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   215篇
  免费   51篇
  国内免费   1篇
  2021年   5篇
  2020年   1篇
  2018年   1篇
  2017年   5篇
  2016年   7篇
  2015年   9篇
  2014年   6篇
  2013年   10篇
  2012年   9篇
  2011年   8篇
  2010年   12篇
  2009年   12篇
  2008年   13篇
  2007年   11篇
  2006年   10篇
  2005年   6篇
  2004年   4篇
  2003年   10篇
  2002年   6篇
  2001年   12篇
  2000年   8篇
  1999年   4篇
  1998年   13篇
  1997年   8篇
  1996年   6篇
  1995年   5篇
  1994年   5篇
  1993年   8篇
  1992年   4篇
  1991年   5篇
  1990年   6篇
  1989年   6篇
  1988年   3篇
  1987年   3篇
  1986年   1篇
  1985年   2篇
  1984年   1篇
  1983年   1篇
  1982年   1篇
  1981年   1篇
  1980年   1篇
  1979年   3篇
  1978年   2篇
  1977年   3篇
  1976年   2篇
  1975年   4篇
  1974年   4篇
排序方式: 共有267条查询结果,搜索用时 15 毫秒
1.
Abstract The gene encoding flavodoxin from Desulfovibrio vulgaris Hildenborough (148 amino acid residues), the first flavoprotein for which a three-dimensional structure has been determined, was cloned with the use of two synthetic oligonucleotides, designed to recognize the coding sequence for amino acid residues 11–19 and 98–103, respectively. The two oligonucleotides were used to screen a library of 900 λ-clones of the D. vulgaris chromosome. A single clone, λFL1, reacting with both probes was isolated. The entire structural gene for flavodoxin is contained in the 15 kb insert of λFL1 as found by nucleic acid sequencing. The codon usage in the flavodoxin gene is strongly biased towards G or C in the third codon position. A table in which codon usage information from all genes of D. vulgaris sequenced to date is combined is presented and should facilitate further gene cloning with oligonucleotide probes.  相似文献   
2.
The micellar weights of dihexanoyl- and diheptanoyllecithin in aqueous solutions are calculated from light scattering and ultracentrifugation data. A monomer-micelle assocation model is used and corrections for the thermodynamic nonideality, on the basis of rigid noninteracting particles, are applied. A few experiments on the influence of high NaCI concentrations (up to 3 M) are described. Dihexanoyllecithin forms micelles with micellar weight of 15 000 to 20 000 and with rather narrow weight distributions. Diheptanoyllecithin micelles however, have broad size distributions with micellar weights of 20 000 up to about 100 000 in the concentration range studied. Micelles are assumed to be spherical or to have sphero-cylindrical shapes depending on the molecular weights. Two models are used: (1) a compact structure, where no attention is paid to the hydrocarbon-water contact (2) micelles with as little hydrocarbon-water contact as possible.  相似文献   
3.
The gene of high molecular weight, multiheme cytochrome c (Hmc) from the sulfate-reducing bacterium Desulfovibrio vulgaris Hildenborough has been overexpressed in Desulfovibrio desulfuricans G200. The recombinant protein has been purified. Its molecular weight (65,600), amino acid composition, and NH2-terminal sequence were found to be identical to those of the wild-type protein. The recombinant protein has been spectroscopically characterized (optical spectrum, EPR, circular dichroism) and compared to the wild-type protein. We have found 16 hemes per molecule by iron analysis and the pyridine hemochrome test. Both high- and low-spin features were observed in the EPR spectrum. A detailed spin quantitation analysis indicates 1 or 2 high-spin hemes and 14 or 15 low-spin hemes per molecule. The redox potentials of the hemes determined by voltammetric techniques gave an average of three different values, 0, -100, and -250 mV (versus NHE), for the wild-type and the recombinant cytochrome. The low potential values are similar to the values observed for the bis(histidinyl) coordinated hemes of cytochrome c3. A comparison of the arrangement of heme binding sites and coordinated histidines in the amino acid sequences of cytochrome c3 and Hmc has shown that the latter contains four domains, three of which are complete c3-like domains, while the fourth represents an incomplete c3-like domain which may contain His-Met coordinated hemes. These data are in agreement with the detailed study of the number and types of hemes reported in this paper.  相似文献   
4.
We have broadly defined the DNA regions regulating esterase6 activity in several life stages and tissue types of D. melanogaster using P- element-mediated transformation of constructs that contain the esterase6 coding region and deletions or substitutions in 5' or 3' flanking DNA. Hemolymph is a conserved ancestral site of EST6 activity in Drosophila and the primary sequences regulating its activity lie between -171 and -25 bp relative to the translation initiation site: deletion of these sequences decrease activity approximately 20-fold. Hemolymph activity is also modulated by four other DNA regions, three of which lie 5' and one of which lies 3' of the coding region. Of these, two have positive and two have negative effects, each of approximately twofold. Esterase6 activity is present also in two male reproductive tract tissues; the ejaculatory bulb, which is another ancestral activity site, and the ejaculatory duct, which is a recently acquired site within the melanogaster species subgroup. Activities in these tissues are at least in part independently regulated: activity in the ejaculatory bulb is conferred by sequences between -273 and -172 bp (threefold decrease when deleted), while activity in the ejaculatory duct is conferred by more distal sequences between -844 and -614 bp (fourfold decrease when deleted). The reproductive tract activity is further modulated by two additional DNA regions, one in 5' DNA (-613 to -284 bp; threefold decrease when deleted) and the other in 3' DNA (+1860 to +2731 bp; threefold decrease when deleted) that probably overlaps the adjacent esteraseP gene. Collating these data with previous studies suggests that expression of EST6 in the ancestral sites is mainly regulated by conserved proximal sequences while more variable distal sequences regulate expression in the acquired ejaculatory duct site.   相似文献   
5.
R Fu  J D Wall    G Voordouw 《Journal of bacteriology》1994,176(2):344-350
The amino acid sequence of DcrA from Desulfovibrio vulgaris Hildenborough, a strictly anaerobic, sulfate-reducing bacterium, indicated homology with the methyl-accepting chemotaxis proteins from enteric bacteria (A. Dolla, R. Fu, M. J. Brumlik, and G. Voordouw, J. Bacteriol. 174:1726-1733, 1992). The homology is restricted to the cytoplasmic C-terminal signaling domain. The periplasmic N-terminal sensor domain was found to contain a unique sequence, CHHCH, corresponding to a consensus c-type heme binding site. A pretreated, DcrA-specific polyclonal antiserum, generated against DcrA protein overproduced in Escherichia coli, was used for immunoprecipitation of 35S-labeled DcrA from D. vulgaris and Desulfovibrio desulfuricans G200(pJRFR2), a transconjugant that overexpresses functional DcrA. Labeling of the latter with the heme precursor 5-amino-[4-14C]levulinic acid, followed by immunoprecipitation, sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and fluorography, confirmed the presence of c-type heme, while labeling with L-[methyl-3H]methionine in the absence of protein synthesis confirmed that DcrA is a methyl-accepting protein. The base liability of the incorporated radioactivity indicated methyl ester formation like that occurring in the methyl-accepting chemotaxis proteins of enteric bacteria. L-[methyl-3H]methionine labeling of D. desulfuricans G200(pJRFR2) under different conditions indicated that methyl labeling of DcrA decreased upon addition of oxygen and increased upon subsequent addition of the reducing agent dithionite. These results indicate that DcrA may serve as a sensor of oxygen concentration and/or redox potential.  相似文献   
6.
A library of 879 recombinant lambda phages, constructed for the genome of Desulfovibrio vulgaris Hildenborough, has been ordered by restriction fingerprinting. Restriction endonuclease HinfI digestion patterns were entered into a data base and sorted into 87 overlapping groups (contigs), with 19 clones remaining unattached. Eight of ten cloned genes of D. vulgaris, including dcrA, which encodes a transmembrane methyl-accepting protein, were assigned to contigs. Probing of a filter containing the lambda DNAs of the library with the labeled, conserved 3' end of the dcrA gene indicated hybridization to 54 clones distributed over multiple contigs. The presence of 11 additional dcr genes (dcrB to dcrL) was confirmed by direct cycled dideoxy sequencing of positive lambda clones. Since the ordered library provides only partial coverage of the D. vulgaris Hildenborough genome, we estimate that the dcr gene family has 16 members spread throughout the genome, making it the second largest gene family found in prokaryotes.  相似文献   
7.
The nucleotide sequence of the hmc operon from Desulfovibrio vulgaris subsp. vulgaris Hildenborough indicated the presence of eight open reading frames, encoding proteins Orf1 to Orf6, Rrf1, and Rrf2. Orf1 is the periplasmic, high-molecular-weight cytochrome (Hmc) containing 16 c-type hemes and described before (W. B. R. Pollock, M. Loutfi, M. Bruschi, B. J. Rapp-Giles, J. D. Wall, and G. Voordouw, J. Bacteriol. 173:220-228, 1991). Orf2 is a transmembrane redox protein with four iron-sulfur clusters, as indicated by its similarity to DmsB from Escherichia coli. Orf3, Orf4, and Orf5 are all highly hydrophobic, integral membrane proteins with similarities to subunits of NADH dehydrogenase or cytochrome c reductase. Orf6 is a cytoplasmic redox protein containing two iron-sulfur clusters, as indicated by its similarity to the ferredoxin domain of [Fe] hydrogenase from Desulfovibrio species. Rrf1 belongs to the family of response regulator proteins, while the function of Rrf2 cannot be derived from the gene sequence. The expression of individual genes in E. coli with the T7 system confirmed the open reading frames for Orf2, Orf6, and Rrf1. Deletion of 0.4 kb upstream from orf1 abolished the expression of Hmc in D. desulfuricans G200, indicating this region to contain the hmc operon promoter. The expression of two truncated hmc genes in D. desulfuricans G200 resulted in stable periplasmic c-type cytochromes, confirming the domain structure of Hmc. We propose that Hmc and Orf2 to Orf6 form a transmembrane protein complex that allows electron flow from the periplasmic hydrogenases to the cytoplasmic enzymes that catalyze the reduction of sulfate. The domain structure of Hmc may be required to allow interaction with multiple hydrogenases.  相似文献   
8.
Immunoglobulins are encoded by a large multigene system that undergoes somatic rearrangement and additional genetic change during the development of immunoglobulin-producing cells. Inducible antibody and antibody-like responses are found in all vertebrates. However, immunoglobulin possessing disulfide-bonded heavy and light chains and domain-type organization has been described only in representatives of the jawed vertebrates. High degrees of nucleotide and predicted amino acid sequence identity are evident when the segmental elements that constitute the immunoglobulin gene loci in phylogenetically divergent vertebrates are compared. However, the organization of gene loci and the manner in which the independent elements recombine (and diversify) vary markedly among different taxa. One striking pattern of gene organization is the "cluster type" that appears to be restricted to the chondrichthyes (cartilaginous fishes) and limits segmental rearrangement to closely linked elements. This type of gene organization is associated with both heavy- and light-chain gene loci. In some cases, the clusters are "joined" or "partially joined" in the germ line, in effect predetermining or partially predetermining, respectively, the encoded specificities (the assumption being that these are expressed) of the individual loci. By relating the sequences of transcribed gene products to their respective germ-line genes, it is evident that, in some cases, joined-type genes are expressed. This raises a question about the existence and/or nature of allelic exclusion in these species. The extensive variation in gene organization found throughout the vertebrate species may relate directly to the role of intersegmental (V<==>D<==>J) distances in the commitment of the individual antibody-producing cell to a particular genetic specificity. Thus, the evolution of this locus, perhaps more so than that of others, may reflect the interrelationships between genetic organization and function.   相似文献   
9.
The structural genes for dissimilatory sulfite reductase (desulfoviridin) from Desulfovibrio vulgaris Hilden-borough were cloned as a 7.2-kbp SacII DNA fragment. Nucleotide sequencing indicated the presence of a third gene, encoding a protein of only 78 amino acids, immediately downstream from the genes for the alpha and beta subunits (dsvA and dsvB). We designated this protein DsvD and the gene encoding it the dsvD gene. The alpha- and beta-subunit sequences are highly homologous to those of the dissimilatory sulfite reductase from Archaeoglobus fulgidus, a thermophilic archaeal sulfate reducer, which grows optimally at 83 degrees C. A gene with significant homology to dsvD was also found immediately downstream from the dsrAB genes of A. fulgidus. The remarkable conservation of gene arrangement and sequence across domain (bacterial versus archaeal) and physical (mesophilic versus thermophilic) boundaries indicates an essential role for DsvD in dissimilatory sulfite reduction and allowed the construction of conserved deoxyoligonucleotide primers for detection of the dissimilatory sulfite reductase genes in the environment.  相似文献   
10.
Flavodoxin from Desulfovibrio vulgaris (Hildenborough) has been expressed at a high level (3-4% soluble protein) in Escherichia coli by subcloning a minimal insert carrying the gene behind the tac promoter of plasmid pDK6. The recombinant protein was readily isolated and its properties were shown to be identical to those of the wild-type protein obtained directly from D. vulgaris, with the exception that the recombinant protein lacks the N-terminal methionine residue. Detailed measurements of the redox potentials of this flavodoxin are reported for the first time. The redox potential, E2, for the couple oxidized flavodoxin/flavodoxin semiquinone at pH 7.0 is -143 mV (25 degrees C), while the value for the flavodoxin semiquinone/flavodoxin hydroquinone couple (E1) at the same pH is -440 mV. The effects of pH on the observed potentials were examined; E2 varies linearly with pH (slope = -59 mV), while E1 is independent of pH at high pH values, but below pH 7.5 the potential becomes less negative with decreasing pH, indicating a redox-linked protonation of the flavodoxin hydroquinone. D. vulgaris apoflavodoxin binds FMN very tightly, with a value of 0.24 nM for the dissociation constant (Kd) at pH 7.0 and 25 degrees C, similar to that observed with other flavodoxins. In addition, the apoflavodoxin readily binds riboflavin (Kd = 0.72 microM; 50 mM sodium phosphate, pH 7.0, 5 mM EDTA at 25 degrees C) and the complex is spectroscopically very similar to that formed with FMN. The redox potentials for the riboflavin complex were determined at pH 6.5 (E1 = -262 mV, E2 = -193 mV; 25 degrees C) and are discussed in the light of earlier proposals that charge/charge interactions between different parts of the flavin hydroquinone play a crucial role in determining E1 in flavodoxin.  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