首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
R. Robinson 《Genetica》1989,79(2):143-145
The predominant colour of the Anatolian Shepherd dog varies from a dark fawn to light red, with a variable black muzzle and face (mask). Evidence is presented that the colour is due to the dominant yellow allele (A y) of the agouti locus. Two other frequent colours are white spotting, due to the piebald allele (s p), and the chinchilla allele (ch). Two rarer colours are the agouti wolf-grey wild type (A +) and a light fawn with a blue facial mask, due to the dilution allele (d).  相似文献   

2.
A new inbred strain JF1 (Japanese Fancy Mouse 1) was established from a strain of fancy mouse. Morphological and genetical analysis indicated that the mouse originated from the Japanese wild mouse, Mus musculus molossinus. JF1 has characteristic coat color, black spots on the white coat, with black eyes. The mutation appeared to be linked to an old mutation piebald (s). Characterization of the causative gene for piebald, endothelin receptor type B (ednrb), demonstrated that the allele in JF1 is same as that of classic piebald allele, suggesting an identical origin of these two mutants. Possibly, classic piebald mutation was introduced from the Japanese tame mouse, which was already reported at the end of the 1700s. We showed that JF1 is a useful strain for mapping of mutant genes on laboratory strains owing to a high level of polymorphisms in microsatellite markers between JF1 and laboratory strains. The clarified genotypes of JF1 for coat color are ``aa BB CC DD ss'. Received: 30 May 1997 / Accepted: 26 August 1997  相似文献   

3.
4.
Domestic yaks (Bos grunniens) exhibit two major coat color variations: a brown vs. wild‐type black pigmentation and a white spotting vs. wild‐type solid color pattern. The genetic basis for these variations in color and distribution remains largely unknown and may be complicated by a breeding history involving hybridization between yaks and cattle. Here, we investigated 92 domestic yaks from China using a candidate gene approach. Sequence variations in MC1R, PMEL and TYRP1 were surveyed in brown yaks; TYRP1 was unassociated with the coloration and excluded. Recessive mutations from MC1R, or p.Gln34*, p.Met73Leu and possibly p.Arg142Pro, are reported in bovids for the first time and accounted for approximately 40% of the brown yaks in this study. The remaining 60% of brown individuals correlated with a cattle‐derived deletion mutation from PMEL (p.Leu18del) in a dominant manner. Degrees of white spotting found in yaks vary from color sidedness and white face, to completely white. After examining the candidate gene KIT, we suggest that color‐sided and all‐white yaks are caused by the serial translations of KIT (Cs6 or Cs29) as reported for cattle. The white‐faced phenotype in yaks is associated with the KIT haplotype Swf. All KIT mutations underlying the serial phenotypes of white spotting in yaks are identical to those in cattle, indicating that cattle are the likely source of white spotting in yaks. Our results reveal the complex genetic origins of domestic yak coat color as either native in yaks through evolution and domestication or as introduced from cattle through interspecific hybridization.  相似文献   

5.
White spotting phenotypes have been intensively studied in horses, and although similar phenotypes occur in the donkey, little is known about the molecular genetics underlying these patterns in donkeys. White spotting in donkeys can range from only a few white areas to almost complete depigmentation and is characterised by a loss of pigmentation usually progressing from a white spot in the hip area. Completely white‐born donkeys are rare, and the phenotype is characterised by the complete absence of pigment resulting in pink skin and a white coat. A dominant mode of inheritance has been demonstrated for spotting in donkeys. Although the mode of inheritance for the completely white phenotype in donkeys is not clear, the phenotype shows similarities to dominant white in horses. As variants in the KIT gene are known to cause a range of white phenotypes in the horse, we investigated the KIT gene as a potential candidate gene for two phenotypes in the donkey, white spotting and white. A mutation analysis of all 21 KIT exons identified a missense variant in exon 4 (c.662A>C; p.Tyr221Ser) present only in a white‐born donkey. A second variant affecting a splice donor site (c.1978+2T>A) was found exclusively in donkeys with white spotting. Both variants were absent in 24 solid‐coloured controls. To the authors’ knowledge, this is the first study investigating genetic mechanisms underlying white phenotypes in donkeys. Our results suggest that two independent KIT alleles are probably responsible for white spotting and white in donkeys.  相似文献   

6.
The conclusion that animal development is guided by a hierarchical system of gene expression and interaction has gained considerable support from recent molecular genetic studies on fruit flies (Drosophila melanogaster) and mice (Mus musculus). They demonstrate that the patterns of organization revealed by terminal differentiation of cells is anticipated by a myriad of transient prepatterns that channel the developing embryo toward its genetically-programmed target. The numerous white spotting mutants in mice exhibit some of the most dramatic and variable patterns of cutaneous melanin pigmentation. Until recently, the mechanisms of action of white spotting genes and their relationship to the developmental genetic hierarchy remained unknown. It now appears that certain white spotting genes may encode growth factors essential for melanoblast development. Others may be related to homeobox genes that play a number of developmental roles, the primary one being the determination of regional organization along the anterior-posterior axis of the early embryo. The patterns of homeobox gene expression are consistent with several of the developmental models for white spotting in mice and other mammals. It is evident that white spotting genes are not solely concerned with the terminal differentiation of melanoblasts into melanocytes. They are heterogeneous with regard to action and level of expression within the developmental hierarchy.  相似文献   

7.
White spotting phenotypes in horses are highly valued in some breeds. They are quite variable and may range from the common white markings up to completely white horses. EDNRB, KIT, MITF, PAX3 and TRPM1 represent known candidate genes for white spotting phenotypes in horses. For the present study, we investigated an American Paint Horse family segregating a phenotype involving white spotting and blue eyes. Six of eight horses with the white‐spotting phenotype were deaf. We obtained whole‐genome sequence data from an affected horse and specifically searched for structural variants in the known candidate genes. This analysis revealed a heterozygous ~63‐kb deletion spanning exons 6–9 of the MITF gene (chr16:21 503 211–21 566 617). We confirmed the breakpoints of the deletion by PCR and Sanger sequencing. PCR‐based genotyping revealed that all eight available affected horses from the family carried the deletion. The finding of an MITF variant fits well with the syndromic phenotype involving both depigmentation and an increased risk for deafness and corresponds to human Waardenburg syndrome type 2A. Our findings will enable more precise genetic testing for depigmentation phenotypes in horses.  相似文献   

8.
White spotting is the absence of melanocytes (pigment cells) from part or all of the locations in the body where they are normally found. At least in the case of the W (kit) locus, white spotting has been attributed to apoptosis. In addition to the death of melanoblasts, white spotting might result from their failure to migrate to their normal locations. These developmental failures are known to be melanocyte-specific in some instances and environment-specific in others. The environment is defined as the tissues surrounding the melanoblast. Patterns of white spotting were examined on mice mutant at the piebald (s), patch (Ph), dominant spotting (W(J2)) rumpwhite (Rw) or belted (bt) loci. The dominant spotting locus has been cloned and found to encode KIT; it has been suggested that Patch encodes the linked alpha-PDGF receptor. Piebald encodes the endothelin beta receptor. In each case, the phenotypes expressed when the allele was backcrossed onto one inbred strain C57BL/6 (B6), were compared with phenotypes expressed when the allele was backcrossed onto a different inbred strain, JU/CtLm (JU). The literature documents genetic loci that influence the extent of the white spotted area; we herein demonstrate that genetic loci also influence the location where the white spot (absence of melanocytes) will occur over the body of the mouse. Spotting occurs in a more anterior direction on JU mice that are piebald, patch or dominant-spotted compared with similar B6 mice. The relationship is reversed in rumpwhite mice, where white spotting is more anterior in the C57BL/6 mice than in the JU mice. The spotting pattern of belted mice was not modified by the background genome. Thus, the Mendelian observations indicate that several loci, which differ in JU compared with B6 mice, influence the size and the location of white spots on the mouse.  相似文献   

9.
White spotting phenotypes in horses may be caused by developmental alterations impairing melanoblast differentiation, survival, migration and/or proliferation. Candidate genes for white‐spotting phenotypes in horses include EDNRB, KIT, MITF, PAX3 and TRPM1. We investigated a German Riding Pony with a sabino‐like phenotype involving extensive white spots on the body together with large white markings on the head and almost completely white legs. We obtained whole genome sequence data from this horse. The analysis revealed a heterozygous 1273‐bp deletion spanning parts of intron 2 and exon 3 of the equine KIT gene (Chr3: 79 579 925–79 581 197). We confirmed the breakpoints of the deletion by PCR and Sanger sequencing. Knowledge of the functional impact of similar KIT variants in horses and other species suggests that this deletion represents a plausible candidate causative variant for the white‐spotting phenotype. We propose the designation W28 for the mutant allele.  相似文献   

10.
Variants in the EDNRB, KIT, MITF, PAX3 and TRPM1 genes are known to cause white spotting phenotypes in horses, which can range from the common white markings up to completely white horses. In this study, we investigated these candidate genes in 169 horses with white spotting phenotypes not explained by the previously described variants. We identified a novel missense variant, PAX3:p.Pro32Arg, in Appaloosa horses with a splashed white phenotype in addition to their leopard complex spotting patterns. We also found three novel variants in the KIT gene. The splice site variant c.1346+1G>A occurred in a Swiss Warmblood horse with a pronounced depigmentation phenotype. The missense variant p.Tyr441Cys was present in several part‐bred Arabians with sabino‐like depigmentation phenotypes. Finally, we provide evidence suggesting that the common and widely distributed KIT:p.Arg682His variant has a very subtle white‐increasing effect, which is much less pronounced than the effect of the other described KIT variants. We termed the new KIT variants W18–W20 to provide a simple and unambiguous nomenclature for future genetic testing applications.  相似文献   

11.
A male mouse with irregular white spotting, typical of piebald, s, arose during an experiment designed to search for mutations induced in spermatogonial cells by ethylnitrosourea (ENU). On being examined cytologically it was found to carry 40 chromosomes but was effectively XXY since one of the two X chromosomes present was distally fused to a Y chromosome. In common with the previously described XXY mice, all of which carried 41 chromosomes, the mouse was sterile with a total absence of germ cells. Because of this, it was not possible to determine if the white spotting was inherited. The spotting could not be related to any observable abnormality of chromosomes known to carry spotting genes, nor could it be linked in any way with the X and Y fusion. It was concluded from the cytological considerations and the time interval (6 months) that had elapsed between mutagen treatment and birth of the offspring, that whereas the spotting was probably the result of ENU damage in a spermatogonial stem cell, the XY fusion was probably a later and spontaneous event.  相似文献   

12.
A specific white spotting phenotype, termed finching or line‐backed spotting, is known for all Pinzgauer cattle and occurs occasionally in Tux‐Zillertaler cattle, two Austrian breeds. The so‐called Pinzgauer spotting is inherited as an autosomal incompletely dominant trait. A genome‐wide association study using 27 white spotted and 16 solid‐coloured Tux‐Zillertaler cattle, based on 777k SNP data, revealed a strong signal on chromosome 6 at the KIT locus. Haplotype analyses defined a critical interval of 122 kb downstream of the KIT coding region. Whole‐genome sequencing of a Pinzgauer cattle and comparison to 338 control genomes revealed a complex structural variant consisting of a 9.4‐kb deletion and an inversely inserted duplication of 1.5 kb fused to a 310‐kb duplicated segment from chromosome 4. A diagnostic PCR was developed for straightforward genotyping of carriers for this structural variant (KITPINZ) and confirmed that the variant allele was present in all Pinzgauer and most of the white spotted Tux‐Zillertaler cattle. In addition, we detected the variant in all Slovenian Cika, British Gloucester and Spanish Berrenda en negro cattle with similar spotting patterns. Interestingly, the KITPINZ variant occurs in some white spotted animals of the Swiss breeds Evolèner and Eringer. The introgression of the KITPINZ variant confirms admixture and the reported historical relationship of these short‐headed breeds with Austrian Tux‐Zillertaler and suggests a mutation event, occurring before breed formation.  相似文献   

13.
A. Wagner 《Genetica》1988,78(2):145-152
Three domestic cat populations from Yugoslavia (Bitola, Ohrid, Struga) were surveyed at six colour loci (orange, non agouti, tabby, dilute, piebald spotting, dominant white) and one coat length locus (long hair, allele l). Statistically significant differences were obtained in frequencies of alleles: O, a, l and W. Human preferences and cultural contacts with other countries especially Turkey are discussed in the paper. Genetic distance analysis did not show that Macedonian populations are more similar to Turkish or Greek ones than populations from Dalmatia, where Turkish influence was absent, although there are similarities between Macedonian and Turkish populations as far as the O allele is concerned. There was a statistically significant correlation between the duration of Turkish rule in the middle and eastern part of the Mediterranean and O and l alleles frequencies.  相似文献   

14.
The genetic basis of the white spotting pattern in Dutch boxer dogs is not known. We studied whether the segregation of white spotting in boxers follows a Mendelian inheritance pattern. Blood samples were collected, along with digital photographs in standard directions of (grand)parents (n=16) and offspring (n=52) from eight litters of Dutch boxers. In order to select heterozygous parents, we selected nonuniform litters, in which at least one puppy was extreme white. On the basis of criteria for the location, the extent of white spotting, and the mean percentage of pigmented area of the foot soles, we classified 10 dogs as solid colored, 27 as flashy, and 15 as extreme white. This was not a significant deviation from the expected 1:2:1 ratio. Because the flashy phenotype seems to be an intermediate between the two homozygotes, white spotting in the Dutch boxer can be considered to be due to a single gene effect, with incomplete dominance. We have evaluated candidate genes c-KIT (KIT) and EDNRB for segregation with white spotting phenotype in these litters. Using polymorphic markers, very near the KIT and EDNRB genes, we found that segregation of the white spotting pattern did not coincide with segregation of these polymorphic markers. Thus neither KIT nor EDNRB are likely to be responsible for white spotting in the Dutch population of boxers.  相似文献   

15.
White spotting phenotypes in horses can range in severity from the common white markings up to completely white horses. EDNRB, KIT, MITF, PAX3 and TRPM1 represent known candidate genes for such phenotypes in horses. For the present study, we re‐investigated a large horse family segregating a variable white spotting phenotype, for which conventional Sanger sequencing of the candidate genes’ individual exons had failed to reveal the causative variant. We obtained whole genome sequence data from an affected horse and specifically searched for structural variants in the known candidate genes. This analysis revealed a heterozygous ~1.9‐kb deletion spanning exons 10–13 of the KIT gene (chr3:77,740,239_77,742,136del1898insTATAT). In continuity with previously named equine KIT variants we propose to designate the newly identified deletion variant W22. We had access to 21 horses carrying the W22 allele. Four of them were compound heterozygous W20/W22 and had a completely white phenotype. Our data suggest that W22 represents a true null allele of the KIT gene, whereas the previously identified W20 leads to a partial loss of function. These findings will enable more precise genetic testing for depigmentation phenotypes in horses.  相似文献   

16.
Element content and expression of genes of interest on single cell types, such as stomata, provide valuable insights into their specific physiology, improving our understanding of leaf gas exchange regulation. We investigated how far differences in stomatal conductance (gs) can be ascribed to changes in guard cells functioning in amphistomateous leaves. gs was measured during the day on both leaf sides, on well-watered and drought-stressed trees (two Populus euramericana Moench and two Populus nigra L. genotypes). In parallel, guard cells were dissected for element content and gene expressions analyses. Both were strongly arranged according to genotype, and drought had the lowest impact overall. Normalizing the data by genotype highlighted a structure on the basis of leaf sides and time of day both for element content and gene expression. Guard cells magnesium, phosphorus, and chlorine were the most abundant on the abaxial side in the morning, where gs was at the highest. In contrast, genes encoding H+-ATPase and aquaporins were usually more abundant in the afternoon, whereas genes encoding Ca2+-vacuolar antiporters, K+ channels, and ABA-related genes were in general more abundant on the adaxial side. Our work highlights the unique physiology of each leaf side and their analogous rhythmicity through the day.  相似文献   

17.
Oxygen consumption rates of adult spring chinook salmon Oncorhynchus tshawytscha increased with swim speed and, depending on temperature and fish mass, ranged from 609 mg O2 h?1 at 30 cm s?1 (c. 0·5 BL s?1) to 3347 mg O2 h?1 at 170 cm s?1 (c. 2·3 BL s?1). Corrected for fish mass, these values ranged from 122 to 670 mg O2 kg?1 h?1, and were similar to other Oncorhynchus species. At all temperatures (8, 12·5 and 17° C), maximum oxygen consumption values levelled off and slightly declined with increasing swim speed >170 cm s?1, and a third‐order polynomial regression model fitted the data best. The upper critical swim speed (Ucrit) of fish tested at two laboratories averaged 155 cm s?1 (2·1 BL s?1), but Ucrit of fish tested at the Pacific Northwest National Laboratory were significantly higher (mean 165 cm s?1) than those from fish tested at the Columbia River Research Laboratory (mean 140 cm s?1). Swim trials using fish that had electromyogram (EMG) transmitters implanted in them suggested that at a swim speed of c. 135 cm s?1, red muscle EMG pulse rates slowed and white muscle EMG pulse rates increased. Although there was significant variation between individual fish, this swim speed was c. 80% of the Ucrit for the fish used in the EMG trials (mean Ucrit 168·2 cm s?1). Bioenergetic modelling of the upstream migration of adult chinook salmon should consider incorporating an anaerobic fraction of the energy budget when swim speeds are ≥80% of the Ucrit.  相似文献   

18.
R. Robinson 《Genetica》1990,82(2):135-137
The phenotypic interaction of three mutant genes Ba, Ds and Wh are quantatively analysed for proportion of white areas in the coat. Each of the genes individually induces a characteristic amount of white spotting which is synergistically enhanced in combination. So much so, that the genotypes Ba+Ds+Wh+ has an almost or completely white coat.  相似文献   

19.
Summary Three not closely linked genes (r 1 s , r 2 s and r 3 s ), a mutation in each of which causes UV- and X-ray-sensitivity of the colony forming ability in haploid and homozygous diploid cells have been identified. In one gene, r 3 s , two mutations have occurred independently. The effect of the mutated genes on long time repair processes as well as the effect on reversion rates from isoleucine-valine dependence to independence has been investigated. A model to explain the mechanisms changed by the mutated genes is discussed.Herrn Prof. Dr. Drs. h. c. Hans Nachtscheim zum 80. Geburtstag gewidmet.  相似文献   

20.
Within the framework of genome‐wide analyses using the novel Axiom® genotyping array, we investigated the distribution of two previously described coat color patterns, namely sabino1 (SBI), associated with the KIT gene (KI16+1037A), and splashed white, associated with the PAX3 gene (ECA6:g.11429753C>T; PAX3C70Y), including a total of 899 horses originating from eight different breeds (Achal Theke, Purebred Arabian, Partbred Arabian, Anglo‐Arabian, Shagya Arabian, Haflinger, Lipizzan and Noriker). Based on the data we collected we were able to demonstrate that, besides Quarter horses, the PAX3C70Y allele is also present in Noriker (seven out of 189) and Lipizzan (three out of 329) horses. The SB1 allele was present in three breeds (Haflinger, 14 out of 98; Noriker, four out of 189; Lipizzan one out of 329). Furthermore, we examined the phenotypes of SB1‐ and PAX3C70Y‐carrier horses for their characteristic white spotting patterns. None of the SB1/sb1‐carrier horses met the criteria defining the Sabino1 pattern according to current applied protocols. From 10 heterozygous PAX3C70Y‐carrier horses, two had nearly a splashed white phenotype. The results of this large‐scale experiment on the genetic association of white spotting patterns in horses underline the influence of gene interactions and population differences on complex traits such as Sabino1 and splashed white.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号