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 46 毫秒
1.
Estimating effective population size (Ne) using linkage disequilibrium (LD) information (Ne(LD)) has the operational advantage of using a single sample. However, Ne(LD) estimates assume discrete generations and its performance are constrained by demographic issues. However, such concerns have received little empirical attention so far. The pedigree of the endangered Gochu Asturcelta pig breed includes individuals classified into discrete filial generations and individuals with generations overlap. Up to 780 individuals were typed with a set of 17 microsatellites. Performance of Ne(LD) was compared with Ne estimates obtained using genealogical information, molecular coancestry (Ne(M)) and a temporal (two‐sample) method (Ne(JR)). Molecular‐based estimates of Ne exceeded those obtained using pedigree data. Estimates of Ne(LD) for filial generations F3 and F4 (17.0 and 17.3, respectively) were lower and steadier than those obtained using yearly or biannual samplings. Ne(LD) estimated for samples including generations overlap could only be compared with those obtained for the discrete filial generations when sampling span approached a generation interval and demographic correction for bias was applied. Single‐sample Ne(M) estimates were lower than their Ne(LD) counterparts. Ne(M) estimates are likely to partially reflect the number of founders rather than population size. In any case, estimates of LD and molecular coancestry tend to covary and, therefore, Ne(M) and Ne(LD) can hardly be considered independent. Demographically adjusted estimates of Ne(JR) and Ne(LD) took comparable values when: (1) the two samples used for the former were separated by one equivalent to discrete generations in the pedigree and (2) sampling span used for the latter approached a generation interval. Overall, the empirical evidence given in this study suggested that the advantage of using single‐sample methods to obtain molecular‐based estimates of Ne is not clear in operational terms. Estimates of Ne obtained using methods based in molecular information should be interpreted with caution.  相似文献   

2.
    
Summary The hybrid produced between a Carbondale haploid strain (-methyl-glucoside rapid fermenter) and a haploid strain (non-fermenter), derived from a hybrid between a homothallic and a heterothallicSaccharomyces, showed an irregular segregation pattern with regard to the fermentation of this sugar.To explain this irregularity, three pairs of alleles,MG 1/mg 1,MG 2/mg 2 andMG 3/mg 3, were assumed to be in quantitative control of the fermetation. Haploid cultures carrying the genotypes (1)mg 1 mg 2 mg 3, (2)MG 1 mg 2 mg 3, (3)mg 1 MG 2 mg 3, (4)mg 1 mg 2 MG 3, (5)MG 1 MG 2 mg 3, (6)MG 1 mg 2 MG 3, (7)mg 1 MG 2 MG 3, and (8)MG 1 MG 2 MG 3, were actually recovered. Strains equipped with: either (1) or (2); either (4) or (6); (3); (5); (7); or (8) are non-fermenters, extremely-slow-fermenters, slow-fermenters, medium-fermenters, semi-rapid-fermenters and rapid-fermenters respectively.The role of these genes in sugar fermentation and the identity or nonidentity of some of these genes with maltose and sucrose genes was discussed.With 2 Figures in the Text  相似文献   

3.
1. The objective was to identify the factors driving spatial and temporal variation in annual production (PA) and turnover (production/biomass) ratio (P/BA) of resident brown trout Salmo trutta in tributaries of the Rio Esva (Cantabrian Mountains, Asturias, north‐western Spain). We examined annual production (total production of all age‐classes over a year) (PA) and turnover (P/BA) ratios, in relation to year‐class production (production over the entire life time of a year‐class) (PT) and turnover (P/BT) ratio, over 14 years at a total of 12 sites along the length of four contrasting tributaries. In addition, we explored whether the importance of recruitment and site depth for spatial and temporal variations in year‐class production (PT), elucidated in previous studies, extends to annual production. 2. Large spatial (among sites) and temporal (among years) variation in annual production (range 1.9–40.3 g m?2 per year) and P/BA ratio (range 0.76–2.4 per year) typified these populations, values reported here including all the variation reported globally for salmonids streams inhabited by one or several species. 3. Despite substantial differences among streams and sites in all production attributes, when all data were pooled, annual (PA) and year‐class production (PT) and annual (P/BA) and year‐class P/BT ratios were tightly linked. Annual (PA) and year‐class production (PT) were similar but not identical, i.e. PT = 0.94 PA, whereas the P/BT ratios were 4 + P/BA ratios. 4. Recruitment (Rc) and mean annual density (NA) were major density‐dependent drivers of production and their relationships were described by simple mathematical models. While year‐class production (PT) was determined (R2 = 70.1%) by recruitment (Rc), annual production (PA) was determined (R2 = 60.3%) by mean annual density (NA). In turn, variation in recruitment explained R2 = 55.2% of variation in year‐class P/BT ratios, the latter attaining an asymptote at P/BT = 6 at progressively higher levels of recruitment. Similarly, variations in mean annual density (NA) explained R2 = 52.1% of variation in annual P/BA, the latter reaching an asymptote at P/BA = 2.1. This explained why P/BT is equal to P/BA plus the number of year‐classes at high but not at low densities. 5. Site depth was a major determinant of spatial (among sites) variation in production attributes. All these attributes described two‐phase trajectories with site depth, reaching a maximum at sites of intermediate depth and declining at shallower and deeper sites. As a consequence, at sites where recruitment and mean annual density reached minimum or maximum values, annual (PA) and year‐class production (PT) and annual (P/BA) and year‐class P/BT ratios also reached minimum and maximum values.  相似文献   

4.
Double-stranded synthetic polydeoxynucleotides of the general form poly[d(GnCn)] · poly[d(GnCn)], poly[d(GnC)] · poly[d(GCn)], and poly[d(AnTn)] · poly[d(AnTn)] have been synthesized. When n = 4 or larger, the CD spectra of polymers of the form poly[d(GnCn)] · poly[d(GnCn)] or poly[d(GnC)] · poly[d(GCn)] closely resemble the spectrum of poly[dG] · poly[dC], suggesting that a string of four continguous guanosine residues is sufficient to induce a conformation resembling that of the polypurine · polypyrimidine. With polymers of the form poly[d(AnTn)] · poly[d(AnTn)], however, the CD spectrum only gradually approaches that of poly[dA] · poly[dT].  相似文献   

5.
The parameters estimated from traditional A/C i curve analysis are dependent upon some underlying assumptions that substomatal CO2 concentration (C i) equals the chloroplast CO2 concentration (C c) and the C i value at which the A/C i curve switches between Rubisco- and electron transport-limited portions of the curve (C i-t) is set to a constant. However, the assumptions reduced the accuracy of parameter estimation significantly without taking the influence of C i-t value and mesophyll conductance (g m) on parameters into account. Based on the analysis of Larix gmelinii’s A/C i curves, it showed the C i-t value varied significantly, ranging from 24 Pa to 72 Pa and averaging 38 Pa. t-test demonstrated there were significant differences in parameters respectively estimated from A/C i and A/C c curve analysis (p<0.01). Compared with the maximum ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) carboxylation rate (Vcmax), the maximum electron transport rate (Jmax) and Jmax/Vcmax estimated from A/C c curve analysis which considers the effects of g m limit and simultaneously fits parameters with the whole A/C c curve, mean Vcmax estimated from A/C i curve analysis (Vcmax-C i) was underestimated by 37.49%; mean Jmax estimated from A/C i curve analysis (Jmax-C i) was overestimated by 17.8% and (Jmax-C i)/(Vcmax-C i) was overestimated by 24.2%. However, there was a significant linear relationship between Vcmax estimated from A/C i curve analysis and Vcmax estimated from A/C c curve analysis, so was it Jmax (p<0.05).  相似文献   

6.
Summary Breeders of self-pollinated legumes commonly use single-seed descent (SSD) or pod-bulk descent (PBD) to produce segregating populations of highly inbred individuals. We presented equations for the expected value of the additive genetic variance within populations derived by SSD (E(V A)SSD) and PBD (E(V A)PBD) in terms of the initial population size (N 0), the number of seed harvested per pod (M), the probability of survival of an individual (), and the generation at which the population is evaluated (S t). Differences between (E(V A)SSD) and (E(V A)PBD) are due to differences in the expected amount of random drift which occurs with the two methods after the S 0 generation. With both methods, random drift occurs when progeny are sampled from heterozygous parents. An additional component of random drift occurs when sampled progeny fail to survive during SSD, or when sampling occurs amoung families during PBD. For values of N 0, M, , and S t that are typical of soybean (Glycine max (L.) Merr.) breeding programs, (E(V A)SSD) will be greater than (E(V A)PBD). The ratio of (E(V A)SSD) to (E(V A)PBD) will: (1) increase as M and increase; (2) approach a value of 1.00 as N 0 increases; and (3) be a curvilinear function of S t. Plant breeders should compare SSD and PBD based upon values of (E(V A)SSD) and (E(V A)PBD) and the expected cost of carrying out the two methods.Contribution No. 2910 of the South Carolina Agricultural Experiment Station, Clemson University  相似文献   

7.
We quantified the effect of stand age and tree species composition on canopy transpiration (EC) by analysing transpiration per unit leaf area (EL) and canopy stomatal conductance (GS) for boreal trees comprising a five stand wildfire chronosequence. A total of 196 sap flux sensors were used on 90 trees consisting of Betula papyrifera Marsh (paper birch; present in the youngest stand), Populus tremuloides Michx (quaking aspen), Pinus banksiana Lamb. (jack pine), and Picea mariana (Mill.) (black spruce). While fine roots were positively correlated with stand EC; leaf area index, basal area, and sapwood area were not. Stands less than 70 years old were dominated by Populus tremuloides and Pinus banksiana and stands greater than 70 years old were composed almost entirely of Picea mariana. As Populus tremuloides and Pinus banksiana increased in size and age, they displayed an increasing sapwood to leaf area ratio (AS : AL), a constant minimum leaf water potential (ΨL), and a constant proportionality between GS at low vapour pressure deficit (Dj GSref) and the sensitivity of GS to D (–δ). In contrast, AS : AL, minimum ΨL, and the proportionally between –δ and GSref decreased with height and age in Picea mariana. A GS model that included the effects of D, AS : AL, tree height, and for Picea mariana an increasing soil to leaf water potential gradient with stand age, was able to capture the effects of contrasting hydraulic properties of Picea mariana, Populus tremuloides and Pinus banksiana during stand development after wildfire.  相似文献   

8.
Temperature responses of carbon assimilation processes were studied in four dominant species from mountain grassland ecosystem, i.e. Holcus mollis (L.), Hypericum maculatum (Cr.), Festuca rubra (L.), and Nardus stricta (L.), using the gas exchange technique. Leaf temperature (T L) of all species was adjusted within the range 13–30 °C using the Peltier thermoelectric cooler. The temperature responses of metabolic processes were subsequently modelled using the Arrhenius exponential function involving the temperature coefficient Q 10. The expected increase of global temperature led to a significant increase of dark respiration rate (R D; Q 10 = 2.0±0.5), maximum carboxylation rate (V Cmax; Q 10 = 2.2±0.6), and maximum electron transport rate (J max; Q 10 = 1.6±0.4) in dominant species of mountain grassland ecosystems. Contrariwise, the ratio between J max and V Cmax linearly decreased with T L [y = −0.884 T L + 5.24; r 2 = 0.78]. Hence temperature did not control the ratio between intercellular and ambient CO2 concentration, apparent quantum efficiency, and photon-saturated CO2 assimilation rate (P max). P max primarily correlated with maximum stomatal conductance irrespective of T L. Water use efficiency tended to decrease with T L [y = −0.21 T L + 8.1; r 2 = 0.87].  相似文献   

9.
A review of the literature revealed that a variety of methods are currently used for fitting net assimilation of CO2–chloroplastic CO2 concentration (A–Cc) curves, resulting in considerable differences in estimating the A–Cc parameters [including maximum ribulose 1·5‐bisphosphate carboxylase/oxygenase (Rubisco) carboxylation rate (Vcmax), potential light saturated electron transport rate (Jmax), leaf dark respiration in the light (Rd), mesophyll conductance (gm) and triose‐phosphate utilization (TPU)]. In this paper, we examined the impacts of fitting methods on the estimations of Vcmax, Jmax, TPU, Rd and gm using grid search and non‐linear fitting techniques. Our results suggested that the fitting methods significantly affected the predictions of Rubisco‐limited (Ac), ribulose 1,5‐bisphosphate‐limited (Aj) and TPU‐limited (Ap) curves and leaf photosynthesis velocities because of the inconsistent estimate of Vcmax, Jmax, TPU, Rd and gm, but they barely influenced the Jmax : Vcmax, Vcmax : Rd and Jmax : TPU ratio. In terms of fitting accuracy, simplicity of fitting procedures and sample size requirement, we recommend to combine grid search and non‐linear techniques to directly and simultaneously fit Vcmax, Jmax, TPU, Rd and gm with the whole A–Cc curve in contrast to the conventional method, which fits Vcmax, Rd or gm first and then solves for Vcmax, Jmax and/or TPU with Vcmax, Rd and/or gm held as constants.  相似文献   

10.
目的:检测老年住院患者分离的鲍曼不动杆菌的主要耐药基因,并研究不同耐药基因型与耐药表型之间的对应关系。方法:用PCR方法检测分离自老年住院患者的不同标本来源的170例非重复鲍曼不动杆菌的耐药基因。检测的耐药基因包括D类碳青霉烯酶:bla_(OXA-51),bla_(OXA-23),bla_(OXA-24),bla_(OXA-58),B类金属碳青霉烯酶:bla_(VIM),bla_(IMP),bla_(SIM),bla_(GIM),bla_(DIM),bla_(NDM-1),以及A类超广谱β-内酰胺酶:blaKPC,共计11种。根据检测结果对菌株进行基因分型,并研究不同基因型与CRAB和CSAB这两种耐药表型之间的对应关系。结果:170株鲍曼不动杆菌的固有基因bla_(OXA-51)均为阳性,此外,主要检出基因为bla_(OXA-23),共124株。另外检测出blaKPC12株,bla OXA-58 6株,bla_(NDM-1)3株,bla_(SIM)2株,bla_(OXA-24)、bla_(VIM)和bla_(DIM)各1株,IMP和GIM未检出。根据检出耐药基因的不同组合,分为bla OXA-51+bla_(OXA-23)阳性为基础的A型(124株)及bla_(OXA-23)阴性为基础的B型(bla_(OXA-51),39株)、C型(bla_(OXA-51)+bla_(OXA-58),6株)、D型(bla_(OXA-51)+bla_(OXA-24),1株)共计四类基因型。从耐药表型来看,128株碳青霉烯耐药菌中有122株bla_(OXA-23)为阳性,在CRAB中占95.3%(122/128),42株碳青霉烯敏感株中,有40株bla_(OXA-23)为阴性,在CSAB中占95.2%(40/42)。结论:老年病房流行的耐碳青霉烯鲍曼不动杆菌的耐药基因型以bla_(OXA-23)阳性为主。其与鲍曼不动杆菌CRAB耐药表型、bla_(OXA-23)阴性与CSAB耐药表型之间有良好的对应关系。  相似文献   

11.
Sex differences in skews of vertebrate lifetime reproductive success are difficult to measure directly. Evolutionary histories of differential skew should be detectable in the genome. For example, male‐biased skew should reduce variation in the biparentally inherited genome relative to the maternally inherited genome. We tested this approach in lek‐breeding ruff (Class Aves, Philomachus pugnax) by comparing genetic variation of nuclear microsatellites (θn; biparental) versus mitochondrial D‐loop sequences (θm; maternal), and conversion to comparable nuclear (Ne) and female (Nef) effective population size using published ranges of mutation rates for each marker (μ). We provide a Bayesian method to calculate Ne (θn = 4Neμn) and Nef (θm = 2Nefμm) using 95% credible intervals (CI) of θn and θm as informative priors, and accounting for uncertainty in μ. In 96 male ruffs from one population, Ne was 97% (79–100%) lower than expected under random mating in an ideal population, where Ne:Nef = 2. This substantially lower autosomal variation represents the first genomic support of strong male reproductive skew in a lekking species.  相似文献   

12.
A novel framework is presented for the analysis of ecophysiological field measurements and modelling. The hypothesis ‘leaves minimise the summed unit costs of transpiration and carboxylation’ predicts leaf‐internal/ambient CO2 ratios (ci/ca) and slopes of maximum carboxylation rate (Vcmax) or leaf nitrogen (Narea) vs. stomatal conductance. Analysis of data on woody species from contrasting climates (cold‐hot, dry‐wet) yielded steeper slopes and lower mean ci/ca ratios at the dry or cold sites than at the wet or hot sites. High atmospheric vapour pressure deficit implies low ci/ca in dry climates. High water viscosity (more costly transport) and low photorespiration (less costly photosynthesis) imply low ci/ca in cold climates. Observed site‐mean ci/ca shifts are predicted quantitatively for temperature contrasts (by photorespiration plus viscosity effects) and approximately for aridity contrasts. The theory explains the dependency of ci/ca ratios on temperature and vapour pressure deficit, and observed relationships of leaf δ13C and Narea to aridity.  相似文献   

13.
Summary Using a direct Monte Carlo simulation, population growth of helper T-cells (N H) and viral cells (N v) is studied for an immune response model with an enhanced spatial inter-cellular interaction relevant to HIV as a function of viral mutation. In the absence of cellular mobility (P mob=0), the helper T-cells grow nonmonotonically before reaching saturation and the viral population grows monotonically before reaching a constant equilibrium. Cellular mobility (P mob=1) enhances the viral growth and reduces the stimulative T-cell growth. Below a mutation threshold (P c), the steady-state density of helper T-cell (p H) is larger than that of the Virus (p v); the density difference Δp o(=pV−pH) remains a constant at P mob=1 while −Δp o→0 as P mutP c at P mob=0. Above the mutation threshold, the difference Δp o in cell density, grows with ΔP=P mutP c monotonically: ΔP o ∞ (ΔP)β ≃ with β≈0.574±0.016 in absence of mobility, while Δp o≈6(ΔP) with P mob=1.  相似文献   

14.
为探究岩溶植物的光合生理适应机制,采用Li-6400XT便携式光合作用测量系统,对广西平果市岩溶区8种适生植物的叶片净光合速率(Pn)、气孔导度(Gs)、胞间CO2浓度(Ci)、蒸腾速率(Tr)、水分利用效率(WUE)和气孔限制值(Ls)等光合特征参数进行了测定分析。结果表明:(1)6个光合特征参数在种内和种间均存在不同程度的变异,并且种内变异均大于种间变异。(2)Gs和Tr的变化主要来源于种间变异(46.72%~49.76%),而Pn、Ci、WUE和Ls变化主要来源于种内变异(48.66%~64.50%)。在生活型水平上,Pn、Gs和Tr的种内变异表现为常绿植物小于落叶植物,而Ci、WUE和Ls则相反。(3)各参数的种间变异均表现为落叶植...  相似文献   

15.
By investigating the R D-C a (dark respiration rate-atmospheric CO2 concentration) and P N (net photosynthetic rate)-C a curves of bamboo (Fargesia denudata) and poplar (Populus cathayanna), we found that: (1) the minimal R D was close to ambient CO2 concentration, and the elevated or decreased atmospheric CO2 concentration enhanced the R D of both species; (2) the response curves of R D-C a were simulated well by quadratic function. This phenomenon might be an inherent property of leaf R D of F. denudata and P. cathayanna. If this was true, it implies that effect of CO2 on R D could be interpreted with the relationship of R D-C a curves and the quadratic function.  相似文献   

16.
C. Fu    D. Li    W. Hu    Y. Wang  † Z. Zhu   《Journal of fish biology》2007,70(2):347-361
The growth and energy budget for F2‘all‐fish’ growth hormone gene transgenic common carp Cyprinus carpio of two body sizes were investigated at 29·2° C for 21 days. Specific growth rate, feed intake, feed efficiency, digestibility coefficients of dry matter and protein, gross energy intake (IE), and the proportion of IE utilized for heat production (HE) were significantly higher in the transgenics than in the controls. The proportion of IE directed to waste products [faecal energy (FE) and excretory energy loss (ZE+UE) where ZE is through the gills and UE through the kidney], and the proportion of metabolizable energy (ME) for recovered energy (RE) were significantly lower in the transgenics than in the controls. The average energy budget equation of transgenic fish was as follows: 100 IE= 19·3 FE+ 6·0 (ZE+UE) + 45·2 HE+ 29·5 RE or 100 ME= 60·5 HE+ 39·5 RE. The average energy budget equation of the controls was: 100 IE= 25·2 FE+ 7·4 (ZE+UE) + 35·5 HE+ 31·9 RE or 100 ME= 52·7 HE+ 47·3 RE. These findings indicate that the high growth rate of ‘all‐fish’ transgenic common carp relative to their non‐transgenic counterparts was due to their increased feed intake, reduced lose of waste productions and improved feed efficiency. The benefit of the increased energy intake by transgenic fish, however, was diminished by their increased metabolism.  相似文献   

17.
Short-chain-length medium-chain-length polyhydroxyalkanoate (SCL-MCL PHA) copolymers are promising as bio-plastics with properties ranging from thermoplastics to elastomers. In this study, the hybrid pathway for the biosynthesis of SCL-MCL PHA copolymers was established in recombinant Escherichia coli by co-expression of β-ketothiolase (PhaA Re ) and NADPH-dependent acetoacetyl-CoA reductase (PhaB Re ) from Ralstonia eutropha together with PHA synthases from R. eutropha (PhaC Re ), Aeromonas hydrophila (PhaC Ah ), and Pseudomonas putida (PhaC2 Pp ) and with (R)-specific enoyl-CoA hydratases from P. putida (PhaJ1 Pp and PhaJ4 Pp ), and A. hydrophila (PhaJ Ah ). When glycerol supplemented with dodecanoate was used as primary carbon source, E. coli harboring various combinations of PhaABCJ produced SCL-MCL PHA copolymers of various monomer compositions varying from C4 to C10. In addition, polymer property analysis suggested that the copolymers produced from this recombinant source have thermal properties (lower glass transition and melting temperatures) superior to polyhydroxybutyrate homopolymer.  相似文献   

18.
We investigate the utility of an improved isotopic method to partition the net ecosystem exchange of CO2 (F) into net photosynthesis (FA) and nonfoliar respiration (FR). Measurements of F and the carbon isotopic content in air at a high‐elevation coniferous forest (the Niwot Ridge AmeriFlux site) were used to partition F into FA and FR. Isotopically partitioned fluxes were then compared with an independent flux partitioning method that estimated gross photosynthesis (GEE) and total ecosystem respiration (TER) based on statistical regressions of night‐time F and air temperature. We compared the estimates of FA and FR with expected canopy physiological relationships with light (photosynthetically active radiation) and air temperature. Estimates of FA and GEE were dependent on light as expected, and TER, but not FR, exhibited the expected dependence on temperature. Estimates of the isotopic disequilibrium D , or the difference between the isotopic signatures of net photosynthesis (δA, mean value ?24.6‰) and ecosystem respiration (δR, mean value ?25.1‰) were generally positive (δAR). The sign of D observed here is inconsistent with many other studies. The key parameters of the improved isotopic flux partitioning method presented here are ecosystem scale mesophyll conductance (gm) and maximal vegetative stomatal conductance (gcmax). The sensitivity analyses of FA, FR, and D to gcmax indicated a critical value of gcmax (0.15 mol m?2 s?1) above which estimates of FA and FR became larger in magnitude relative to GEE and TER. The value of D decreased with increasing values of gm and gcmax, but was still positive across all values of gm and gcmax. We conclude that the characterization of canopy‐scale mesophyll and stomatal conductances are important for further progress with the isotope partitioning method, and to confirm our anomalous isotopic disequilibrium findings.  相似文献   

19.
S Kubota  K Ikeda  J T Yang 《Biopolymers》1983,22(10):2237-2252
A series of sequential polypeptides (LysiRj)n (R is Leu, Ser, or Gly) and random copolypeptides, (Lysx, Leuy)n, were synthesized. Their conformation in NaDodSO4 solution was determined by CD. Only (Lys-Leu)n, (Lys-Ser)n, and (Lys3-Ser)n adopt a stable β-form in the surfactant solution; (Lys-Ser2)n, (Lys-Ser3)n, (Lys2-Ser2)n, and (Lys2-Ser)n have an unstable β-form, which reverts to an unordered form in high NaDodSO4 concentrations, even though both Ser and DodSO-bound Lys+ are β-formers. In contrast, (Lys-Gly)n remains unordered in NaDodSO4 solution. On the other hand, Lys-rich (Lys2-Leu)n forms an unstable helix and (Lys2-Leu2)n a stable helix in NaDodSO4 solution. In 25 mM NaDodSO4 (Lysx, Leuy)n also forms a helix up to x = 75 and reverts to the β-form at x = 90. This compares with the helical conformation of (Lysx, Alay)n up to x = 65 and its β-form at x = 90, suggesting that Leu is an even stronger helix-former than Ala. Our results may provide a plausible explanation for the increase in helicity and disruption of the β-form for many proteins in NaDodSO4 solution, that is, the polypeptide chain of a protein usually favors a helical conformation over a β-form in the presence of excess surfactant.  相似文献   

20.
Based on a Cambridge Structural Database (CSD) search, a meta‐analysis of 116 structures of alanine H3NCαH(CH3)C′(O)O and its derivatives H3NCαH(CH3)C′(O)O(H/R/M), protonated, esterified, or coordinated at the carboxylic group, shows that in the first step of a chirality chain, the L configuration at Cα induces (M) and (P) conformations with respect to rotation around the central C′─Cα bond. In the second step, the (M) and (P) conformations selectively distort the planar carboxylic group CαC'(Ocis)Otrans to asymmetric flat (R) and (S) tetrahedra. High diastereoselectivities are caused by the two players attraction N…Ocis and repulsion Otrans…CMe, which work together in (L,M,R) configurations but against each other in (L,P,S) configurations.  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