Next Article in Journal
Evaluating the Oxidative Stress in Inflammation: Role of Melatonin
Previous Article in Journal
MicroRNA Profiling of B Cell Subsets from Systemic Lupus Erythematosus Patients Reveals Promising Novel Biomarkers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Characterization of Bovine SMO Gene and Effects of Its Genetic Variations on Body Size Traits in Qinchuan Cattle (Bos taurus)

1
College of Animal Science and Technology, Northwest A&F University, Yangling 712100, Shaanxi, China
2
National Beef Cattle Improvement Center of Northwest A&F University, Yangling 712100, Shaanxi, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this study.
Int. J. Mol. Sci. 2015, 16(8), 16966-16980; https://doi.org/10.3390/ijms160816966
Submission received: 23 March 2015 / Revised: 2 June 2015 / Accepted: 30 June 2015 / Published: 27 July 2015
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)

Abstract

:
Smoothened (Smo)-mediated Hedgehog (Hh) signaling pathway governs the patterning, morphogenesis and growth of many different regions within animal body plans. This study evaluated the effects of genetic variations of the bovine SMO gene on economically important body size traits in Chinese Qinchuan cattle. Altogether, eight single nucleotide polymorphisms (SNPs: 1–8) were identified and genotyped via direct sequencing covering most of the coding region and 3ʹUTR of the bovine SMO gene. Both the p.698Ser.>Ser. synonymous mutation resulted from SNP1 and the p.700Ser.>Pro. non-synonymous mutation caused by SNP2 mapped to the intracellular C-terminal tail of bovine Smo protein; the other six SNPs were non-coding variants located in the 3ʹUTR. The linkage disequilibrium was analyzed, and five haplotypes were discovered in 520 Qinchuan cattle. Association analyses showed that SNP2, SNP3/5, SNP4 and SNP6/7 were significantly associated with some body size traits (p < 0.05) except SNP1/8 (p > 0.05). Meanwhile, cattle with wild-type combined haplotype Hap1/Hap1 had significantly (p < 0.05) greater body length than those with Hap2/Hap2. Our results indicate that variations in the SMO gene could affect body size traits of Qinchuan cattle, and the wild-type haplotype Hap1 together with the wild-type alleles of these detected SNPs in the SMO gene could be used to breed cattle with superior body size traits. Therefore, our results could be helpful for marker-assisted selection in beef cattle breeding programs.

1. Introduction

Smoothened (Smo)-mediated Hedgehog (Hh) signaling pathway is a critical regulator of cell growth and patterning during embryonic development, and also is involved in stem cell renewal and tissue homeostasis in adult animals [1,2,3]. Perturbations in the pathway are related to birth defects and various cancers [4]. Smo, a seven-transmembrane protein, belongs to the Frizzled (FzD) class of G-protein-coupled receptor (GPCR) superfamily and serves as the obligatory signal transducer of the Hh pathway [5,6,7]. Specifically, Smo transforms the extracellular Hh protein signal into an intracellular glima-associated transcription factors (Gli1-3) protein signal, thereby activating intranuclear target genes [8,9]. As a core element of the Hh pathway, SMO gene is conserved from flies to humans [1], and enormous progress has been made in revealing not only its molecular structure and functional mechanism but also its cellular and developmental functions. Functional studies in model animals have shown that SMO gene participates in both osteogenesis and myogenesis through Hh pathway.
In mice, the forced expression of a constitutively actived SMO allele promoted chondrocyte proliferation, while chondrocyte-specific SMO knockout mice showed a marked decrease in chondrocyte proliferation and developed shorter long bones than wild-type littermates [10]. Similarly, the removal of SMO from perichondrial cells led to bone collar defects and abolished development of primary spongiosa. Meanwhile, in chimeric mice, cells genetically deficient in SMO exhibited a cell-autonomous defect in osteoblast differentiation in bone collar and primary spongiosa [11]. In zebrafish, inactive Smo mutant embryos displayed phenotypic defects, including abnormalities in body size, cartilage, pectoral fins, central nervous system [12]. Compared to wild-type zebrafish embryos, Smo mutant embryos lacked slow muscle, and their fast muscle fibers were disorganized [13]. A recent study reported that the fine coordination of Smo activity by the miR-30 family controlled the specification and differentiation of distinct muscle cell types of zebrafish embryos [14]. Collectively, these findings confirm SMO contributes to the development and growth of bone and muscle, suggesting that SMO is an attractive candidate gene for the selection of growth-related traits in livestock.
Little study on the bovine SMO gene has been performed, since most research has focused on model animals and human. Based on its role in osteogenesis and myogenesis as demonstrated in mice and zebrafish, we proposed the hypothesis that variations of the SMO gene affected body size traits in cattle. Here, we do research on Qinchuan cattle, for as a famous indigenous cattle breed in China, its growth rate and underdeveloped hind hip need to be improved to be comparable to imported beef cattle breeds [15]. Molecular characterization of the bovine SMO was analyzed using bioinformatics. Screening of genetic variations was performed by direct sequencing. Then the genetic associations with body size traits in Qinchuan cattle were examined. Our results are potentially beneficial for further research in enhancing the economic traits of beef cattle.

2. Results and Discussion

2.1. Sequence Homology, Inferred Phylogenetic Tree and Sequence Alignments

SMO gene maps to bovine chromosome 4, consists of 12 exons divided by 11 introns and 3ʹUTR, and encodes 780 amino acids. BLAST analysis revealed that the amino acid sequence of the bovine SMO (NP_001179149.1) shared high identity with other vertebrates. The percent identity with human (NP_005622.1), rat (NP_036939.1), mouse (NP_795970.3), chicken (XP_414970.4), zebrafish (NP_571102.1) and xenopus (NP_001128704.1) were 94%, 93%, 92%, 79%, 71% and 70%, respectively. The relatively high identity among mammalians (92%–94%) suggested that the SMO gene was more conserved within this group. The phylogenetic tree (Figure 1) intuitively showed the relationship between bovine SMO and the potential evolutionary process; bovine, human, mouse and rat SMO fell into one evolutionarily related group, while chicken, xenopus and zebrafish SMO were dispersed at more distant branches.
Figure 1. Phylogenetic tree for amino acid sequences of the SMO gene in seven species with bootstrap confidence values at the branch nodes. Branch lengths indicated the evolutionary distances.
Figure 1. Phylogenetic tree for amino acid sequences of the SMO gene in seven species with bootstrap confidence values at the branch nodes. Branch lengths indicated the evolutionary distances.
Ijms 16 16966 g001
Research on model animals and human has revealed that Smo protein consists of three domains: the extracellular N-terminal domain, a seven heptahelical transmembrane domain, and the intracellular C-terminal tail [16]. Using the already known information of other species, to better understand the structure of bovine Smo protein, we performed multiple alignments of the amino acid sequences among the above mentioned species. Figure 2 showed the aligned Smo amino acid sequences; asterisks denoted the fully conserved amino acid residues. Based on the recently determined three-dimensional structure of human Smo (hSmo) [16], analysis indicated that in the seven species, the most highly conserved region covered the seven heptahelical transmembrane domain (residues 224–534 of hSmo), in accordance with the common feature in all GPCRs [17]. The regions comprising the transmembrane domain were highlighted with a red line. The upstream and downstream the transmembrane domain were the extracellular N-terminal and intracellular C-terminal domains, respectively.
Figure 2. Aligned multiple amino acid sequences of the SMO gene in seven species. To make the assessment of alignment much easier, amino acid residues with different chemical properties are differentiated with different background colors. Asterisks denote positions that have been fully conserved; the “:” and “.” characters indicate positions where conservative and semi-conserved amino acid mutations have happened, respectively. And the red line highlights the inferred transmembrane domain.
Figure 2. Aligned multiple amino acid sequences of the SMO gene in seven species. To make the assessment of alignment much easier, amino acid residues with different chemical properties are differentiated with different background colors. Asterisks denote positions that have been fully conserved; the “:” and “.” characters indicate positions where conservative and semi-conserved amino acid mutations have happened, respectively. And the red line highlights the inferred transmembrane domain.
Ijms 16 16966 g002

2.2. Analysis of Sequence Variations in Qinchuan Cattle SMO Gene

We amplified and sequenced most of the coding region and 3ʹUTR of the SMO gene. Compared with the SMO gene reference sequence (GenBank accession No. AC_000161), a total of eight SNPs (Figure 3), including two coding variants (SNP1: g.22935C>T, SNP2: g.22939T>C) in exon 12 and six non-coding variants (SNP3: g.23232C>T, SNP4: g.23283C>A, SNP5: g.23329C>T, SNP6: g.23458T>G, SNP7: g.23633T>C, SNP8: g.23641C>A) in the 3ʹUTR, were identified in 520 Qinchuan cattle. SNP1 (g.22935C>T) resulted in a p.698Ser.>Ser. synonymous mutation and SNP2 (g.22939T>C) led to a p.700Ser.>Pro. missense mutation. According to the above aligned Smo amino acid sequences (Figure 2), both the p.698Ser.>Ser. synonymous mutation and the p.700Ser.>Pro. non-synonymous mutation were located in the intracellular C-terminal tail of bovine Smo. In addition, compared with the public SNP information of the bovine SMO gene provided by NCBI, SNP1 and SNP8 were identified as two novel variations and will be submitted into the SNP data bank.
Figure 3. The sequencing maps of the eight detected single nucleotide polymorphisms (SNPs) within the bovine SMO gene. SNP1: g.22935C>T; SNP2: g.22939T>C; SNP3: g.23232C>T; SNP4: g.23283C>A; SNP5: g.23329C>T; SNP6: g.23458T>G; SNP7: g.23633T>C; SNP8: g.23641C>A.
Figure 3. The sequencing maps of the eight detected single nucleotide polymorphisms (SNPs) within the bovine SMO gene. SNP1: g.22935C>T; SNP2: g.22939T>C; SNP3: g.23232C>T; SNP4: g.23283C>A; SNP5: g.23329C>T; SNP6: g.23458T>G; SNP7: g.23633T>C; SNP8: g.23641C>A.
Ijms 16 16966 g003
All eight SNPs were successfully genotyped by PCR-direct sequencing. The detailed SNP information for each cattle was provided in Table S1. The two novel SNPs (SNP1 and SNP8) exhibited two different genotypes: wild-type homozygote and heterozygote; whereas, the others displayed three kinds of genotypes: wild-type homozygote, heterozygote and mutant homozygote (Table 1).
Table 1 also contained the results of genotypic and allelic frequencies, genetic diversity parameters (He, Ne and PIC) and Hardy-Weinberg equilibrium. The wild-type alleles of all eight SNP loci were predominant in the studied population. Results of χ2 test indicated that genotype distributions of the eight mutations were all in Hardy-Weinberg equilibrium (χ2 < χ20.05). The PIC value (if PIC < 0.25, 0.25 < PIC < 0.5, or PIC > 0.5, low, medium or high polymorphism, respectively [18]) manifested that the two novel SNPs were in low polymorphism, while the rest were located on medium genetic diversity level in the studied population.
Table 1. Genotypic, allelic frequencies and genetic diversity of eight SNP loci within the SMO gene in Qinchuan cattle populations.
Table 1. Genotypic, allelic frequencies and genetic diversity of eight SNP loci within the SMO gene in Qinchuan cattle populations.
LociGenotypic FrequenciesTotalAllelic Frequenciesχ2 (HWE)PICHeNe
Wild TypeHybrid SubtypeMutant TypeWild TypeMutant Type
SNP1/8CC/CCCT/CATT/AA520C/CT/A0.55840.05960.06141.0655
0.93650.063500.96830.0317
SNP2TTTCCC520TC1.01860.37080.49171.9673
0.30770.51350.17880.56440.4356
SNP3/5CC/CCCT/CTTT/TT520C/CT/T0.02920.29200.35501.5505
0.59040.35770.05190.76920.2308
SNP4CCCAAA520CA0.25850.34960.45151.8231
0.42500.46150.11350.65580.3442
SNP6/7TT/TTTG/TCGG/CC520T/TG/C0.16340.35270.45731.8426
0.41350.46540.12120.64620.3538
HWE, Hardy-Weinberg equilibrium; χ20.05 = 5.991; χ20.01 = 9.210.

2.3. Linkage Disequilibrium and Haplotype Analysis

Table 2 illustrated the results of linkage disequilibrium (LD). According to the r2-value, a pairwise measure of LD, SNP1 and SNP8 loci were in perfect LD (r2 = 1), so were SNP3 and SNP5 (r2 = 1), and SNP6 and SNP7 (r2 = 1). r2 = 1 indicates that, during the history of the sample, two SNPs have not been separated by recombination, and observations at one SNP provide complete information about the other SNP [19]. Thus, SNP1 and SNP8 loci were abbreviated as a single SNP1/8 locus, SNP3 and SNP5 as SNP3/5, and SNP6 and SNP7 as SNP6/7. In addition, SNP2, SNP3/5, SNP4 and SNP6/7 loci exhibited strong LD (r2 > 0.33) with each other. This may be a result of selection which can result in nonrandom associations among SNPs and thus elevate levels of LD in a gene; especially selection, during domestication, aiming at alleles of a gene can significantly elevate levels of LD in a given region [20,21].
Table 2. Estimated values of linkage disequilibrium analysis for eight detected SNPs.
Table 2. Estimated values of linkage disequilibrium analysis for eight detected SNPs.
r2SNP2SNP3SNP4SNP5SNP6SNP7SNP8
SNP10.0420.0080.0620.0080.0600.0601.000
SNP2 0.3890.6800.3890.7100.7100.042
SNP3 0.5711.0000.5480.5480.008
SNP4 0.5710.3850.3850.062
SNP5 0.5480.5480.008
SNP6 1.0000.060
SNP7 0.060
If r2 = 1, SNP pairs are in perfect linkage disequilibrium; If r2 > 0.33, SNP pairs are in strong linkage disequilibrium.
Subsequently, a total of five haplotypes were identified (Table 3). Theoretically, the number of the inferred haplotypes should be 256 (28), but both the perfect and strong LDs between SNP marker pairs notably decreased the number. Obviously, Hap1 (-CTCCCTTC-), the combination of wild-type alleles, was the most frequent (0.558), followed by Hap2 (0.234), Hap3 (0.092), Hap4 (0.085), and lastly the combination of mutant-type alleles Hap5 (-TCCACGCA-) (0.031). The high-frequency haplotypes might have been in the population for a long time and more adaptive to the environment or selection [22,23]. Consequently, new mutants are more likely to originate from the common haplotypes, meaning that rarer variants are more closely related to the common haplotypes and represent recent mutations [24].
Table 3. Information of haplotypes of the SMO gene in Qinchuan cattle population.
Table 3. Information of haplotypes of the SMO gene in Qinchuan cattle population.
HaplotypeSNPsFrequency
12345678
Hap1CTCCCTTC0.564
Hap2CCTATGCC0.231
Hap3CCCCCGCC0.091
Hap4CCCACTTC0.082
Hap5TCCACGCA0.032

2.4. Effects of Single Marker on Body Size Traits

Association analysis between the single SNP markers and body size traits were performed (Table 4). At the SNP1/8 locus, there were no significant effects on body size traits (p > 0.05) (data not shown).
At the SNP2 locus, animals with wild-type genotype TT had significantly greater body length (p = 0.005), wither height (p = 0.011), hip height (p = 0.022), hip width (p = 0.004) and heart girth (p = 0.018) than those with mutant genotype CC; significant difference was also found between genotype TC and CC for hip width (p = 0.015). As mentioned above, SNP2 led to the p.700Ser.>Pro. non-synonymous mutation in the intracellular C-terminal tail of bovine Smo. Intriguingly, in common with other GPCRs, phosphorylation of Smo regulates the switch between on/off signaling states [6]; or rather the activation of Smo is triggered by hyperphosphorylating Ser/Thr residues within C-terminal cytoplasmic tail during Hh signaling [25,26,27,28]. Substitutions of phosphorylation sites with un-phosphorylatable residues render Smo inactive and diminish Hh signal activity, whereas phosphor-mimetic Smo variants display overactive Hh signal [25,29,30]. Moreover, with the increase in the number of phosphor-mimetic mutations in its C-terminal tail, Smo activity exhibits a progressive elevation. [30]. Herein, we deduced that individuals with mutant allele SNP2-C showed significantly lower body size traits than those with wild-type allele SNP2-T (p < 0.05) might be caused by the attenuation of Smo activation via the substitution of Ser. residue with un-phosphorylatable residue Pro. Subsequently, the attenuation might undergo cascade amplification, consequently influencing the transcription of downstream target genes.
Table 4. Associations between different genotypes of SNPs detected in SMO and body size traits in Qinchuan cattle.
Table 4. Associations between different genotypes of SNPs detected in SMO and body size traits in Qinchuan cattle.
LociGenotypesBody Size Traits (Mean ± SE)
BL (cm)WH (cm)HH (cm)RL (cm)HW (cm)CD (cm)HG (cm)
SNP 2TT (160)135.422 ± 0.934 a121.937 ± 0.654 a124.422 ± 0.562 a42.263 ± 0.34939.025 ± 0.451 a59.156 ± 0.539164.131 ± 1.397 a
TC (267)133.333 ± 0.723120.936 ± 0.506123.414 ± 0.43542.049 ± 0.27038.577 ± 0.349 a58.470 ± 0.417162.534 ± 1.067
CC (93)130.565 ± 1.225 b118.790 ± 0.858 b121.925 ± 0.738 b41.215 ± 0.45836.645 ± 0.592 b57.301 ± 0.112157.871 ± 1.808 b
p-value 10.0050.0110.0220.2080.004/0.0150.0810.018
SNP3/5CC/CC (307)134.718 ± 0.673 a121.406 ± 0.473 a123.785 ± 0.40742.231 ± 0.25238.827 ± 0.327 a58.866 ± 0.389163.383 ± 0.995 a
CT/CT (186)132.239 ± 0.865120.470 ± 0.608123.312 ± 0.52341.677 ± 0.32437.995 ± 0.42058.081 ± 0.500161.409 ± 1.279
TT/TT (27)127.963 ± 2.271 b117.352 ± 1.595 b120.741 ± 1.37240.926 ± 0.85035.741 ± 1.101 b56.685 ± 1.312154.037 ± 3.356 b
p-value 10.0140.0460.1020.4240.0220.3350.024
SNP4CC (221)135.262 ± 0.794 a121.776 ± 0.558 a124.122 ± 0.48042.235 ± 0.29738.905 ± 0.385 a59.102 ± 0.458163.833 ± 1.176
CA (240)132.650 ± 0.762120.512 ± 0.535123.198 ± 0.46041.854 ± 0.28538.292 ± 0.37058.160 ± 0.440161.723 ± 1.128
AA (59)130.186 ± 1.537 b118.847 ± 1.079 b122.025 ± 0.92841.407 ± 0.57636.678 ± 0.746 b57.381 ± 0.887157.949 ± 2.275
p-value 10.0100.0490.1360.6050.0250.2570.066
SNP6/7TT/TT (215)134.935 ± 0.808 a121.800 ± 0.565124.277 ± 0.48642.326 ± 0.30138.972 ± 0.390 a58.988 ± 0.465163.579 ± 1.189 a
TG/TC (242)132.963 ± 0.761120.512 ± 0.533123.054 ± 0.45841.864 ± 0.28438.326 ± 0.36858.227 ± 0.439162.388 ± 1.121
GG/CC (63)130.508 ± 1.492 b118.992 ± 1.045122.214 ± 0.89841.127 ± 0.55636.476 ± 0.720 b57.651 ± 0.860156.698 ± 2.197 b
p-value 10.0280.0550.1320.1760.0070.5160.018
a,b Means with different superscripts are significantly different (p < 0.05); 1 p-values after modified Bonferroni correction for trait-wise multiple tests.
At the SNP3/5 locus, animals with wild-type genotype CC/CC showed significantly larger body length (p = 0.014), wither height (p = 0.046), hip width (p = 0.022) and heart girth (p = 0.024) compared to those with mutant genotype TT/TT. At the SNP4 locus, individuals with genotype CC demonstrated higher mean values for body length (p = 0.010), wither height (p = 0.049) and hip width (p = 0.025) than those with genotype AA. At the SNP6/7 locus, the body length (p = 0.028), hip width (p = 0.007) and heart girth (p = 0.018) of animals with wild-type genotype TT/TT were significantly larger than those of animals with mutant genotype GG/CC. Although these SNPs identified in the 3ʹUTR were non-coding DNA variants, there is growing indication that 3ʹUTR variants actively participate in modifying gene expression patterns. A G > A transition in the 3ʹUTR of the GDF8 gene created microRNA target sites for mir1 and mir206, consequently promoting the muscular hypertrophy phenotype in Texel sheep [31]. Additionally, an A > G substitution in the 3ʹUTR of the pig PPARA gene associated with adipose tissue accumulation was found located near the putative target sequence for mir224 and potentially increased the mir224 binding to the PPARA, thus reducing PPARA transcript level [32]. Accordingly, we hypothesized that variants in the 3ʹUTR of the bovine SMO gene might work in a manner parallel to SNPs in the 3ʹUTR of the GDF8 or PPARA gene, such as creating microRNA binding sites or possibly influencing the combination of the potential microRNAs with the SMO gene, thereby influencing the transcription level of SMO.
Collectively, wild-type individuals had larger body size traits in our Qinchuan cattle population, and wild-type alleles (SNP2-T, SNP3/5-C/C, SNP4-C, and SNP6/7-T/T) appeared to be beneficial for improving body size traits in cattle breeding programs.

2.5. Effects of Haplotype Combinations on Body Size Traits

Haplotype combinations may provide greater power than a single marker for genetic disease and trait associations [33]. Consequently, we analyzed the haplotype combinations of eight SNPs, and a total of 14 diplotypes (combined genotypes or haplotypes) were identified. We selected five diplotypes for association analysis; those with frequency far lower than 0.05 were not chosen. Individuals with wild-type diplotype Hap1/Hap1 (CC-TT-CC-CC-CC-TT-TT-CC) displayed significantly greater body length (p = 0.031) than those with Hap2/Hap2 (CC-CC-TT-AA-TT-GG-CC-CC) (Table 5), which suggested that diplotype Hap1/Hap1 could be used as a molecular marker in selection of preferable body size traits in cattle. Likewise, the results were in agreement with the conclusion of the effect of one SNP, the wild-type alleles (SNP2-T, SNP3/5-C/C, SNP4-C, and SNP6/7-T/T) related with greater body size traits. On the other hand, statistical analyses implied that mutations of these detected loci within SMO might lead to a decrease in body size traits.
Table 5. Associations of daplotypes with body size traits in Qinchuan cattle.
Table 5. Associations of daplotypes with body size traits in Qinchuan cattle.
DiplotypeBody Size Traits (Mean ± SE)
BL (cm)WH (cm)HH (cm)RL (cm)HW (cm)CD (cm)HG (cm)
Hap1/Hap1 (160)135.422 ± 0.952 a121.938 ± 0.668124.422 ± 0.57842.263 ± 0.35439.025 ± 0.45659.156 ± 0.547164.131 ± 1.375
Hap1/Hap2 (139)132.327 ± 1.022120.478 ± 0.716123.183 ± 0.62041.705 ± 0.38038.252 ± 0.48958.119 ± 0.587161.935 ± 1.476
Hap1/Hap3 (59)135.246 ± 1.568121.619 ± 1.099123.636 ± 0.95142.339 ± 0.58338.898 ± 0.75059.237 ± 0.900163.525 ± 2.265
Hap1/Hap4 (53)133.387 ± 1.655121.491 ± 1.160123.943 ± 1.00342.509 ± 0.61538.679 ± 0.79258.387 ± 0.950161.519 ± 2.390
Hap2/Hap2 (27)127.963 ± 2.318 b117.352 ± 1.625120.741 ± 1.40640.926 ± 0.86235.741 ± 1.10956.685 ± 1.331154.037 ± 3.348
p-value 10.0310.0940.1581.0000.0640.8660.055
a,b Means with different superscripts are significantly different (p < 0.05); 1 p-values after modified Bonferroni correction for trait-wise multiple tests.
Marker-assisted selection (MAS) based on genetic variation is more effective and powerful than traditional selection methodologies for genetically improving livestock economic traits, for example, growth traits, milk traits, reproduction traits, meat quality traits [34]. It is well established that genes involved in Hh signaling, including IHH, SHH, DHH, PTCH1, SMO and GLI1-3, play crucial roles in the growth, patterning and morphogenesis of various animal tissues [1,8,9]. Genome-wide association analysis in human has documented that variants within genes in Hh signaling are associated with adult height [35]. Studies in mice and zebrafish have demonstrated that SMO contributes to osteogenesis and myogenesis, consequently influencing the development and growth of bone and muscle [10,11,12,13,14]. These findings suggest that the SMO gene, as one mediating Hh signaling, could be a potential candidate gene related to animal body size traits. Our results showed that some variations within SMO were associated with body size traits in Qinchuan cattle, making it useful as a molecular marker in the MAS program for beef cattle.

3. Experimental Section

3.1. Animal Source, Data Collection and Genomic DNA Preparation

520 adult animals (18–24 months old; unrelated for at least three generations) were selected from the following farms: Fineness Breeding Center of Qinchuan Cattle (Yangling, Shaanxi, China), Reserved Farm of Qinchuan Cattle (Fufeng, Shaanxi, China) and Qinchuan Cattle Farm (Qian County, Shaanxi, China). Seven body measurement traits (body length (BL),wither height (WH), hip height (HH), rump length (RL), hip width (HW), chest depth (CD), heart girth (HG)) were measured for statistical analysis as Gilbert et al. [36] described. Meanwhile, blood samples were collected from the jugular vein and immediately treated with 2.0% heparin. Genomic DNA samples were extracted from blood using the standard phenol-chloroform protocol [37].

3.2. Bioinformatic Study

The amino acid sequences of the SMO gene for different species (Bos taurus, Homo sapiens, Rattus norvegicus, Mus musculus, Gallus gallus, Xenopus laevis and Danio rerio) were acquired using BLAST provided by NCBI. A phylogenetic tree for SMO was constructed using MEGA6.06 software. Multiple sequence alignment for orthologous Smo proteins was performed using Clustalx2.0 software.

3.3. Sequence Variant Detection

Primers (Table S2) used to amplify the bovine SMO gene were designed based on the published nucleotide sequence (GenBank accession No. AC_000161). Polymerase chain reaction (PCR) was conducted in 30 µL reaction mixtures, containing 10 pM of each primer, 50 ng templates DNA, 15 µL 2× Reaction Mix (500 M dNTP, 20 mM Tris-HCl (pH 8.3), 100 mM KCl, 3 mM MgCl2, other stabilizer and enhancer), and 0.3U Golden DNA polymerase (Tiangen Biotech, Beijing, China). The PCR was performed in a thermal cycler (Eppendorf, Germany) with the following procedure: initial denaturing at 95 °C for 5 min, followed by 35 cycles of 30 s at 94 °C, 30 s at the optimum annealing temperature (Table S1), 72 °C for 1 min, and ended with a final elongation of 10 min at 72 °C. PCR products were detected by electrophoresis on 1.5% (w/v) agarose gel (containing 200 ng/mL ethidium bromide) and purified by Axygen kits (MBI Fermentas, Amherst, NY, USA), and then sequenced in both directions in an ABI PRIZM377 DNA analyzer (Perkin-Elmer, Waltham, MA, USA). Sequence maps were imported into SeqMan of DNASTAR software (version 7.1) and analyzed to search for variations.

3.4. Genotyping

Primers (Table S2) were redesigned for genotyping eight detected SNPs via direct sequencing. PCR conditions, sequencing and sequence analysis were as described above.

3.5. Statistical Analyses

Genotype frequencies were calculated by direct counting. Allele frequencies, Hardy-Weinberg equilibrium (HWE), heterozygosity (He), and effective allele numbers (Ne) were analyzed based on genotype frequencies as Nei and Rychoudhury [38] described; polymorphism information content (PIC) was obtained via Botstein’s methods [18]. Linkage disequilibrium between all pairs of biallelic loci, as well as haplotypes, was analyzed using the Partition-Ligation Combination-Subdivision EM (PL-CSEM) algorithm of SHEsis software [39,40].
The associations between single SNP marker and body measurement traits were analyzed with SPSS software (version .19.0) using the general linear models (GLM) procedure and the following model:
Yijklm = μ + Gi + Aj +Fk +Sl + Sm + εijklm
where Yijklm = trait measured on each individual; μ = overall mean; Gi = fixed effect of the ith genotype; Aj = fixed effect of the jth age; Fk = fixed effect of the kth farm; Sl = fixed effect of the lth sex; Sm = fixed effect of the mth sire; and εijklm = random error.
For the association analysis between combined genotypes and the body size traits, the statistical model was similar to the model 1 with a slight modification, which was that Gi was the fixed effect associated with the ith combined genotypes. Moreover, to obtain more robust results, all the p values of the statistical results were corrected by Bonferroni correction which was used to account for multiple tests.

4. Conclusions

In summary, we analyzed the molecular characterization of the bovine SMO together with the SMO of several other species, and identified eight SNPs and five haplotypes in the coding region and 3ʹUTR of SMO in 520 Qinchuan cattle. Genotyping and association analyses demonstrated that wild-type alleles of some detected SNPs appeared to be more beneficial for selecting cattle with superior body size traits, and could be used as molecular markers for the improvement of Qinchuan cattle and perhaps other breedtypes that would benefit from marker-assisted selection for increased body size. Nevertheless, these results should be considered as preliminary, and further studies should be conducted to validate observed associations in a broad variety of cattle breeds. Further studies also should explore the molecular mechanisms responsible for these SNPs in the SMO gene associated with the variations in body size traits.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1422-0067/16/08/16966/s1.

Acknowledgments

The research was supported by the National 863 Program of China (#2013 AA102505 & #2011AA100307-02), the National Natural Science Foundation (#31272411), the National Beef and Yak Industrial Technology System (#CARS-38), the National Science-technology Support Plan Projects (#2012BAD28B04-03) and the Technical Innovation Engineering Project of Shaanxi Province (#2014 KTZB02-02-01).

Author Contributions

Lin-Sen Zan and Ya-Ran Zhang conceived and designed the experiments. Ya-Ran Zhang performed the experiments and wrote the manuscript. Lin-Sheng Gui mainly assisted in analyzing the data. Yao-Kun Li and Bi-Jie Jiang provided constructive suggestions for the discussion, and also assisted in the language modification. Hong-Cheng Wang and Ying-Ying Zhang helped to collect the samples and data, and disposal the data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ingham, P.W.; McMahon, A.P. Hedgehog signaling in animal development: Paradigms and principles. Genes Dev. 2001, 15, 3059–3087. [Google Scholar] [CrossRef] [PubMed]
  2. Beachy, P.A.; Karhadkar, S.S.; Berman, D.M. Tissue repair and stem cell renewal in carcinogenesis. Nature 2004, 432, 324–331. [Google Scholar] [CrossRef] [PubMed]
  3. Jiang, J.; Hui, C.C. Hedgehog signaling in development and cancer. Dev. Cell 2008, 15, 801–812. [Google Scholar] [CrossRef] [PubMed]
  4. Sharpe, H.J.; Wang, W.; Hannoush, R.N.; de Sauvage, F.J. Regulation of the oncoprotein smoothened by small molecules. Nat. Chem. Biol. 2015, 11, 246–255. [Google Scholar] [CrossRef] [PubMed]
  5. Alcedo, J.; Ayzenzon, M.; von Ohlen, T.; Noll, M.; Hooper, J.M. The Drosophila smoothened gene encodes a seven-pass membrane protein, a putative receptor for the hedgehog signal. Cell 1996, 86, 221–232. [Google Scholar] [CrossRef]
  6. Ayers, K.L.; Therond, P.P. Evaluating smoothened as a G-protein-coupled receptor for Hedgehog signaling. Trends Cell Biol. 2010, 20, 287–298. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, C.; Wu, H.; Katritch, V.; Han, G.W.; Huang, X.P.; Liu, W.; Siu, F.Y.; Roth, B.L.; Cherezov, V.; Stevens, R.C. Structure of the human smoothened receptor bound to an antitumour agent. Nature 2013, 497, 338–343. [Google Scholar] [CrossRef] [PubMed]
  8. McMahon, A.P.; Ingham, P.W.; Tabin, C.J. Developmental roles and clinical significance of hedgehog signaling. Curr. Top. Dev. Biol. 2003, 53, 1–114. [Google Scholar] [PubMed]
  9. Robbins, D.J.; Fei, D.L.; Riobo, N.A. The Hedgehog signal transduction network. Sci. Signal. 2012, 5, re6. [Google Scholar] [CrossRef] [PubMed]
  10. Long, F.; Zhang, X.M.; Karp, S.; Yang, Y.; McMahon, A.P. Genetic manipulation of hedgehog signaling in the endochondral skeleton reveals a direct role in the regulation of chondrocyte proliferation. Development 2001, 128, 5099–5108. [Google Scholar] [PubMed]
  11. Long, F.; Chung, U.I.; Ohba, S.; McMahon, J.; Kronenberg, H.M.; McMahon, A.P. Ihh signaling is directly required for the osteoblast lineage in the endochondral skeleton. Development 2004, 131, 1309–1318. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, W.; Burgess, S.; Hopkins, N. Analysis of the zebrafish smoothened mutant reveals conserved and divergent functions of hedgehog activity. Development 2001, 128, 2385–2396. [Google Scholar] [PubMed]
  13. Henry, C.A.; Amacher, S.L. Zebrafish slow muscle cell migration induces a wave of fast muscle morphogenesis. Dev. Cell 2004, 7, 917–923. [Google Scholar] [CrossRef] [PubMed]
  14. Ketley, A.; Warren, A.; Holmes, E.; Gering, M.; Aboobaker, A.A.; Brook, J.D. The miR-30 MicroRNA family targets smoothened to regulate Hedgehog signalling in Zebrafish early muscle development. PLoS ONE 2013, 8, e65170. [Google Scholar] [CrossRef] [PubMed]
  15. Adoligbe, C.; Zan, L.; Farougou, S.; Wang, H.; Ujjan, J.A. Bovine GDF10 gene polymorphism analysis and its association with body measurement traits in Chinese indigenous cattle. Mol. Biol. Rep. 2012, 39, 4067–4075. [Google Scholar] [CrossRef] [PubMed]
  16. Rognan, D.; Mus-Veteau, I. Three-Dimensional structure of the smoothened receptor: Implications for drug discovery. Top. Med. Chem. 2014. [Google Scholar] [CrossRef]
  17. Venkatakrishnan, A.J.; Deupi, X.; Lebon, G.; Tate, C.G.; Schertler, G.F.; Babu, M.M. Molecular signatures of G-protein-coupled receptors. Nature 2013, 494, 185–194. [Google Scholar] [CrossRef] [PubMed]
  18. Botstein, D.; White, R.L.; Skolnick, M.; Davis, R.W. Construction of a genetic linkage map in man using restriction fragment length polymorphisms. Am. J. Hum. Genet. 1980, 32, 314–331. [Google Scholar] [PubMed]
  19. Ardlie, K.G.; Kruglyak, L.; Seielstad, M. Patterns of linkage disequilibrium in the human genome. Genetics 2002, 3, 299–309. [Google Scholar] [CrossRef] [PubMed]
  20. Saunders, M.A.; Slatkin, M.; Garner, C.; Hammer, M.F.; Nachman, M.W. The extent of linkage disequilibrium caused by selection on G6PD in humans. Genetics 2005, 171, 1219–1229. [Google Scholar] [CrossRef] [PubMed]
  21. Clark, R.M.; Linton, E.; Messing, J.; Doebley, J.F. Pattern of diversity in the genomic region near the maize domestication gene tb1. Proc. Natl. Acad. Sci. USA 2004, 101, 700–707. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, Y.Z.; Wang, K.Y.; He, H.; Shen, Q.W.; Lei, C.Z.; Lan, X.Y.; Zhang, C.L.; Chen, H. Haplotype distribution in the GLI3 gene and their associations with growth traits in cattle. Gene 2013, 513, 141–146. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, G.; Zhang, S.; Wei, S.; Zhang, Y.; Li, Y.; Fu, C.; Zhao, C.; Zan, L. Novel polymorphisms of SIX4 gene and their association with body measurement traits in Qinchuan cattle. Gene 2014, 539, 107–110. [Google Scholar] [CrossRef] [PubMed]
  24. Posada, D.; Crandall, K.A. Intraspecific gene genealogies: Trees grafting into networks. Trends Ecol. Evol. 2001, 16, 37–45. [Google Scholar] [CrossRef]
  25. Apionishev, S.; Katanayeva, N.M.; Marks, S.A.; Kalderon, D.; Tomlinson, A. Drosophila Smoothened phosphorylation sites essential for Hedgehog signal transduction. Nat. Cell Biol. 2005, 7, 86–92. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, Y.; Li, S.; Tong, C.; Zhao, Y.; Wang, B.; Liu, Y.; Jia, J.; Jiang, J. G protein-coupled receptor kinase 2 promotes high-level Hedgehog signaling by regulating the active state of Smo through kinase-dependent and kinase-independent mechanisms in Drosophila. Genes Dev. 2010, 24, 2054–2067. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, Y.; Jiang, J. Decoding the phosphorylation code in Hedgehog signal transduction. Cell Res. 2013, 23, 186–200. [Google Scholar] [CrossRef] [PubMed]
  28. Maier, D.; Cheng, S.; Faubert, D.; Hipfner, D.R. A broadly conserved G-protein-coupled receptor kinase phosphorylation mechanism controls Drosophila smoothened activity. PLoS Genet. 2014, 10, e1004399. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, C.; Williams, E.H.; Guo, Y.; Lum, L.; Beachy, P.A. Extensive phosphorylation of Smoothened in Hedgehog pathway activation. Proc. Natl. Acad. Sci. USA 2004, 101, 17900–17907. [Google Scholar] [CrossRef] [PubMed]
  30. Jia, J.; Tong, C.; Wang, B.; Luo, L.; Jiang, J. Hedgehog signaling activity of smoothened requires phosphorylation by protein kinase A and casein kinase I. Nature 2004, 432, 1045–1050. [Google Scholar] [CrossRef] [PubMed]
  31. Clop, A.; Marcq, F.; Takeda, H.; Pirottin, D.; Tordoir, X.; Bibé, B.; Bouix, J.; Caiment, F.; Elsen, J.M.; Eychenne, F.; et al. A mutation creating a potential illegitimate microRNA target site in the myostatin gene affects muscularity in sheep. Nat. Genet. 2006, 38, 813–818. [Google Scholar] [CrossRef] [PubMed]
  32. Stachowiak, M.; Szydlowski, M.; Flisikowski, K.; Flisikowska, T.; Bartz, M.; Schnieke, A.; Switonski, M. Polymorphism in 3ʹ untranslated region of the pig PPARA gene influences its transcript level and is associated with adipose tissue accumulation. J. Anim. Sci. 2014, 92, 2363–2371. [Google Scholar] [CrossRef] [PubMed]
  33. Akey, J.; Jin, L.; Xiong, M.M. Haplotypes vs. single marker linkage disequilibrium tests: What do we gain? Eur. J. Hum. Genet. 2001, 9, 291–300. [Google Scholar] [CrossRef] [PubMed]
  34. Inbeagha-Awemu, E.M.; Kgwatalala, P.; Zhao, X. A critical analysis of production-associated DNA polymorphisms in the genes of cattle, goat, sheep, and pig. Mamm. Genome 2008, 19, 591–617. [Google Scholar] [CrossRef] [PubMed]
  35. Weedon, M.N.; Lango, H.; Lindgren, C.M.; Wallace, C.; Evans, D.M.; Mangino, M.; Freathy, R.M.; Perry, J.R.; Stevens, S.; Hall, A.S.; et al. Genome-wide association analysis identifies 20 loci that influence adult height. Nat. Genet. 2008, 40, 575–583. [Google Scholar] [CrossRef] [PubMed]
  36. Gilbert, R.P.; Bailey, D.R.; Shannon, N.H. Linear body measurements of cattle before and after 20 years of selection for postweaning gain when fed two different diets. J. Anim. Sci. 1993, 71, 1712–1720. [Google Scholar] [PubMed]
  37. Mullenbach, R.; Lagoda, P.J.; Welter, C. An efficient salt-chloroform extraction of DNA from blood and tissues. Trends Genet. 1989, 5, 39. [Google Scholar]
  38. Nei, M.; Roychoudhury, A.K. Sampling variance of heterozygosity and genetic distance. Genetics 1974, 76, 379–390. [Google Scholar] [PubMed]
  39. Li, Z.; Zhang, Z.; He, Z.; Tang, W.; Li, T.; Zeng, Z.; He, L.; Shi, Y. A partition-ligation-combination-subdivision EM algorithm for haplotype inference with multiallelic markers: Update of the SHEsis (http://analysis.bio-x.cn). Cell Res. 2009, 19, 519–523. [Google Scholar] [CrossRef] [PubMed]
  40. Shi, Y.Y.; He, L. SHEsis a powerful software platform for analyses of linkage disequilibrium, haplotype construction, and genetic association at polymorphism loci. Cell Res. 2005, 15, 97–98. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Zhang, Y.-R.; Gui, L.-S.; Li, Y.-K.; Jiang, B.-J.; Wang, H.-C.; Zhang, Y.-Y.; Zan, L.-S. Molecular Characterization of Bovine SMO Gene and Effects of Its Genetic Variations on Body Size Traits in Qinchuan Cattle (Bos taurus). Int. J. Mol. Sci. 2015, 16, 16966-16980. https://doi.org/10.3390/ijms160816966

AMA Style

Zhang Y-R, Gui L-S, Li Y-K, Jiang B-J, Wang H-C, Zhang Y-Y, Zan L-S. Molecular Characterization of Bovine SMO Gene and Effects of Its Genetic Variations on Body Size Traits in Qinchuan Cattle (Bos taurus). International Journal of Molecular Sciences. 2015; 16(8):16966-16980. https://doi.org/10.3390/ijms160816966

Chicago/Turabian Style

Zhang, Ya-Ran, Lin-Sheng Gui, Yao-Kun Li, Bi-Jie Jiang, Hong-Cheng Wang, Ying-Ying Zhang, and Lin-Sen Zan. 2015. "Molecular Characterization of Bovine SMO Gene and Effects of Its Genetic Variations on Body Size Traits in Qinchuan Cattle (Bos taurus)" International Journal of Molecular Sciences 16, no. 8: 16966-16980. https://doi.org/10.3390/ijms160816966

Article Metrics

Back to TopTop