPs at CSN1S2 were associated with the milk yield trait (p < 0.05) and 10 SNPs were associated with milk pH (p < 0.05); moreover, 6 SNPs were associated with the SCS trait (p < 0.01) and 6 SNPs were associated with milk NaCl content (Table 2). At the CSN3 gene, 3 out of 12 genotyped SNPs were associated with milk protein content at p < 0.001, and 2 SNPs were associated with protein content at p < 0.01. Furthermore, only two SNPs were associated with fat content (p < 0.05) and showed contextual association with milk energy (p < 0.01). Finally, CSN3 SNP rs658757070 was associated with milk NaCl content.

3.3. LD Analysis

Linkage disequilibrium (LD) analysis was performed on doe (Figure 3 and Figure 4) and buck (Figures S2 and S3) populations.
After the exclusion of SNPs out of HWE and those with MAF values lower than 0.05, the casein cluster’s haplotype structure included 35 SNPs for does and 38 SNPs for bucks, spanning about 231 kbp, with an average distance between SNPs of 5.3 kb, ranging from 3 bp to 102 kbp. Block 1 (28 kb) in the doe population included two tag SNPs of the CSN1S1 gene located in the promoter region and one SNP of the CSN2 gene located in exon 9 (Figure 1). The same structure was visible in the buck population, in which block 2 included three SNPs from the CSN1S1 gene and two SNPs from the CSN2 gene. The CSN1S2 and CSN3 genes displayed two haplotype blocks each in both populations. In both populations, block 6 consisted of six SNPs, and had seven different haplotype combinations, while block 1 and block 5 comprised three SNPs, and had three possible haplotype combinations (Figure 3 and Figure S1).

4. Discussion

The genotyping of 44 SNPs at casein genes, in a sample population of 825 does and 153 bucks, revealed the high variability of the Sarda breed goat, as all analyzed SNPs were polymorphic, and of these 15% were rare (MAF < 0.05). Bucks also displayed 9% monomorphic SNPs, probably due to a different selection pressure on males and females, related to milk production, and to the effects of positive and purifying selection on nucleotide variation and diversity [24].
In Sarda goats, genotyping of the functional allele variants at the four casein genes revealed the occurrence of all investigated alleles in does and bucks, including rare defective variants such as CSN1S1 E (with a frequency of 0.03) and null alleles, only at Ca-sensitive casein genes, with frequencies lower than 0.01 Vacca 2014 [25]. The high variability of Sarda goats is probably due to the strategies adopted by the breeders on the island [26]. In addition to these strategies to maintain variability, farmers have also made, in the past, unplanned crosses with exotic breeds to increase milk production, but this proved to be a counterproductive practice in the long run [16]. The variability of casein cluster genes and structure has been reported in livestock species [27,28,29,30] and in other species, such as donkeys [31] camelids [32,33], elephants [34,35] and llamas [36], contributing to the reconstructing phylogenesis of the DNA region and bringing new insights into the casein function and formation of casein micelles (reviewed in [37]).

4.1. Association Analysis—The Outstanding Role of CSN1S2

In the present study, we revealed an association of casein gene variants with milk protein and fat content. This association extends to milk total solids and milk energy. The association of the casein genes’ variability with milk protein and lipid content has been reported in different goat breeds [38,39], and it has also been highlighted in other dairy species, such as sheep [40,41] and cattle [42,43]. Martin et al. (2017) [44], in a GWAS study, found that the casein cluster region was associated with protein content and fat content in French goats. Massender et al. (2023) [45] found that casein genes were of key economic importance in Canadian Saanen and Alpine goats, and revealed the presence of several candidate positional and functional SNPs in the region.
The SNP panel used in the present investigation allowed us to highlight, in this context, the outstanding role played by the CSN1S2 gene as regards its association with milk yield (MY), in addition to the association with milk protein content, shared among the four casein genes, and the association with fat content, shared among the Ca-sensitive caseins genes. SNPs associated with MY were located in the CSN1S2 gene body, at introns 13 and 14 and exon 17. That association might be due to LD with an unknown causal mutation, or to unknown biological pathways, which still need to be elucidated. The CSN1S2 gene is phylogenetically the most recent in the casein cluster [13], and variation in this gene might have given a selective advantage in Sarda dairy goats. MY is a polygenic quantitative trait, of high economic and productive importance, and more studies are needed to better understand the biological mechanisms that regulate this type of productive trait. The association of CSN1S2 gene variants with milk fat and total solid concentration has been revealed in Chinese goats (Yue 2013 [46]), with daily milk yield and milk coagulation parameters in Sarda goats [47] and milk protein and casein contents in sheep [41]; no association between CSN2 and CSN1S2 genes and milk protein or dry matter contents were found in Murciano Granadina goats [48]. The CSN1S2 gene was associated with Somatic Cell Score (SCS) in the present study, another critical parameter, considered an indirect indicator of milk hygienic quality, important in the physiology of lactation in goats [49,50].

4.2. CSN1S1 and CSN2 Genes Are Associated with Milk Protein and Fat Content

The CSN1S1 and CSN2 genes are convergently transcribed and are just 10.9 kbp apart on goat chromosome 6. Linkage Disequilibrium (LD) analysis showed the occurrence of a haplotype block including SNPs rs664719033, mapping to the CSN1S1 upstream region, and rs658253664, mapping to the CSN2 exon 9 (block 1) in the does population. Similarly, LD block 2 included SNPs from the CSN1S1 and CSN2 genes in bucks. The two genes have a total of four SNPs associated with milk protein and lipid content, and consequently also with total solids and milk energy.
In French dairy goats, the CSN1S1 gene has been reported to explain about 40% of protein content variation, and its genotype can be used as a positional and functional candidate to better predict genomic breeding values [51]. Different alleles of the αs1-casein gene have been reported to affect milk protein content, protein yield, fat content, and milk yield [13,52] in dairy goats. In addition, polymorphism of the caprine CSN1S1 gene has been reported to affect the technological properties of milk, such as cheese yield and cheese curd formation [53].
Variants at the CSN1S1 gene have been associated with four different levels of as1-casein in milk; in Sarda goats, the CSN1S1 B allele was reported to be the most frequent and was associated with an as1-casein content in milk of 4.94 g/L per allele [54], higher than the values reported for the CSN1S1 A variant [14]. The CSN1S1 F “weak” allele had a frequency of 0.25, and the remaining defective alleles were rare [25]. The defective alleles of as1-casein can influence in different ways the correct formation of casein micelles, causing variations in the morphology of the mammary gland epithelial cells [39]. Some authors do not detect any effect of casein variants CSN1S1 and CSN3 on the physicochemical characteristics of Saanen goat milk [55]. Some studies have revealed an effect of CSN2 on milk quality traits and milk renneting properties, except for milk yield, in Sarda goats [53].

4.3. CSN3 Variation Is Associated with Protein Content

The CSN3 gene showed a highly significant association with proteins, as did all the casein genes. Weak associations were found with lipid content, total solids, and milk energy. This is in accordance with the structural role played by k-casein in stabilizing casein micelles, and with a fundamental role in the coagulation processes that occur in the newborn’s stomach or during transformation processes aimed at cheesemaking [56]. The variability of this gene in goat breeds has been reported [57]. Association analysis in Sarda goats revealed a highly significant effect of variation at the CSN3 gene on milk coagulation parameters [53]. Susilorini et al. (2022) [58] indicated an association between CSN3 gene polymorphism and milk yield and composition in indigenous Indonesian goats.
Rahmatalla et al. (2021) [59] analyzed the casein cluster in search of new rare variants in both wild and farmed Indian goats, and revealed that most novel SNP variants occurred in the endangered Nubian ibex. The genetic variability of local breeds has been shaped over time in their environment, and represents a precious reservoir of biodiversity that should be protected. Although local breeds, such as the Sarda goat, are less productive than transboundary breeds, they are adapted to the local climate. They can be reared under extensive farming systems in a natural environment [60].

5. Conclusions

Variability of the Sarda does at casein genes was associated with milk protein content. In addition, CSN1S1 and CSN2, which were in linkage disequilibrium, were also associated with the lipid content, and then with milk total solids and milk energy. The CSN1S2 gene differed from the other genes because, in addition to its association with milk protein and fat content, it also showed an association with milk yield and other milk traits (pH value, SCS, NaCl). The CSN3 gene variants were associated with milk protein content. This information might be used in selection schemes and might guide future investigations to better understand lactation’s biology, potentially disclosing the direct link between genotype and phenotype.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ani14010056/s1, Figure S1: Distribution of SNPs on Casein genes; Figure S2: LD structure of the casein gene cluster in a population of 153 Sarda bucks; Figure S3: Description of haplotype blocks in 153 Sarda bucks; Table S1: List of SNPs of 153 Sarda bucks, mapping to the casein genes, with population parameters; Table S2: Descriptive statistics of milk yield and composition of 825 milk samples of Sarda goats; Table S3: Analysis of variance (F-value and significance) of each SNP for milk yield and composition of 825 Sarda does milk samples.

Author Contributions

Conceptualization, M.L.D. and G.M.V.; data curation, M.P.; formal analysis, M.L.D., M.P. and A.N.; methodology, V.L.; project administration, G.M.V.; writing—original draft, M.L.D.; writing—review and editing, M.L.D., M.P., A.N., V.L. and G.M.V. All authors have read and agreed to the published version of the manuscript.

Funding

Research supported by the Regional Government of Sardinia (Progetto Strategico Sulcis; CUP J73C17000070007, Cagliari, Italy). The authors thank the farmers for giving access to their flocks, and thank the APAs (Provincial Farmers Associations) of Cagliari and Nuoro (Italy), and the firms Società Agricola is Crabaxius (Fluminimaggiore, Italy), Azienda Agricola Murgia Antonello (Sant’Anna Arresi, Italy), Latteria Sociale Santadi (Santadi, Italy), Azienda Agricola F.lli Secci s.s. (Iglesias, Italy), and Agricola Allevatori Tallaroga, Soc. Coop. (Villamassargia, Italy) for their support in sample collection. We also thank ARA Sardegna (Regional Farmers Association of Sardinia, Cagliari, Italy) for support in chemical milk analysis.

Institutional Review Board Statement

Milk samples were collected by experienced technicians for recording in the herd book, which was not directly linked to the present study. Blood samples for DNA extraction reported in the present study were approved by the Ethical Animal Care and Experimental Use Committee (Organismo Preposto al Benessere e alla Sperimentazione Animale, OPBSA) at the University of Sassari (protocol number 0122930, approved on 28 September 2021).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Silanikove, N. The physiological basis of adaptation in goats to harsh environments. Small Rumin. Res. 2000, 35, 181–193. [Google Scholar] [CrossRef]
  2. FAO. Website of the FAO Pastoral Knowledge Hub. 2021. Available online: https://www.fao.org/pastoralist-knowledge-hub/en/ (accessed on 20 June 2023).
  3. Nori, M. Assessing the Policy Frame in Pastoral Areas of Europe; EUI RSC Working Paper 2022/03; PASTRES and Global Governance Programme: Florence, Italy, 2022. [Google Scholar]
  4. Paraskevopoulou, C.; Theodoridis, A.; Johnson, M.; Ragkos, A.; Arguile, L.; Smith, L.; Vlachos, D.; Arsenos, G. Sustainability Assessment of Goat and Sheep Farms: A Comparison between European Countries. Sustainability 2020, 12, 3099. [Google Scholar] [CrossRef]
  5. Paschino, P.; Stocco, G.; Dettori, M.L.; Pazzola, M.; Marongiu, M.L.; Pilo, C.E.; Cipolat-Gotet, C.; Vacca, G.M. Characterization of milk composition, coagulation properties, and cheese-making ability of goats reared in extensive farms. J. Dairy Sci. 2020, 103, 5830–5843. [Google Scholar] [CrossRef] [PubMed]
  6. DAD-IS FAO, the Domestic Animal Diversity Information System. Available online: https://www.fao.org/dad-is/en/ (accessed on 15 May 2023).
  7. Scano, P.; Caboni, P. Seasonal Variations of Milk Composition of Sarda and Saanen Dairy Goats. Dairy 2022, 3, 528–540. [Google Scholar] [CrossRef]
  8. Martin, P.; Ollivier-Bousquet, M.; Grosclaude, F. Genetic polymorphism of caseins: A tool to investigate casein micelle organization. Int. Dairy J. 1999, 9, 163–171. [Google Scholar] [CrossRef]
  9. Barillet, F.; Arranz, J.-J.; Carta, A. Mapping quantitative trait loci for milk production and genetic polymorphisms of milk proteins in dairy sheep. Genet. Sel. Evol. 2005, 37, S109–S123. [Google Scholar] [CrossRef] [PubMed]
  10. Rijnkels, M. Multispecies comparison of the casein gene loci and evolution of casein gene family. J. Mammary Gland Biol. Neoplasia 2002, 7, 327–345. [Google Scholar] [CrossRef]
  11. Leroux, C.; Mazure, N.; Martine, P. Mutations Away from Splice Site Recognition Sequences Might cis-Modulate Alternative Splicing of Goat aS1-Casein Transcripts. Structural organization of the relevant gene. J. Biol. Chem. 1992, 267, 6147–6157. [Google Scholar] [CrossRef]
  12. Jansa Perez, M.; Leroux, C.; Sanchez Bonastre, A.; Martin, P. Occurrence of a LINE element in the 3′ UTR of an allelic form of the goat as1-casein gene associated with a reduced level of protein synthesis. Gene 1994, 147, 179–187. [Google Scholar] [CrossRef]
  13. Carillier Jacquin, C.; Larroque, H.; Granié, R.C. Including αs1 casein gene information in genomic evaluations of French dairy goats. Genet. Sel. Evol. 2016, 48, 54. [Google Scholar] [CrossRef]
  14. Caravaca, F.; Carrizosa, J.; Urrutia, B.; Baena, F.; Jordana, J.; Amills, M.; Badaoui, B.; Sanchez, A.; Angiolillo, A.; Serradilla, J.M. Effect of αs1-casein (CSN1S1) and κ-casein (CSN3) genotypes on milk composition in Murciano-Granadina goats. J. Dairy Sci. 2009, 92, 2960–2964. [Google Scholar] [CrossRef] [PubMed]
  15. Pizarro Inostroza, M.G.; Landi, V.; Navas González, F.J.; Jurado, J.M.L.; Amparo Martínez, M.; Fernández Álvarez, J.; Delgado Bermejo, J.V. Non-parametric association analysis of additive and dominance effects of casein complex SNPs on milk content and quality in Murciano-Granadina goats. J. Anim. Breed Genet. 2020, 137, 407–422. [Google Scholar] [CrossRef] [PubMed]
  16. Usai, M.G.; Casu, S.; Molle, G.; Decandia, M.; Ligios, S.; Carta, A. Using cluster analysis to characterize the goat farming system in Sardinia. Livest. Sci. 2006, 104, 63–76. [Google Scholar] [CrossRef]
  17. International Standard ISO 9622 and IDF 141:2013. ISO; Milk and Liquid Milk Products: Determination of Fat, Protein, Casein, Lactose, and pH Content. ISO-IDF (International Organization for Standardization and International Dairy Federation): Brussels, Belgium, 2013.
  18. ISO 6731:2010 (IDF 21:2010); Milk, Cream and Evaporated Milk. Determination of Total Solids Content (Reference method). International Organisation for Standardisation: Geneva, Switzerland, 2010.
  19. NRC. Nutrient Requirements of Dairy Cattle, 7th ed.; National Academy Press: Washington, DC, USA, 2001. [Google Scholar] [CrossRef]
  20. Shook, G.E. Genetic improvement of mastitis through selection on somatic cell count. Vet. Clin. N. Am. Food Anim. Pract. 1993, 9, 563–577. [Google Scholar] [CrossRef] [PubMed]
  21. International Standard ISO 21187 and IDF 196:2004; Milk: Quantitative Determination of Bacteriological Quality—Guidance for Establishing and Verifying a Conversion Relationship between Routine Method Results and Anchor Method Results. ISO: Geneva, Switzerland; IDF: Brussels, Belgium, 2004.
  22. Barrett, J.C.; Fry, B.; Maller, J.; Daly, M.J. Haploview: Analysis and visualization of LD and haplotype maps. Bioinformatics 2005, 21, 263–265. [Google Scholar] [CrossRef]
  23. Gabriel, S.B.; Schaffner, S.F.; Nguyen, H.; Moore, J.M.; Roy, J.; Blumenstiel, B.; Altshuler, D. The structure of haplotype blocks in the human genome. Science 2002, 23, 2225–2229. [Google Scholar] [CrossRef]
  24. Cvijović, I.; Good, B.H.; Desai, M.M. The Effect of Strong Purifying Selection on Genetic Diversity. Genetics 2018, 209, 1235–1278. [Google Scholar] [CrossRef]
  25. Vacca, G.M.; Dettori, M.L.; Piras, G.; Manca, F.; Paschino, P.; Pazzola, M. Goat casein genotypes are associated with milk production traits in the Sarda breed. Anim. Genet. 2014, 45, 723–731. [Google Scholar] [CrossRef]
  26. Piras, D.; Doro, M.G.; Casu, G.; Melis, M.P.; Vaccargiu, S.; Piras, I.; Debora, P.; Stradoni, R.; Frongia, B.; Lai, G.; et al. Haplotype affinities resolve a major component of goat (Capra hircus) mtDNA D-loop diversity and reveal specific features of the Sardinian stock. PLoS ONE 2012, 7, e30785. [Google Scholar] [CrossRef]
  27. Palmeri, M.; Mastrangelo, S.; Sardina, M.T.; Portolano, B. Genetic Variability at αs2-casein Gene in Girgentana Dairy Goat Breed. Ital. J. Anim. Sci. 2014, 13, 2997. [Google Scholar] [CrossRef]
  28. Darwish, A.M.; El Nady, G.H.; Ali, N.I.; Abdelsalam, A.Z.E. Evaluation of β-casein variants in Egyptian goat, sheep and cattle by allele specific PCR and relevance to β-casomorphin. Indian J. Anim. Res. 2018, 52, 799–804. [Google Scholar] [CrossRef]
  29. Criscione, A.; Tumino, S.; Avondo, M.; Marletta, D.; Bordonaro, S. Casein haplotype diversity in seven dairy goat breeds. Arch. Anim. Breed. 2019, 62, 447–454. [Google Scholar] [CrossRef] [PubMed]
  30. Asim, M.; Saif-ur-Rehman, M.; Hassan, F. Comparative genomic characterization and screening of regulatory regions of casein gene family in Bos taurus and Bubalus bubalis. J. Glob. Innov. Agric. Sci. 2022, 10, 147–158. [Google Scholar] [CrossRef]
  31. Cunsolo, V.; Saletti, R.; Muccilli, V.; Gallina, S.; Di Francesco, A.; Foti, S. Proteins and bioactive peptides from donkey milk: The molecular basis for its reduced allergenic properties. Food Res. Int. 2017, 99, 41–45. [Google Scholar] [CrossRef]
  32. Pauciullo, A.; Shuiep, E.T.; Ogah, M.D.; Cosenza, G.; Di Stasio, L.; Erhardt, G. Casein Gene Cluster in Camelids: Comparative Genome Analysis and New Findings on Haplotype Variability and Physical Mapping. Front. Genet. 2019, 10, 748. [Google Scholar] [CrossRef]
  33. Mutery, A.A.; Rais, N.; Mohamed, W.K.; Abdelaziz, T. Genetic Diversity in Casein Gene Cluster in a Dromedary Camel (C. dromedarius) Population from the United Arab Emirates. Genes 2021, 12, 1417. [Google Scholar] [CrossRef]
  34. Runthala, A.; Mbye, M.; Ayyash, M.; Xu, Y.; Kamal-Eldin, A. Caseins: Versatility of Their Micellar Organization in Relation to the Functional and Nutritional Properties of Milk. Molecules 2023, 28, 2023. [Google Scholar] [CrossRef]
  35. Madende, M.; Kemp, G.; Stoychev, S.; Osthoff, G. Characterisation of African elephant beta casein and its relevance to the chemistry of caseins and casein micelles. Int. Dairy J. 2018, 85, 112–120. [Google Scholar] [CrossRef]
  36. Pauciullo, A.; Erhardt, G. Molecular Characterization of the Llamas (Lama glama) Casein Cluster Genes Transcripts (CSN1S1, CSN2, CSN1S2, CSN3) and Regulatory Regions. PLoS ONE 2015, 10, e0124963. [Google Scholar] [CrossRef]
  37. Rahmatalla, S.A.; Arends, D.; Brockmann, G.A. Review: Genetic and protein variants of milk caseins in goats. Front. Genet. 2022, 13, 995349. [Google Scholar] [CrossRef]
  38. Martin, P.; Szymanowska, M.; Zwierzchowski, L.; Leroux, C. The impact of genetic polymorphisms on the protein composition of ruminant milks. Reprod. Nutr. Dev. 2002, 42, 433–459. [Google Scholar] [CrossRef] [PubMed]
  39. Rodrigues, M.T.; Soares, M.A.M.; Zacaroc, A.A.; Silva, M.M.C.; Garcia, O.S.R.; Magalhães, A.C.M. Differences in the defective alleles E and F for the locus CSN1S1 in goats affects the profile of milk caseins. Small Rumin. Res. 2015, 123, 47–54. [Google Scholar] [CrossRef]
  40. Giambra, I.J.; Brandt, H.; Erhardt, G. Milk protein variants are highly associated with milk performance traits in East Friesian Dairy and Lacaune sheep. Small Rumin. Res. 2014, 121, 382–394. [Google Scholar] [CrossRef]
  41. Luigi-Sierra, M.G.; Marmol-Sanchez, E.; Amills, M. Comparing the diversity of the casein genes in the Asian mouflonand domestic sheep. Anim. Genet. 2020, 51, 470–475. [Google Scholar] [CrossRef] [PubMed]
  42. Cecchinato, A.; Chessa, S.; Ribeca, C.; Cipolat-Gotet, C.; Bobbo, T.; Casellas, J.; Bittante, G. Genetic variation and effects of candidate-gene polymorphisms on coagulation properties, curd firmness modeling and acidity in milk from Brown Swiss cows. Animal 2015, 9, 1104–1112. [Google Scholar] [CrossRef] [PubMed]
  43. Marina, H.; Pelayo, R.; Suárez-Vega, A.; Gutiérrez-Gil, B.; Esteban-Blanco, C.; Arranz, J.J. Genome-wide association studies (GWAS) and post-GWAS analyses for technological traits in Assaf and Churra dairy breeds. J. Dairy Sci. 2021, 104, 11850–11866. [Google Scholar] [CrossRef] [PubMed]
  44. Martin, P.; Palhière, I.; Maroteau, C.; Bardou, P.; Canale-Tabet, K.; Sarry, J.; Tosser-Klopp, G. A genome scan for milk production traits in dairy goats reveals two new mutations in Dgat1 reducing milk fat content. Sci. Rep. 2017, 7, 1872. [Google Scholar] [CrossRef] [PubMed]
  45. Massender, E.; Oliveira, H.R.; Brito, L.F.; Maignel, L.; Jafarikia, M.; Baes, C.F.; Sullivan, B.; Schenkel, F.S. Genome-wide association study for milk production and conformation traits in Canadian Alpine and Saanen dairy goats. J. Dairy Sci. 2023, 106, 1168–1189. [Google Scholar] [CrossRef]
  46. Yue, X.P.; Fang, Q.; Zhang, X.; Mao, C.C.; Lan, X.Y.; Chen, H.; Lei, C.Z. Effects of CSN1S2 Genotypes on Economic Traits in Chinese Dairy Goats. Asian Australas. J. Anim. Sci. 2013, 26, 911–915. [Google Scholar]
  47. Deniz, D.; Sena, A.; Hale, S.; Buse, V.; Faruk, B. Identification of the frequency of CSN1S2 gene alleles and the effects of these alleles and parity on milk yield and composition in Saanen goats. Large Anim. Rev. 2021, 27, 91–96. [Google Scholar]
  48. Zidi, A.; Amills, M.; Jordana, J.; Carrizosa, J.; Urrutia, B.; Serradilla, J.M. Variation at the 3′-UTR of the goat aS2- and b-casein genes is not associated with milk protein and dry matter contents in Murciano-Granadina goats. Anim. Genet. 2014, 46, 91–99. [Google Scholar] [CrossRef]
  49. Jimenez-Granado, R.; Molina, A.; Ziadi, C.; Sanchez, M.; Muñoz-Mejías, E.; Demyda-Peyrás, S.; Menendez-Buxadera, A. Genetic Parameters of Somatic Cell Score in Florida Goats Using Single and Multiple Traits Models. Animals 2022, 12, 1009. [Google Scholar] [CrossRef] [PubMed]
  50. Lianou, D.T.; Michael, C.K.; Vasileiou, N.G.C.; Liagka, D.V.; Mavrogianni, V.S.; Caroprese, M.; Fthenakis, G.C. Association of Breed of Sheep or Goats with Somatic Cell Counts and Total Bacterial Counts of Bulk-Tank Milk. Appl. Sci. 2021, 11, 7356. [Google Scholar] [CrossRef]
  51. Teissier, M.; Larroque, H.; Granié, R.C. Weighted single-step genomic BLUP improves accuracy of genomic breeding values for protein content in French dairy goats: A quantitative trait infuenced by a major gene. Genet. Sel. Evol. 2018, 50, 31. [Google Scholar] [CrossRef] [PubMed]
  52. Grosclaude, F.; Mahé, M.F.; Brignon, G.; Di Stasio, L.; Jeunet, R. A Mendelian polymorphism underlying quantitative variations of goat αs1-casein. Genet. Sel. Evol. 1987, 19, 399–412. [Google Scholar] [CrossRef] [PubMed]
  53. Martin, P.; Leroux, C. Le gène caprin spécifiant la caséine αs1: Un suspect tout désigné aux effets aussi multiples qu’inattendus. INRA Prod. Anim. HS2000 2000, 13, 125–132. [Google Scholar] [CrossRef]
  54. Balia, F.; Pazzola, M.; Dettori, M.L.; Mura, M.C.; Luridiana, S.; Carcangiu, V.; Piras, G.; Vacca, G.M. Effect of CSN1S1 gene polymorphism and stage of lactation on milk yield and composition of extensively reared goats. J. Dairy Res. 2013, 80, 129–137. [Google Scholar] [CrossRef]
  55. Deniz, D.; Sena, A.; Hale, S.; Faruk, B. Determination the effect of CSN1S1, CSN3 and AGPAT6 genes and lactation rank on physicochemical properties of goat milk. Large Anim. Rev. 2020, 26, 167–171. [Google Scholar]
  56. Oftedal, O.T. The evolution of milk secretion and its ancient origins. Animal 2012, 6, 355–368. [Google Scholar] [CrossRef]
  57. Di Gerlando, R.; Tortorici, L.; Sardina, M.T.; Monteleone, G.; Mastrangelo, S.; Portolano, B. Molecular Characterisation of κ–Casein Gene in Girgentana Dairy Goat Breed and Identification of Two New Alleles. Ital. J. Anim. Sci. 2015, 14, 3464. [Google Scholar] [CrossRef]
  58. Susilorini, T.E.; Furqon, A.; Septian, W.A.; Wulandari, D.; Suyadi, S. Polymorphism of Kappa-Casein Gene (CSN3|Alw44I) and Its Effect on Milk Yield and Compositions in Indigenous Senduro Goat. Adv. Anim. Vet. Sci. 2022, 10, 1333–1338. [Google Scholar] [CrossRef]
  59. Rahmatalla, S.A.; Arends, D.; Ahmed, A.S.; Hassan, L.M.A.; Krebs, S.; Reissmann, M.; Brockmann, G.A. Capture Sequencing to Explore and Map Rare Casein Variants in Goats. Front. Genet. 2021, 12, 620253. [Google Scholar] [CrossRef] [PubMed]
  60. Ajmone-Marsan, P.; Boettcher, P.J.; Colli, L.; Ginja, C.; Kantanen, J.; Lenstra, J.A. (Eds.) Genomic Characterization of Animal Genetic Resources—Practical Guide; FAO: Rome, Italy, 2023. [Google Scholar] [CrossRef]
Figure 1. Sarda goats, reared on the island of Sardina.
Figure 1. Sarda goats, reared on the island of Sardina.
Animals 14 00056 g001
Figure 2. Geographic map showing the distribution of sampling farms in Sardinia. Made with Google MyMaps, https://www.google.com/intl/it/maps/about/mymaps/ (Last accessed on 16 November 2023).
Figure 2. Geographic map showing the distribution of sampling farms in Sardinia. Made with Google MyMaps, https://www.google.com/intl/it/maps/about/mymaps/ (Last accessed on 16 November 2023).
Animals 14 00056 g002
Figure 3. LD structure of the casein gene cluster DNA region in a population of 825 Sarda does. The structural organization of the four casein genes respects gene positions on goat chromosome 6, as represented in Genome Data Viewer (https://www.ncbi.nlm.nih.gov/genome/gdv/?org=capra-hircus, last accessed on 18 November 2023). LD plot of pairwise normalized coefficient of linkage disequilibrium (D′): coloured in red, D′ = 1.0 and logarithm of the odds (LOD) ≥ 2.0; coloured in light blue, D′ = 1.0 and LOD < 1.0.
Figure 3. LD structure of the casein gene cluster DNA region in a population of 825 Sarda does. The structural organization of the four casein genes respects gene positions on goat chromosome 6, as represented in Genome Data Viewer (https://www.ncbi.nlm.nih.gov/genome/gdv/?org=capra-hircus, last accessed on 18 November 2023). LD plot of pairwise normalized coefficient of linkage disequilibrium (D′): coloured in red, D′ = 1.0 and logarithm of the odds (LOD) ≥ 2.0; coloured in light blue, D′ = 1.0 and LOD < 1.0.
Animals 14 00056 g003
Figure 4. Haplotype blocks defined by the 44 SNPs genotyped in the four casein genes in 825 Sarda does. Haplotype frequencies are reported near each haplotype within a block. For each crossing area, a value of the multiallelic normalized coefficient of linkage disequilibrium (D′) is given for the haplotypes displayed, which represents the level of recombination between the two blocks. SNP numbering is reported in Table 1.
Figure 4. Haplotype blocks defined by the 44 SNPs genotyped in the four casein genes in 825 Sarda does. Haplotype frequencies are reported near each haplotype within a block. For each crossing area, a value of the multiallelic normalized coefficient of linkage disequilibrium (D′) is given for the haplotypes displayed, which represents the level of recombination between the two blocks. SNP numbering is reported in Table 1.
Animals 14 00056 g004
Table 2. Least square means of Sarda goat milk traits according to the different genes, SNPs and genotypes (n = 825).
Table 2. Least square means of Sarda goat milk traits according to the different genes, SNPs and genotypes (n = 825).
GeneSNPGenotypenMYFat %Protein %Lactose %Total SolidspHSCSLBCNaClMilk Energy
CSN1S1rs155505526CC330.975.16 ab4.234.7815.15 ab6.765.811.84213.53.81 ab
TC961.045.43 a4.224.7615.40 a6.735.781.84212.63.91 a
TT2941.035.08 b4.114.7714.94 b6.735.701.85214.93.75 b
rs664719033CC3011.025.62 a4.09 A4.6415.31 A6.716.411.70229.6 b3.93 A
TC3631.045.41 b3.97 B4.6415.02 B6.726.601.66237.3 ab3.82 B
TT1341.065.28 b3.87 C4.6314.74 B6.726.351.58242.7 a3.74 B
rs155505528AA1961.075.22 B3.90 C4.6414.72 B6.726.431.60240.8 a3.73 B
GA3621.045.48 A3.99 B4.6415.10 A6.726.561.65235.6 ab3.85 A
GG2491.025.63 A4.09 A4.6415.32 A6.716.451.72230.3 b3.94 A
CSN2rs658253664CC3121.025.61 A4.09 A4.6415.29 A6.716.411.70230.1 b3.93 B
TC3641.045.41 AB3.97 B4.6415.02 B6.726.611.66237.2 ab3.82 B
TT1291.075.25 B3.86 C4.6314.72 C6.736.361.57242.3 a3.73 B
rs155505539CC931.015.594.10 A4.6615.276.716.02 B1.60226.5 b3.91
TC3531.035.494.03 AB4.6415.136.716.57 B1.71233.5 ab3.86
TT3621.065.403.94 B4.6314.986.726.52 B1.63239.0 a3.82
rs155505541AA90.895.62 ab4.28 AB4.6715.53 A6.696.031.82216.8 b3.98 A
GA1441.025.70 a4.07 A4.6515.38 AB6.716.411.69228.4 ab3.95 A
GG6511.055.42 b3.98 B4.6415.00 B6.726.521.65237.1 a3.82 B
rs155505544CC2401.025.64 A4.08 A4.6515.31 A6.716.311.68229.0 B3.93 A
TC3771.035.43 B4.00 A4.6315.06 AB6.716.641.69236.6 A3.84 B
TT1801.075.30 B3.90 B4.6314.80 B6.736.511.58241.8 A3.76 B
CSN1 S2CNS1S2_437AA2241.025.604.09 A4.6415.31 A6.716.321.67230.93.93 A
GA3801.045.414.00 AB4.6314.99 B6.726.581.67237.73.82 B
GG1981.075.423.90 B4.6614.99 B6.736.531.63236.03.81 B
CNS1S2_670GG2591.015.604.09 A4.6315.29 A6.71 a6.401.67232.23.93 a
TG3711.045.413.99 B4.6314.98 B6.72 b6.571.67236.83.82 B
TT1761.065.393.89 B4.6614.97 B6.73 a6.481.63235.93.79 B
CNS1S2_679CC1761.075.413.89 B4.6614.98 B6.73 a6.461.63236.13.80 B
TC3711.045.403.99 B4.6314.98 B6.72 b6.591.67237.23.82 B
TT2591.015.594.09 A4.6315.29 A6.71 b6.401.67232.23.92 A
CNS1S2_347AA1741.075.433.90 b4.6615.016.736.481.63235.93.81 b
GA3651.055.403.99 ab4.6414.98 a6.726.591.67236.93.82 ab
GG2631.015.584.08 a4.6315.27 b6.716.371.67232.23.92 a
rs660278987AA3381.00 b5.60 A4.08 A4.6415.30 A6.71 b6.29 B1.68230.9 B3.93 A
CA3581.07 a5.36 B3.97 B4.6214.92 B6.71 b6.72 A1.66240.0 A3.80 B
CC1081.07 a5.37 AB3.87 B4.6814.93 AB6.74 a6.43 B1.61233.5 B3.78 B
rs268293098CC381.015.694.074.6215.536.70 b6.65 ab1.66230.33.98
TC2791.025.504.044.6115.126.71 b6.71 a1.68237.93.87
TT4851.055.433.974.6615.026.73 a6.38 b1.65234.13.83
rs155505547CC4811.055.433.97 b4.6515.026.72 a6.39 b1.65234.23.83
GC2811.015.514.04 a4.6115.136.71 b6.70 a1.69237.63.88
GG371.025.684.06 a4.6515.526.71 b6.35 b1.59227.13.98
rs654867803AA1041.07 a5.35 B3.87 C4.69 a14.91 A6.74 a6.37 B1.61232.5 B3.77 B
GA3511.07 a5.34 B3.96 B4.62 b14.89 A6.71 b6.70 A1.66239.8 A3.79 B
GG3301.00 b5.60 A4.08 A4.64 ab15.30 B6.71 b6.26 B1.69230.9 B3.93 A
rs650859238CC3371.00 b5.61 A4.08 A4.6415.30 A6.71 b6.29 B1.68230.9 a3.93 A
TC3561.07 a5.36 B3.97 B4.6214.92 B6.72 ab6.71 A1.67239.7 b3.80 B
TT1101.06 a5.37 B3.87 C4.6814.91 B6.74 a6.49 AB1.60234.3 ab3.78 B
rs666578482AA3361.00 b5.59 a4.08 A4.6415.29 A6.71 b6.28 B1.68231.0 b3.92 A
GA3571.07 a5.36 b3.96 B4.6214.92 B6.72 b6.71 A1.67239.9 a3.80 B
GG1081.07 b5.36 b3.87 B4.6814.91 B6.74 a6.47 AB1.60234.1 b3.78 B
rs268293100CC1091.07 a5.37 b3.87 B4.6814.926.74 a6.44 AB1.61234.0 a3.78 B
TC3551.07 a5.36 b3.97 AB4.6214.926.71 b6.72 A1.66239.8 a3.80 B
TT3351.00 b5.60 a4.07 A4.6415.296.71 b6.27 B1.67230.8 b3.92 A
rs638259886CC3361.09 a5.41 B3.95 B4.6415.00 AB6.70 a6.65 AB1.65236.6 AB3.81 AB
CG1321.07 a5.31 B3.98 B4.5714.87 B6.68 b6.83 A1.69244.6 A3.77 B
GG3341.02 b5.66 A4.09 A4.6215.38 A6.68 b6.21 B1.69230.9 B3.94 A
CSN3rs669162082CC4321.035.55 a4.05 A4.6415.23 A6.726.461.67233.43.90 A
TC3061.065.36 b3.96 B4.6414.93 B6.726.491.64236.63.80 B
TT611.085.44 ab3.92 B4.6514.93 B6.716.691.66234.83.81 B
rs658757070CC3071.065.383.92 B4.6414.90 b6.726.651.65238.5 a3.80 b
TC3911.025.504.03 A4.6315.14 ab6.716.451.68235.5 ab3.87 a
TT1051.045.534.12 B4.6715.28 a6.736.321.65227.1 b3.90 a
rs635706338AA2761.065.393.93 B4.6414.93 b6.726.621.65238.43.81
TA3311.035.494.01 AB4.6415.09 ab6.726.441.67235.33.86
TT1201.025.564.12 A4.6615.30 a6.726.431.65229.63.92
rs651457123AA2791.075.403.92 B4.6414.93 b6.726.631.66238.93.81
AT3981.035.494.01 AB4.6315.11 ab6.716.451.67235.03.86
TT1231.025.564.12 A4.6615.30 a6.736.411.65228.93.92
rs644683886AA4310115.57 a4.23 A4.5915.39 a6.706.361.67231.93.94 A
GA29310375.58 a4.06 AB4.6315.27 ab6.696.421.74233.33.91 A
GG46210735.44 b3.97 B4.6315.02 b6.696.581.64237.53.83 B
rs155505560GG4531.025.484.024.6515.136.726.36 b1.65233.03.86
TG3141.075.423.984.6214.976.716.69 a1.67238.93.82
TT351.015.613.924.6215.216.736.33 b1.74236.33.92
For each SNP, least squares means with different superscripts are different. Capital letters = p < 0.01; lowercase letters = p < 0.05.
Table 3. Least square means of Sarda goat milk traits according to the different haplotype blocks.
Table 3. Least square means of Sarda goat milk traits according to the different haplotype blocks.
GeneSNPGenotypenMYFat %Protein %Lactose %Total SolidspHSCSLBCNaClMilk Energy
Haplotype Block 1CGC2450.995.534.10 A4.6515.22 a6.716.521.73232.73.89 a
Blocks TAT1251.055.283.87 B4.6314.74 b6.736.231.58240.63.74 b
Block 2TGT1771.055.283.86 b4.63 b14.756.736.371.52243.8 a3.74
CGC451.115.494.06 a4.69 a15.156.756.271.45228.2 b3.88
Block 3CAGTG2221.025.594.09 A4.6415.306.71 b6.341.65229.93.93
CGTCA1001.015.463.89 B4.6515.056.74 a6.441.62236.13.81
Block 4ATCGCACT1011.045.504.04 a4.6615.166.71 b5.951.61230.23.88
CTCATGCC1011.035.383.90 b4.6814.936.74 a6.461.64232.93.78
For each SNP, least squares means with different superscripts are different. Capital letters = p < 0.01; lowercase letters = p < 0.05.
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

Dettori, M.L.; Pazzola, M.; Noce, A.; Landi, V.; Vacca, G.M. Variations in Casein Genes Are Associated with Milk Protein and Fat Contents in Sarda Goats (Capra hircus), with an Important Role of CSN1S2 for Milk Yield. Animals 2024, 14, 56. https://doi.org/10.3390/ani14010056

AMA Style

Dettori ML, Pazzola M, Noce A, Landi V, Vacca GM. Variations in Casein Genes Are Associated with Milk Protein and Fat Contents in Sarda Goats (Capra hircus), with an Important Role of CSN1S2 for Milk Yield. Animals. 2024; 14(1):56. https://doi.org/10.3390/ani14010056

Chicago/Turabian Style

Dettori, Maria Luisa, Michele Pazzola, Antonia Noce, Vincenzo Landi, and Giuseppe Massimo Vacca. 2024. "Variations in Casein Genes Are Associated with Milk Protein and Fat Contents in Sarda Goats (Capra hircus), with an Important Role of CSN1S2 for Milk Yield" Animals 14, no. 1: 56. https://doi.org/10.3390/ani14010056

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