首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   81篇
  免费   4篇
  2022年   1篇
  2015年   4篇
  2014年   10篇
  2013年   5篇
  2012年   1篇
  2011年   3篇
  2010年   7篇
  2009年   4篇
  2008年   5篇
  2007年   1篇
  2006年   5篇
  2005年   2篇
  2004年   1篇
  2003年   5篇
  2002年   3篇
  2001年   2篇
  2000年   4篇
  1999年   5篇
  1995年   1篇
  1992年   1篇
  1991年   1篇
  1990年   4篇
  1989年   1篇
  1982年   1篇
  1978年   1篇
  1977年   1篇
  1974年   2篇
  1972年   2篇
  1971年   1篇
  1967年   1篇
排序方式: 共有85条查询结果,搜索用时 15 毫秒
1.
Three amino acid residues in the active site of lipoamide dehydrogenase from Azotobacter vinelandii were replaced with other residues. His450, the active-site base, was replaced with Ser, Tyr or Phe. Pro451, from X-ray analysis found to be in cis conformation positioning the backbone carbonyl of His450 close to N3 of the flavin, was changed to Ala. Glu455, from X-ray analysis expected to be involved in modulating the pKa of the base (His450), was replaced with Asp and Gln. The general conclusion is that mutation of the His-Glu diad impairs intramolecular electron transfer between the disulfide/dithiol and the FADH-/FAD. The wild-type enzyme functions according to a ping-pong mechanism in the physiological reaction in which the formation of NADH is rate-limiting. Above pH 8.0 the enzyme is strongly inhibited by the product NADH. The pH dependence of the steady-state kinetics using the NAD+ analog 3-acetylpyridine adenine dinucleotide (AcPyAde+) reveals a pKa of 8.1 in the pKm AcPyAde+ plot indicating that this pKa is related to the deprotonation of His450 [Benen, J., Berkel van, W., Zak, Z., Visser, T., Veeger, C. & Kok de, A. (1991) Eur. J. Biochem. 202, 863-872] and to the inhibition by NADH. The mutations considerably affect turnover. Enzymes with the mutations Pro451----Ala, His450----Phe and His450----Tyr appear to be almost inactive in both directions. Enzyme His450----Ser is minimally active, V at the pH optimum being 0.5% of wild-type activity in the physiological reaction. Rapid reaction kinetics show that for the His450-mutated enzymes the reductive half reaction using reduced 6,8-thioctic acid amide [Lip(SH)2] is rate-limiting and extremely slow when compared using reduced 6,8-thioctic acid amide [Lip(SH)2] is rate-limiting and extremely slow when compared to the wild-type enzyme. For enzyme Pro451----Ala it is concluded that the loss of activity is due to over-reduction by Lip(SH)2 and NADH. The Glu455-mutated enzymes are catalytically competent but show strong inhibition by the product NADH (enzyme Glu455----Asp more than Glu455----Gln). The inhibition can largely be overcome by using AcPyAde+ instead of NAD+ in the physiological reaction. The rapid reaction kinetics obtained for enzymes Glu455----Asp and Glu455----Gln deviate from the wild-type enzyme. It is concluded that this difference is due to cooperativity between the active sites in this dimeric enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   
2.
Muscarinic activation of tracheal smooth muscle (TSM) involves a M3AChR/heterotrimeric-G protein/NPR-GC coupling mechanism. G protein activators Mastoparan (MAS) and Mastoparan-7 stimulated 4- and 10-fold the NPR-GC respectively, being insensitive to PTX and antibodies against Gαi/o subfamily. Muscarinic and MAS stimulation of NPR-GC was blocked by antibodies against C-terminal of Gαq16, whose expression was confirmed by RT-PCR. However, synthetic peptides from C-terminal of Gαq15/16 stimulated the NPR-GC. Coupling of αq16 to M3AChR is supported by MAS decreased [3H]QNB binding, being abolished after M3AChR-4-DAMP-alkylation. Anti-i3M3AChR antibodies blocked the muscarinic activation of NPR-GC, and synthetic peptide from i3M3AChR (M3P) was more potent than MAS increasing GTPγ [35S] and decreasing the [3H]QNB activities. Coupling between NPR-GC and Gαq16 was evaluated by using trypsin-solubilized-fraction from TSM membranes, which displayed a MAS-sensitive-NPR-GC activity, being immunoprecipitated with anti-Gαq16, also showing an immunoreactive heterotrimeric-G-β -subunit. These data support the existence of a novel transducing cascade, involving Gαq16β γ coupling M3AChR to NPR-GC.  相似文献   
3.

Background

A major hindrance to the development of high yielding biofuel feedstocks is the ability to rapidly assess large populations for fermentable sugar yields. Whilst recent advances have outlined methods for the rapid assessment of biomass saccharification efficiency, none take into account the total biomass, or the soluble sugar fraction of the plant. Here we present a holistic high-throughput methodology for assessing sweet Sorghum bicolor feedstocks at 10 days post-anthesis for total fermentable sugar yields including stalk biomass, soluble sugar concentrations, and cell wall saccharification efficiency.

Results

A mathematical method for assessing whole S. bicolor stalks using the fourth internode from the base of the plant proved to be an effective high-throughput strategy for assessing stalk biomass, soluble sugar concentrations, and cell wall composition and allowed calculation of total stalk fermentable sugars. A high-throughput method for measuring soluble sucrose, glucose, and fructose using partial least squares (PLS) modelling of juice Fourier transform infrared (FTIR) spectra was developed. The PLS prediction was shown to be highly accurate with each sugar attaining a coefficient of determination (R 2 ) of 0.99 with a root mean squared error of prediction (RMSEP) of 11.93, 5.52, and 3.23 mM for sucrose, glucose, and fructose, respectively, which constitutes an error of <4% in each case. The sugar PLS model correlated well with gas chromatography–mass spectrometry (GC-MS) and brix measures. Similarly, a high-throughput method for predicting enzymatic cell wall digestibility using PLS modelling of FTIR spectra obtained from S. bicolor bagasse was developed. The PLS prediction was shown to be accurate with an R 2 of 0.94 and RMSEP of 0.64 μg.mgDW-1.h-1.

Conclusions

This methodology has been demonstrated as an efficient and effective way to screen large biofuel feedstock populations for biomass, soluble sugar concentrations, and cell wall digestibility simultaneously allowing a total fermentable yield calculation. It unifies and simplifies previous screening methodologies to produce a holistic assessment of biofuel feedstock potential.
  相似文献   
4.
Drosophila melanogaster thioredoxin reductase-1 (DmTrxR-1) is a key flavoenzyme in dipteran insects, where it substitutes for glutathione reductase. DmTrxR-1 belongs to the family of dimeric, high Mr thioredoxin reductases, which catalyze reduction of thioredoxin by NADPH. Thioredoxin reductase has an N-terminal redox-active disulfide (Cys57-Cys62) adjacent to the flavin and a redox-active C-terminal cysteine pair (Cys489'-Cys490' in the other subunit) that transfer electrons from Cys57-Cys62 to the substrate thioredoxin. Cys489'-Cys490' functions similarly to Cys495-Sec496 (Sec = selenocysteine) and Cys535-XXXX-Cys540 in human and parasite Plasmodium falciparum enzymes, but a catalytic redox center formed by adjacent Cys residues, as observed in DmTrxR-1, is unprecedented. Our data show, for the first time in a high Mr TrxR, that DmTrxR-1 oscillates between the 2-electron reduced state, EH2, and the 4-electron state, EH4, in catalysis, after the initial priming reduction of the oxidized enzyme (Eox) to EH2. The reductive half-reaction consumes 2 eq of NADPH in two observable steps to produce EH4. The first equivalent yields a FADH--NADP+ charge-transfer complex that reduces the adjacent disulfide to form a thiolate-flavin charge-transfer complex. EH4 reacts with thioredoxin rapidly to produce EH2. In contrast, Eox formation is slow and incomplete; thus, EH2 of wild-type cannot reduce thioredoxin at catalytically competent rates. Mutants lacking the C-terminal redox center, C489S, C490S, and C489S/C490S, are incapable of reducing thioredoxin and can only be reduced to EH2 forms. Additional data suggest that Cys57 attacks Cys490' in the interchange reaction between the N-terminal dithiol and the C-terminal disulfide.  相似文献   
5.
DNA in viruses and cells exists in highly condensed, tightly packaged states. We have undertaken an in vitro study of the kinetics of DNA condensation by the trivalent cation hexaammine cobalt (III) with the aim of formulating a quantitative, mechanistic model of the condensation process. Experimental approaches included total intensity and dynamic light scattering, electron microscopy, and differential sedimentation. We determined the average degree of condensation, the distribution of condensate sizes, and the fraction of uncondensed DNA as a function of reaction time for a range of [DNA] and [Co(NH(3))(3+)(6)]. We find the following: (1) DNA condensation occurs only above a critical [Co(NH(3))(3+)(6)] for a given DNA and salt concentration. At the onset of condensation, [Co(NH(3))(3+)(6)]/[DNA-phosphate] is close to the average value of 0.54, which reflects the 89-90% charge neutralization criterion for condensation. (2) The equilibrium weight average hydrodynamic radius of the condensates first decreases, then increases with increasing [Co(NH(3))(3+)(6)] as they undergo a transition from intramolecular (monomolecular) to intermolecular (multimolecular) condensation. However, is insensitive to [DNA]. (3) The uncondensed DNA fraction decays approximately exponentially with time. The equilibrium uncondensed DNA fraction and relaxation time decrease with increasing [Co(NH(3))(3+)(6)] but are insensitive to [DNA]. (4) The condensation rate in its early stages is insensitive to [DNA] but proportional to [Co(NH(3))(3+)(6)](xs) = [Co(NH(3))(3+)(6)] - [Co(NH(3))(3+)(6)](crit). (5) Data for low [DNA] and low [Co(NH(3))(3+)(6)] at early stages of condensation are most reliable for kinetic modeling since under these conditions there is minimal clumping and network formation among separate condensates. A mechanism with initial monomolecular nucleation and subsequent bimolecular association and unimolecular dissociation steps with rate constants that depend on the number of DNA molecules in the condensate, accounts reasonably well for these observations.  相似文献   
6.
Arscott LD  Veine DM  Williams CH 《Biochemistry》2000,39(16):4711-4721
Glutathione reductase catalyzes the reduction of glutathione disulfide by NADPH. The FAD of the reductase is reduced by NADPH, and reducing equivalents are passed to a redox-active disulfide to complete the first half-reaction. The nascent dithiol of two-electron reduced enzyme (EH(2)) interchanges with glutathione disulfide forming two molecules of glutathione in the second half-reaction. It has long been assumed that a mixed disulfide (MDS) between one of the nascent thiols and glutathione is an intermediate in this reaction. In addition to the nascent dithiol composed of Cys(45) and Cys(50), the enzyme contains an acid catalyst, His(456), having a pK(a) of 9.2 that protonates the first glutathione (residue numbers refer to the yeast enzyme sequence). Reduction of yeast glutathione reductase by glutathione and reoxidation of EH(2) by glutathione disulfide indicate that the mixed disulfide accumulates, in particular, at low pH. The reaction of glutathione disulfide with EH(2) is stoichiometric in the absence of an excess of glutathione. The equilibrium position among E(ox), MDS, and EH(2) is determined by the glutathione concentration and is not markedly influenced by pH between 6.2 and 8.5. The mixed disulfide is the principal product in the reaction of glutathione with oxidized enzyme (E(ox)) at pH 6. 2. Its spectrum can be distinguished from that of EH(2) by a slightly lower thiolate (Cys(50))-FAD charge-transfer absorbance at 540 nm. The high GSH/GSSG ratio in the cytoplasm dictates that the mixed disulfide will be the major enzyme species.  相似文献   
7.

Background

Although simulation studies show that combining multiple breeds in one reference population increases accuracy of genomic prediction, this is not always confirmed in empirical studies. This discrepancy might be due to the assumptions on quantitative trait loci (QTL) properties applied in simulation studies, including number of QTL, spectrum of QTL allele frequencies across breeds, and distribution of allele substitution effects. We investigated the effects of QTL properties and of including a random across- and within-breed animal effect in a genomic best linear unbiased prediction (GBLUP) model on accuracy of multi-breed genomic prediction using genotypes of Holstein-Friesian and Jersey cows.

Methods

Genotypes of three classes of variants obtained from whole-genome sequence data, with moderately low, very low or extremely low average minor allele frequencies (MAF), were imputed in 3000 Holstein-Friesian and 3000 Jersey cows that had real high-density genotypes. Phenotypes of traits controlled by QTL with different properties were simulated by sampling 100 or 1000 QTL from one class of variants and their allele substitution effects either randomly from a gamma distribution, or computed such that each QTL explained the same variance, i.e. rare alleles had a large effect. Genomic breeding values for 1000 selection candidates per breed were estimated using GBLUP modelsincluding a random across- and a within-breed animal effect.

Results

For all three classes of QTL allele frequency spectra, accuracies of genomic prediction were not affected by the addition of 2000 individuals of the other breed to a reference population of the same breed as the selection candidates. Accuracies of both single- and multi-breed genomic prediction decreased as MAF of QTL decreased, especially when rare alleles had a large effect. Accuracies of genomic prediction were similar for the models with and without a random within-breed animal effect, probably because of insufficient power to separate across- and within-breed animal effects.

Conclusions

Accuracy of both single- and multi-breed genomic prediction depends on the properties of the QTL that underlie the trait. As QTL MAF decreased, accuracy decreased, especially when rare alleles had a large effect. This demonstrates that QTL properties are key parameters that determine the accuracy of genomic prediction.

Electronic supplementary material

The online version of this article (doi:10.1186/s12711-015-0124-6) contains supplementary material, which is available to authorized users.  相似文献   
8.
9.
The objective of this simulation study was to compare the effect of the number of QTL and distribution of QTL variance on the accuracy of breeding values estimated with genomewide markers (MEBV). Three distinct methods were used to calculate MEBV: a Bayesian Method (BM), Least Angle Regression (LARS) and Partial Least Square Regression (PLSR). The accuracy of MEBV calculated with BM and LARS decreased when the number of simulated QTL increased. The accuracy decreased more when QTL had different variance values than when all QTL had an equal variance. The accuracy of MEBV calculated with PLSR was affected neither by the number of QTL nor by the distribution of QTL variance. Additional simulations and analyses showed that these conclusions were not affected by the number of individuals in the training population, by the number of markers and by the heritability of the trait. Results of this study show that the effect of the number of QTL and distribution of QTL variance on the accuracy of MEBV depends on the method that is used to calculate MEBV.  相似文献   
10.
Thioredoxin reductase (TrxR) catalyzes the reduction of thioredoxin (Trx) by NADPH. Like other members of the pyridine nucleotide-disulfide oxidoreductase enzyme family, the enzyme from Drosophila melanogaster is a homodimer, and each catalytically active unit consists of three redox centers: FAD and an N-terminal Cys-57/Cys-62 redox-active disulfide from one monomer and a Cys-489'/Cys-490' C-terminal redox-active disulfide from the second monomer. Because dipteran insects such as D. melanogaster lack glutathione reductase, thioredoxin reductase (DmTrxR) is particularly important; in addition to its normal functions, it also reduces GSSG for antioxidant protection. DmTrxR, used as a model for the enzyme from the malaria vector, Anopheles gambiae, has been shown to cycle in catalysis between the two-electron and four-electron reduced states, EH2 and EH4 [Bauer, H. et al. (2003) J. Biol. Chem. 278, 33020-33028]. His-464' acts as an acid-base catalyst of the dithiol-disulfide interchange reactions required in catalysis. The H464'Q enzyme has only 2% of the wild-type activity, emphasizing the importance of this residue. The pH dependence of Vmax for wild-type DmTrxR has pKa values of 6.4 and 9.3 on the DmTrxR-DmTrx-2 complex, whereas H464'Q DmTrxR only has an observable pKa at 6.4, indicating that the pKa at pH 9.3 is contributed mainly by His-464'. The pKa at pH 6.4 has been assigned to Cys-57 and Cys-490'; the thiolate on Cys-490' is the nucleophile in the reduction of Trx. In contrast to wild-type DmTrxR, H464'Q DmTrxR does not stabilize a thiolate-FAD charge-transfer complex in the presence of excess NADPH. The rates of steps in both the reductive and the oxidative half-reactions are markedly diminished in H464'Q DmTrxR as compared to those of wild-type enzyme, indicating that His-464' is involved in both half-reactions.  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