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.
Zeins are the major storage proteins in corn seeds organized in protein bodies located in the endosperm. They are soluble in alcoholic solution and depict a high tendency to aggregation. The Z19 α-zein aggregates obtained by heating show a particular and interesting temperature-dependent behavior. This work was aimed at determining not only the effect of temperature on the aggregation behavior, but also the effect of the sulfhydryl groups and disulphide bonds on the thermal aggregation process under non-aqueous conditions. Z19 α-zein was chemically modified to obtain different sulfhydryl groups and disulphide-bonds content. Far-UV CD, ANS emission fluorescence, and dynamic light scattering, as well as differential scanning calorimetry, were performed to characterize this protein. Removal of these disulphide-bonds and alkylation of all the sulfhydryl groups in the protein promoted the lowest Tm of 57.36 °C, eliminated aggregation, enhanced protein flexibility, and diminished thermal stability. These results suggest that the disulphide linkage could be the driving force for the Z19 α-zein aggregation.  相似文献   

2.
The genetic linkage relationships of the human glycosphingolipid beta-galactosidases were determined using human--mouse somatic cell hybrids. A new method was devised for the estimation of human galactosylceramide, lactosylceramide, and GMI-ganglioside beta-galactosidase activities in the presence of their mouse counterparts, which takes advantage of the reproducible specific activity of lysosomal hydrolases under a given set of culture conditions and is based on differences in both pH optima and sensitivity to chloride ion. Human and mouse chromosomes were identified by their characteristic banding patterns obtained after quinacrine staining, and the optimum glycolipid beta-galactosidase activity was determined for three different substrates. A ratio was defined for each activity which was the specific activity at the human pH optimum divided by the specific activity at the mouse pH optimum. Linear regression analysis was used to test for concordant segregation between pH ratios for each enzyme and the frequency of occurrence of different human chromosomes in the man--mouse somatic hybrid clones. The results obtained from two independent series of hybrid clones indicated that human beta-galactosidase activities consistently segregated with human chromosome 12 in these somatic cell hybrids.  相似文献   

3.
β2-Microglobulin (β2M) is believed to have arisen in a basal jawed vertebrate (gnathostome) and is the essential L chain that associates with most MHC class I molecules. It contains a distinctive molecular structure called a constant-1 Ig superfamily domain, which is shared with other adaptive immune molecules including MHC class I and class II. Despite its structural similarity to class I and class II and its conserved function, β2M is encoded outside the MHC in all examined species from bony fish to mammals, but it is assumed to have translocated from its original location within the MHC early in gnathostome evolution. We screened a nurse shark bacterial artificial chromosome library and isolated clones containing β2M genes. A gene present in the MHC of all other vertebrates (ring3) was found in the bacterial artificial chromosome clone, and the close linkage of ring3 and β2M to MHC class I and class II genes was determined by single-strand conformational polymorphism and allele-specific PCR. This study satisfies the long-held conjecture that β2M was linked to the primordial MHC (Ur MHC); furthermore, the apparent stability of the shark genome may yield other genes predicted to have had a primordial association with the MHC specifically and with immunity in general.  相似文献   

4.
 We constructed a genetic map of a cross between the Swiss winter wheat (Triticum aestivum L.) variety Forno and the Swiss winter spelt (Triticum spelta L.) variety Oberkulmer. For the linkage analysis,176 polymorphic RFLP probes and nine microsatellites were tested on 204 F5 recombinant inbred lines (RILs) of Forno×Oberkulmer revealing 242 segregating marker loci. Thirty five percent of these loci showed significant (P>0.05) deviation from a 1 : 1 segregation, and the percentage of Forno alleles ranged from 21% to 83% for individual marker loci. Linkage analysis was performed with the program MAPMAKER using the Haldane mapping function. Using a LOD threshold of 10, we obtained 37 linkage groups. After finding the best order of marker loci within linkage groups by multi-point analysis we assembled the linkage groups into 23 larger units by lowering the LOD threshold. All except one of the 23 new linkage groups could be assigned to physical chromosomes or chromosome arms according to hybridisation patterns of nulli-tetrasomic lines of Chinese Spring and published wheat maps. This resulted in a genetic map comprising 230 marker loci and spanning 2469 cM. Since the analysed population is segregating for a wide range of agronomically important traits, this genetic map is an ideal basis for the identification of quantitative trait loci (QTLs) for these traits. Received: 3 August 1998 / Accepted: 28 November 1998  相似文献   

5.
The present report is dealing with the identification, in various unrelated proteins, of protein fragments sharing local sequence and structure similarities with the chymosin-sensitive linkage surrounding the Phe-Met/Ile bond of κ-caseins. In all these proteins, this linkage is observed within an exposed β-strand-like structure, as also predicted for κ-caseins. The structure of one of these fragments, included in glutamine synthetase, particularly superimposes well with the conformation observed for a chymosin inhibitor (CP-113972) within the complex it forms with chymosin and can be similarly accommodated by specificity pockets within the enzyme substrate binding cleft. The effect of the enzyme activity of chymosin was thus tested on glutamine synthetase. Chymosin cut the latter at the Phe-Met linkage, suggesting that this system may locally resemble the κ-casein/chymosin complex.  相似文献   

6.
Two previously described QTL mapping methods, which combine linkage analysis (LA) and linkage disequilibrium analysis (LD), were compared for their ability to detect and map multiple QTL. The methods were tested on five different simulated data sets in which the exact QTL positions were known. Every simulated data set contained two QTL, but the distances between these QTL were varied from 15 to 150 cM. The results show that the single QTL mapping method (LDLA) gave good results as long as the distance between the QTL was large (> 90 cM). When the distance between the QTL was reduced, the single QTL method had problems positioning the two QTL and tended to position only one QTL, i.e. a "ghost" QTL, in between the two real QTL positions. The multi QTL mapping method (MP-LDLA) gave good results for all evaluated distances between the QTL. For the large distances between the QTL (> 90 cM) the single QTL method more often positioned the QTL in the correct marker bracket, but considering the broader likelihood peaks of the single point method it could be argued that the multi QTL method was more precise. Since the distances were reduced the multi QTL method was clearly more accurate than the single QTL method. The two methods combine well, and together provide a good tool to position single or multiple QTL in practical situations, where the number of QTL and their positions are unknown.  相似文献   

7.
The low-energy orientational oscillations of the peptide groups of an -helix are considered and the value of the frequency is estimated to be in agreement with experiments. Approximate formulae are derived for the projection of a dipole moment on the helix axis and for the helix parameters. Within the framework of a three-chain model, the asymptotics of the soliton solution is obtained using a discrete approach.The analysis of -helix geometry exhibits two types of low-frequency oscillations of the -helix. The first one is connected with atom movements along the helix axis with the peptide groups twisting around the helix axis. Accordingly, it changes the hydrogen bond lengths between neighbouring peptide groups. In the second case, the slopes of the peptide groups to the helix axis oscillate without the helix parameters changing. Here, the energy of interactions between peptide-group dipoles is changed and, as a result, the oscillations have an optical nature. The frequency of the optical orientational oscillations is approximately 100 cm-1.  相似文献   

8.
Annual (Lolium multiflorum Lam.) and perennial (L. perenne L.) ryegrass are two common forage and turfgrass species grown throughout the world. Perennial ryegrass is most commonly used for turfgrass purposes, and contamination by annual ryegrass, through physical seed mixing or gene flow, can result in a significant reduction in turfgrass quality. Seed certifying agencies in the United States currently use a test called seedling root fluorescence (SRF) to detect contamination between these species. The SRF test, however, can be inaccurate and therefore, the development of additional markers for species separation is needed. Male and female molecular-marker linkage maps of an interspecific annual × perennial ryegrass mapping population were developed to determine the map location of the SRF character and to identify additional genomic regions useful for species separation. A total of 235 AFLP markers, 81 RAPD markers, 16 comparative grass RFLPs, 106 SSR markers, 2 isozyme loci and 2 morphological characteristics, 8-h flowering, and SRF were used to construct the maps. RFLP markers from oat and barley and SSR markers from tall fescue and other grasses allowed the linkage groups to be numbered, relative to the Triticeae and the International Lolium Genome Initative reference population P150/112. The three-generation population structure allowed both male and female maps to be constructed. The male and female maps each have seven linkage groups, but differ in map length with the male map being 537 cm long and the female map 712 cm long. Regions of skewed segregation were identified in both maps with linkage groups 1, 3, and 6 of the male map showing the highest percentage of skewed markers. The (SRF) character mapped to linkage group 1 in both the male and female maps, and the 8-h flowering character was also localized to this linkage group on the female map. In addition, the Sod-1 isozyme marker, which can separate annual and perennial ryegrasses, mapped to linkage group 7. These results indicate that Lolium linkage groups 1 and 7 may provide additional markers and candidate genes for use in ryegrass species separation.Communicated by C. Möllers  相似文献   

9.
The relationship between linkage disequilibrium (LD) and recombination fraction can be used to infer the pattern of genetic variation and evolutionary process in humans and other systems. We described a computational framework to construct a linkage–LD map from commonly used biallelic, single-nucleotide polymorphism (SNP) markers for outcrossing plants by which the decline of LD is visualized with genetic distance. The framework was derived from an open-pollinated (OP) design composed of plants randomly sampled from a natural population and seeds from each sampled plant, enabling simultaneous estimation of the LD in the natural population and recombination fraction due to allelic co-segregation during meiosis. We modified the framework to infer evolutionary pasts of natural populations using those marker types that are segregating in a dominant manner, given their role in creating and maintaining population genetic diversity. A sophisticated two-level EM algorithm was implemented to estimate and retrieve the missing information of segregation characterized by dominant-segregating markers such as single methylation polymorphisms. The model was applied to study the relationship between linkage and LD for a non-model outcrossing species, a gymnosperm species, Torreya grandis, naturally distributed in mountains of the southeastern China. The linkage–LD map constructed from various types of molecular markers opens a powerful gateway for studying the history of plant evolution.  相似文献   

10.
Protein aggregates can be stabilised by disulphide bridges. The whey protein β-lactoglobulin (β-lac) contains a disulphide bridge and a free cysteine that are shielded from the solvent by an α-helix. These groups are important in the thiol–disulphide exchange that occurs during aggregation and gelation of β-lac. Replica exchange molecular dynamics simulations show that the exposure mechanism is very different for the two buried groups. While melting of the α-helix enhances exposure of the free cysteine, it does not for the buried bridge. These findings shed light on the molecular mechanism of the first step of β-lac denaturation and aggregation.  相似文献   

11.
As genome and cDNA sequencing projects progress, a tremendous amount of sequence information is becoming publicly available. These sequence resources can be exploited for gene discovery and marker development. Simple sequence repeat (SSR) markers are among the most useful because of their great variability, abundance, and ease of analysis. By in silico analysis of 10,232 non-redundant expressed sequence tags (ESTs) in pepper as a source of SSR markers, 1,201 SSRs were found, corresponding to one SSR in every 3.8 kb of the ESTs. Eighteen percent of the SSR–ESTs were dinucleotide repeats, 66.0% were trinucleotide, 7.7% tetranucleotide, and 8.2% pentanucleotide; AAG (14%) and AG (12.4%) motifs were the most abundant repeat types. Based on the flanking sequences of these 1,201 SSRs, 812 primer pairs that satisfied melting temperature conditions and PCR product sizes were designed. 513 SSRs (63.1%) were successfully amplified and 150 of them (29.2%) showed polymorphism between Capsicum annuum ‘TF68’ and C. chinense ‘Habanero’. Dinucleotide SSRs and EST–SSR markers containing AC-motifs were the most polymorphic. Polymorphism increased with repeat length and repeat number. The polymorphic EST–SSRs were mapped onto the previously generated pepper linkage map, using 107 F2 individuals from an interspecific cross of TF68 × Habanero. One-hundred and thirtynine EST–SSRs were located on the linkage map in addition to 41 previous SSRs and 63 RFLP markers, forming 14 linkage groups (LGs) and spanning 2,201.5 cM. The EST–SSR markers were distributed over all the LGs. This SSR-based map will be useful as a reference map in Capsicum and should facilitate the use of molecular markers in pepper breeding.Gibum Yi and Je Min Lee equally contributed to this work.  相似文献   

12.
What chance did Mendel's experiments give him of noticing linkage?   总被引:1,自引:0,他引:1  
L Douglas  E Novitski 《Heredity》1977,38(2):253-257
The a priori probability of noticeable linkage among all conceivable experiments of the size reported by Mendel cannot reasonably be taken as greater than 24-36 per cent; and therefore, the frequently heard opinion that his chances of encountering linkage were high, approaching 99-4 per cent, appears to be mistaken.  相似文献   

13.
Aminoacyl-tRNAs are the biologically active substrates for peptide bond formation in protein synthesis. The stability of the acyl linkage in each aminoacyl-tRNA, formed through an ester bond that connects the amino acid carboxyl group with the tRNA terminal 3′-OH group, is thus important. While the ester linkage is the same for all aminoacyl-tRNAs, the stability of each is not well characterized, thus limiting insight into the fundamental process of peptide bond formation. Here, we show, by analysis of the half-lives of 12 of the 22 natural aminoacyl-tRNAs used in peptide bond formation, that the stability of the acyl linkage is effectively determined only by the chemical nature of the amino acid side chain. Even the chirality of the side chain exhibits little influence. Proline confers the lowest stability to the linkage, while isoleucine and valine confer the highest, whereas the nucleotide sequence in the tRNA provides negligible contribution to the stability. We find that, among the variables tested, the protein translation factor EF-Tu is the only one that can protect a weak acyl linkage from hydrolysis. These results suggest that each amino acid plays an active role in determining its own stability in the acyl linkage to tRNA, but that EF-Tu overrides this individuality and protects the acyl linkage stability for protein synthesis on the ribosome.  相似文献   

14.
15.
16.
Summary The DNA·DNA duplex ·d(GCGCAAAACGCG) (designated duplex III) containing a 3-thioformacetal (3-TFMA) linkage in the center of the sequence was characterized in detail by two- and three-dimensional homonuclear NMR spectroscopy. The NMR results were analyzed and compared with those of two duplexes of the same sequence: One is an unmodified reference sequence and the other contains a formacetal (OCH2O) linkage at the central T^T step (designated duplex I and duplex II, respectively). In general, the NMR spectra of duplex III closely resemble those of the analogous duplexes I and II, suggesting an overall B-type structure adopted by the 3-TFMA-modified duplex III. Nonetheless, the detection of several distinct spectral features originating from the protons at the modification site is indicative of a local conformation that is clearly different from the corresponding region in duplexes I and II. The 3-thioformacetal linker, in contrast to the formacetal (FMA) linkage, cannot be accommodated in a conformation usually found in natural nucleic acid duplexes. As a consequence, the 3-TFMA-modified T6 sugar adopts an O4-endo form (an intermediate structure between the usual C2-endo and C3-endo forms). This change is accompanied by a change in the (C4–C3–S3–CH2) dihedral angle and by subsequent adjustments of other torsion angles along the backbone. Notably, this conformational readjustment at the T6–T7 backbone linkage is localized; its collective result has negligible effect on base-base stacking of the T6 and T7 residues. A close examination of the COSY data in all three duplexes reveals a subtle variation in sugar geometry, with more S-type character adopted by the modified duplexes II and III. The results of this study illustrate that, although the difference between FMA and 3-TFMA linkages is merely in the substitution of the T6(O3) in the former by a sulfur atom in the latter, the stereoelectronic difference in a single atom can induce significant local structural distortion in an otherwise well-structured oligonucleotide duplex.Supplementary material available from the authors: One table containing J12, J12 and J34 of duplexes I, II and III.  相似文献   

17.
This study demonstrates the relative ease of generating high-density linkage maps using the AFLP® technology. Two high-density AFLP linkage maps of Zea mays L. were generated based on: (1) a B73 × Mo17 recombinant inbred population and (2) a D32 × D145 immortalized F2 population. Although AFLP technology is in essence a mono-allelic marker system, markers can be scored quantitatively and used to deduce zygosity. AFLP markers were generated using the enzyme combinations EcoRI/MseI and PstI/MseI. A total of 1539 and 1355 AFLP markers have been mapped in the two populations, respectively. Among the mapped PstI/MseIAFLP markers we have included fragments bounded by a methylated PstI site (mAFLP markers). Mapping these mAFLP markers shows that the presence of C-methylation segregates in perfect accordance with the primary target sequence, leading to Mendelian inheritance. Simultaneous mapping of PstI/MseIAFLP and PstI/MseI mAFLP markers allowed us to identify a number of epi-alleles, showing allelic variation in the CpNpG methylation only. However, their frequency in maize is low. Map comparison shows that, despite some rearrangements, most of the AFLP markers that are common in both populations, map at similar positions. This would indicate that AFLP markers are predominantly single-locus markers. Changes in map order occur mainly in marker-dense regions. These marker-dense regions, representing clusters of mainly EcoRI/MseI AFLP and PstI/MseI mAFLP markers, co- localize well with the putative centromeric regions of the maize chromosomes. In contrast, PstI/MseImarkers are more uniformly distributed over the genome.  相似文献   

18.
We report a dense genetic linkage map of Heliconius erato, a neotropical butterfly that has undergone a remarkable adaptive radiation in warningly colored mimetic wing patterns. Our study exploited natural variation segregating in a cross between H. erato etylus and H. himera to localize wing color pattern loci on a dense linkage map containing amplified fragment length polymorphisms (AFLP), microsatellites, and single-copy nuclear loci. We unambiguously identified all 20 autosomal linkage groups and the sex chromosome (Z). The map spanned a total of 1430 Haldane cM and linkage groups varied in size from 26.3 to 97.8 cM. The average distance between markers was 5.1 cM. Within this framework, we localized two major color pattern loci to narrow regions of the genome. The first gene, D, responsible for red/orange elements, had a most likely placement in a 6.7-cM region flanked by two AFLP markers on the end of a large 87.5-cM linkage group. The second locus, Sd, affects the melanic pattern on the forewing and was found within a 6.3-cM interval between flanking AFLP loci. This study complements recent linkage analysis of H. erato's comimic, H. melpomene, and forms the basis for marker-assisted physical mapping and for studies into the comparative genetic architecture of wing-pattern mimicry in Heliconius.  相似文献   

19.
Can divergence in a mating trait increase local adaption by increasing ecological divergence? Servedio and Bürger propose that “pseudomagic traits,” tightly linked complexes consisting of an ecological locus under divergent selection and a locus acting as a mating cue, can effectively mimic pleiotropy. Such pseudomagic traits can form even when linkage between ecological and mating loci is limited.  相似文献   

20.
Frequent loss of heterogeneity in prostate cancer cells and linkage studies of families affected by hereditary prostate cancer (HPC) have implied that the short arm of chromosome 8, specifically 8p22-23, may harbor a prostate-cancer-susceptibility gene. In a recent study, seven potentially important mutations in the macrophage scavenger receptor 1 gene (MSR1), located at 8p22, were observed in families affected with HPC, and an indication of co-segregation between these mutations and prostate cancer was reported. In an attempt to confirm linkage at 8p22-23, we performed linkage analyses in 57 families affected with HPC (ascertained throughout Sweden) by using 13 markers on the short arm of chromosome 8. In the complete set of families, evidence for prostate cancer linkage was observed at 8p22-23, with a peak hold of 1.08 (P=0.03), observed at D8S1731, approximately 1 cM centromeric to the MSR1 gene. At marker D8S1135, the closest marker to MSR1, a hlod of 1.07 (P=0.03) was observed. Evidence of linkage was seen in families with early-onset HPC and in families with a small number of affected individuals. The peak multipoint non-parametric linkage score was 2.01 (P=0.03) at D8S552 in the 14 pedigrees with mean age at onset <65 years, and 2.25 (P=0.01) at D8S1731 in the 36 pedigrees with fewer than five affected family members. Thus, we have confirmed evidence for prostate cancer linkage at 8p22-23. Follow-up studies to evaluate the possible association between prostate cancer and genes in this region, especially the MSR1 gene, are warranted.  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