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   96篇
  免费   8篇
  2020年   2篇
  2016年   1篇
  2015年   4篇
  2014年   2篇
  2013年   5篇
  2012年   8篇
  2011年   5篇
  2010年   7篇
  2009年   3篇
  2008年   5篇
  2007年   6篇
  2006年   4篇
  2005年   6篇
  2004年   6篇
  2003年   5篇
  2002年   3篇
  2001年   1篇
  2000年   4篇
  1999年   2篇
  1998年   1篇
  1997年   2篇
  1996年   1篇
  1995年   2篇
  1994年   1篇
  1992年   1篇
  1991年   2篇
  1990年   1篇
  1989年   1篇
  1987年   1篇
  1985年   1篇
  1984年   2篇
  1980年   2篇
  1978年   2篇
  1977年   1篇
  1974年   1篇
  1973年   1篇
  1972年   1篇
  1966年   1篇
排序方式: 共有104条查询结果,搜索用时 31 毫秒
1.
H Deng  J Zheng  D Sloan  J Burgner  R Callender 《Biochemistry》1989,28(4):1525-1533
The binding of the coenzymes NAD+ and NADH to lactate dehydrogenase causes significant changes in the Raman spectra of both of these molecules relative to spectra obtained in the absence of enzyme. The molecular motions of the bound adenine moiety of both NAD+ and NADH as well as adenine containing analogues of these coenzymes produce Raman bands that are essentially identical, suggesting that the binding of adenine to the enzyme is the same regardless of the nicotinamide head-group nature. We also have observed that the molecular motions of the bound adenine moiety are different from both those obtained when it is in either water, various hydrophobic solvents, or various other solvent compositions. Protonation of the bound adenine ring at the 3-position is offered as a possible explanation. Significant shifts are observed in both the stretching frequency of the carboxamide carbonyl of NAD+ and the rocking motion of the carboxamide NH2 group of NADH. These shifts are probably caused by hydrogen bonding with the enzyme. The interaction energies of these hydrogen-bonding patterns are discussed. The aromatic nature of the nicotinamide moiety of NAD+ appears to be unchanged upon binding. Pronounced changes in the Raman spectrum of the nicotinamide moiety of NADH are observed upon binding; some of these changes are understood and discussed. Finally, these results are compared to analogous results that were recently reported for liver alcohol dehydrogenase [Chen et al. (1987) Biochemistry 26, 4776-4784]. In general, the coenzyme binding properties are found to be quite similar, but not identical, for the two enzymes.  相似文献   
2.
On the origin of the lactate dehydrogenase induced rate effect   总被引:2,自引:0,他引:2  
J W Burgner  W J Ray 《Biochemistry》1984,23(16):3636-3648
To evaluate the ability of lactate dehydrogenase to facilitate the bond making/breaking steps for both the addition of pyruvate enol to NAD (pyruvate adduct reaction) and the normal redox reaction, the ability of the enzyme to facilitate the tautomerization of bound pyruvate is assessed. In addition, the equilibrium constants for the adduct reaction are obtained for both bound and free reactants from the ratio of the rate constants in the forward and reverse reactions (at pH 7). The latter comparison indicates that the enzyme facilitates bond making/breaking in the (forward) pyruvate adduct reaction by a factor of about 10(11) M. Similar comparisons suggest that reactant immobilization accounts for about 1000 M of this 10(11) M rate effect. Since the (pH-independent) rate constant for the ketonization of bound pyruvate enol assisted by the external buffer, imidazolium ion, is 2 X 10(7) M-1 s-1 and the corresponding rate constant for free pyruvate enol, again assisted by imidazolium ion, is 35 M-1 s-1 [Burger, J. W., II, & Ray, W. J., Jr. (1978) Biochemistry 17, 1664], the enzyme facilitates the bond making/breaking steps associated with the conversion of bound HO-C less than to bound O = C less than by a factor of about 10(6)-fold. The product of the above two rate enhancement factors and the rate factor suggested previously for the environmental effect on NAD produced by its binding to lactate dehydrogenase, 100-fold, is 10(11) M, and it accounts for the bond making/breaking effects exerted by the enzyme in the pyruvate adduct reaction. The rate constant for oxidation of ethanol (a model for lactate) by 1-methylnicotinamide (a model for NAD) is about 5 X 10(-12) M-1 s-1 at 25 degrees C in pure ethanol (delta H for this reaction is about 30 kcal/mol). The ratio of the rate constants for E X NAD X Lac----E X NADH X Pyr and the above model reaction is estimated as about 10(14) M in water; i.e., the LDH-induced rate effect is about 10(14) M. The product of the values for the above rate factors for the normal redox reaction is about 10(12) M. Although the value of this product is less certain than that for the adduct reaction, these rate factors do account for much of the LDH-induced rate effect.  相似文献   
3.
J W Burgner  W J Ray 《Biochemistry》1984,23(16):3620-3626
The binary complex of NAD and lactate dehydrogenase reacts reversibly with cyanide to produce a complex (E X NAD-CN) whose noncovalent interactions are similar to those in the E X NADH complex (where E is one-fourth of the tetrameric dehydrogenase). The reaction apparently is a simple bimolecular nucleophilic addition at the 4 position of the bound nicotinamide ring; viz., cyanide does not bind to the enzyme prior to reaction. The value of the dissociation constant for E X NAD-CN is about 1 X 10(-6) M and is independent of pH over the range of 6-8. The equilibrium constant for the reaction of cyanide with E X NAD is about 400-fold larger than that for the nonenzymic process after a statistical correction. This increment in Ke is accounted for by a 220-fold increase in the rate of the forward enzymic reaction (20 M-1 s-1) as compared with an approximately 2-fold decrease for the reverse process (9 X 10(-5) s-1). Thus, the increased value of the rate constant for bond formation in the enzymic reaction is attributed to an equilibrium binding effect that is translated almost entirely into a rate effect on that step (bond formation). Since the nonenzymic reaction is sensitive to solvent composition, this equilibrium binding effect likely is produced by environmental effects at the nicotinamide/dehydronicotinamide part of the coenzyme binding site on the enzyme.  相似文献   
4.
Total Ca content and that fraction of Ca sensitive to removal by the chelator ethylene glycol-bis(β-aminoethyl ether)N,N,N',N'-tetraacetate (EGTA) have been investigated in the mouse 3T3 cell as a function of growth stage, transformation with SV40 virus, and serum levels of the media. Cells were allowed to grow through several doublings in media containing (45)Ca. The cellular content of (45)Ca was used to access total cell Ca. That fraction of (45)Ca removed by EGTA was presumed to represent primarily surface-localized Ca. The data are expressed on a per cell volume basis to compensate for size differences as a function of growth stage and transformation. During exponential growth phase, the 3T3 cell contains 525pmol Ca/μl cell volume. Of this, approx. 457 pmol/μl is not removable by EGTA and, presumably, is cytoplasmically located. This value is in close agreement with previous studies on the HeLa cell (470 pmol Ca/μl cell water after the removal of the surface Ca). The low level of EGTA- removable Ca present in the 3T3 cell during early exponential growth (68 pmol Ca/μl cell volume) increases progressively with increasing cell density, and upon quiescence it is sevenfold greater. In contrast, SV40- transformed 3T3 cells growing exponentially possess total levels of Ca which are approximately two-thirds the levels of the normal 3T3 cell. However, their EGTA-sensitive Ca is not significantly different from that of exponentially growing, normal 3T3 cells. As the transformed cells continue to grow at high density, their total ca and their sensitivity to EGTA do not change, in contrast to the normal 3T3 cell. Thus, an increase in Ca associated with the cell surface appears to be correlated with growth inhibition. This has been investigated further by regulating growth of the normal and transformed cell with alterations in the serum level of the media. In 4 percent calf serum the normal cell is stopped from continued proliferation. Growth stoppage under these conditions is characterized by a nearly fourfold increase in EGTA-removable Ca, similar to the increase observed upon quiescence in depleted 10 percent serum. Similar treatment of the transformed cell does not reduce its growth rate, nor does it significantly alter Ca distribution. However, at 0.5 percent medium serum levels, the SV40 3T3 growth rate is substantially reduced and, under these conditions, EGTA-removable Ca increases twofold.  相似文献   
5.
6.
7.
S Chen  J W Burgner  J M Krahn  J L Smith  H Zalkin 《Biochemistry》1999,38(36):11659-11669
Single tryptophan residues were incorporated into each of three peptide segments that play key roles in the structural transition of ligand-free, inactive glutamine phosphoribosylpyrophosphate (PRPP) amidotransferase to the active enzyme-substrate complex. Intrinsic tryptophan fluorescence and fluorescence quenching were used to monitor changes in a phosphoribosyltransferase (PRTase) "flexible loop", a "glutamine loop", and a C-terminal helix. Steady state fluorescence changes resulting from substrate binding were used to calculate binding constants and to detect the structural rearrangements that coordinate reactions at active sites for glutamine hydrolysis and PRTase catalysis. Pre-steady state kinetics of enzyme.PRPP and enzyme.PRPP.glutamine complex formation were determined from stopped-flow fluorescence measurements. The kinetics of the formation of the enzyme.PRPP complex were consistent with a model with two or more steps in which rapid equilibrium binding of PRPP is followed by a slow enzyme isomerization. This isomerization is ascribed to the closing of the PRTase flexible loop and is likely the rate-limiting step in the reaction of PRPP with NH(3). The pre-steady state kinetics for binding glutamine to the binary enzyme. PRPP complex could also be fit to a model involving rapid equilibrium binding of glutamine followed by an enzyme isomerization step. The changes monitored by fluorescence account for the interconversions between "end state" structures determined previously by X-ray crystallography and define an intermediate enzyme.PRPP conformer.  相似文献   
8.

Background  

Leber's congenital amaurosis (LCA) is a severe form of retinal dystrophy. Mutations in the RPE65 gene, which is abundantly expressed in retinal pigment epithelial (RPE) cells, account for approximately 10–15% of LCA cases. In this study we used the high turnover, and rapid breeding and maturation time of the Rpe65 -/- knockout mice to assess the efficacy of using rAAV-mediated gene therapy to replace the disrupted RPE65 gene. The potential for rAAV-mediated gene treatment of LCA was then analyzed by determining the pattern of RPE65 expression, the physiological and histological effects that it produced, and any improvement in visual function.  相似文献   
9.
To assess the hypothesis that nitric oxide is critical in the pathogenesis of cerebral malaria, we analysed genetic variation in the proximal promoter region of NOS2A, the gene encoding inducible nitric oxide synthase. Sequencing 72 Gambian chromosomes revealed 11 single nucleotide polymorphisms in 2.5 kB (theta=8.6 x 10(-4)). Genotyping 104 nuclear families identified six common haplotypes. A single haplotype, uniquely defined by the NOS2A-1659T allele, was associated with cerebral malaria by a transmission disequilibrium test of 334 affected children and their parents (P=0.02). An independent case-control study of 505 different children from the same population replicated the allelic association with cerebral malaria (odds ratio: 1.31, P=0.04). Taken together these data indicate a weak but significant association of the NOS2A locus with susceptibility to cerebral malaria. Despite high linkage disequilibrium across the region studied, this association would not have been detected without the initial construction of a dense marker set for haplotype tagging.  相似文献   
10.
Biosynthesis of the isoprenoid precursor isopentenyl diphosphate (IPP) proceeds via two distinct pathways. Sequence comparisons and microbiological data suggest that multidrug-resistant strains of gram-positive cocci employ exclusively the mevalonate pathway for IPP biosynthesis. Bacterial mevalonate pathway enzymes therefore offer potential targets for development of active site-directed inhibitors for use as antibiotics. We used the PCR and Enterococcus faecalis genomic DNA to isolate the mvaS gene that encodes 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) synthase, the second enzyme of the mevalonate pathway. mvaS was expressed in Escherichia coli from a pET28 vector with an attached N-terminal histidine tag. The expressed enzyme was purified by affinity chromatography on Ni(2+)-agarose to apparent homogeneity and a specific activity of 10 micromol/min/mg. Analytical ultracentrifugation showed that the enzyme is a dimer (mass, 83.9 kDa; s(20,w), 5.3). Optimal activity occurred in 2.0 mM MgCl(2) at 37(o)C. The DeltaH(a) was 6,000 cal. The pH activity profile, optimum activity at pH 9.8, yielded a pK(a) of 8.8 for a dissociating group, presumably Glu78. The stoichiometry per monomer of acetyl-CoA binding was 1.2 +/- 0.2 and that of covalent acetylation was 0.60 +/- 0.02. The K(m) for the hydrolysis of acetyl-CoA was 10 microM. Coupled conversion of acetyl-CoA to mevalonate was demonstrated by using HMG-CoA synthase and acetoacetyl-CoA thiolase/HMG-CoA reductase from E. faecalis.  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