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   36篇
  免费   4篇
  2017年   1篇
  2016年   1篇
  2015年   1篇
  2013年   2篇
  2012年   3篇
  2011年   3篇
  2010年   1篇
  2009年   5篇
  2008年   5篇
  2007年   3篇
  2006年   5篇
  2005年   1篇
  1999年   2篇
  1998年   1篇
  1984年   2篇
  1983年   1篇
  1980年   1篇
  1977年   2篇
排序方式: 共有40条查询结果,搜索用时 15 毫秒
11.
M Merkx  M W Pinkse  B A Averill 《Biochemistry》1999,38(31):9914-9925
The pH dependence of the catalytic parameters k(cat) and K(M) has been determined for the Fe(III)Fe(II)- and Fe(III)Zn(II)-forms of bovine spleen purple acid phosphatase (BSPAP). The parameter k(cat) was found to be maximal at pH 6.3, and a pK(a) of 5.4-5.5 was obtained for the acidic limb of the k(cat) vs pH profile. Two different EPR spectra were detected for the phosphate complex of the mixed-valent diiron enzyme; their relative amounts depended on the pH, with an apparent pK(a) of 6. The EPR spectra of Fe(III)Fe(II)-BSPAP.PO(4) and Fe(III)Zn(II)-BSPAP.PO(4) at pH 5.0 are similar to those previously reported for Fe(III)Fe(II)-Uf.PO(4) and Fe(III)Zn(II)-Uf.PO(4) complexes at pH 5.0. At higher pH, a new Fe(III)Fe(II)-BSPAP.PO(4) species is formed, with apparent g-values of 1.94, 1.71, and 1.50. The EPR spectrum of Fe(III)Zn(II)-BSPAP does not show significant changes upon addition of phosphate up to 30 mM at pH 6.5, suggesting that phosphate binds only to the spectroscopically silent Zn(II). To determine whether the phosphate complexes were good structural models for the enzyme substrate complexes, these complexes were studied using rapid-freeze EPR and stopped-flow optical spectroscopy. The stopped-flow studies showed the absence of burst kinetics at pH 7.0, which indicates that substrate hydrolysis is rate limiting, rather than phosphate release. The EPR spectrum of Fe(III)Fe(II)-BSPAP.p-NPP is similar, but not identical, to that of the corresponding phosphate complex, both at pH 5 and pH 6.5. We propose that both phosphate and p-NPP bridge the two metal ions at low pH. At higher pH where the enzyme is optimally active, we propose that hydroxide competes with phosphate and p-NPP for coordination to Fe(III) and that both phosphate and p-NPP coordinate only to the divalent metal ion.  相似文献   
12.
M W Pinkse  M Merkx  B A Averill 《Biochemistry》1999,38(31):9926-9936
Purple acid phosphatases (PAPs) employ a dinuclear Fe(3+)Fe(2+) or Fe(3+)Zn(2+) center to catalyze the hydrolysis of phosphate monoesters. The interaction of fluoride with bovine spleen purple acid phosphatase (BSPAP) has been studied using a combination of steady-state kinetics and spectroscopic methods. For FeZn-BSPAP, the nature of the inhibition changes from noncompetitive at pH 6.5 (K(i(comp)) approximately K(i(uncomp)) approximately 2 mM) to uncompetitive at pH 5.0 (K(i(uncomp)) = 0.2 mM). The inhibition constant for AlZn-BSPAP at pH 5.0 (K(i) = 3 microM) is approximately 50-70-fold lower than that observed for both FeZn-BSAP and GaZn-BSPAP, suggesting that fluoride binds to the trivalent metal. Fluoride binding to the enzyme-substrate complex was found to be remarkably slow; hence, the kinetics of fluoride binding were studied in some detail for FeZn-, AlZn-, and FeFe-BSPAP at pH 5.0 and for FeZn-BSPAP at pH 6.5. Since the enzyme kinetics studies indicated the formation of a ternary enzyme-substrate-fluoride complex, the binding of fluoride to FeZn-BSPAP was studied using optical and EPR spectroscopies, both in the presence and absence of phosphate. The characteristic optical and EPR spectra of FeZn-BSPAP. F and FeZn-BSPAP.PO(4).F are similar at pH 5.0 and pH 6.5, indicating the formation of similar fluoride complexes at both pHs. A structural model for the ternary enzyme-(substrate/phosphate)-fluoride complexes is proposed that can explain the results from both the spectroscopic and the enzyme kinetics experiments. In this model, fluoride binds to the trivalent metal replacing the water/hydroxide ligand that is essential for the hydrolysis reaction to take place, while phosphate or the phosphate ester coordinates to the divalent metal ion.  相似文献   
13.
A combination of techniques to separate and quantify the native proteins associated with a particular transition metal ion from a cellular system has been developed. The procedure involves four steps: (1) labeling of the target proteins with a suitable short-lived radioisotope (suitable isotopes are 64Cu, 67Cu, 187W, 99Mo, 69Zn, 56Mn, 65Ni); (2) separation of intact soluble holoproteins using native isoelectric focusing combined with blue native polyacrylamide gel electrophoresis into native–native 2D gel electrophoresis; (3) spot visualization and quantification using autoradiography; and (4) protein identification with tandem mass spectrometry. The method was applied to the identification of copper proteins from a soluble protein extract of wild-type Escherichia coli K12 using the radioisotope 64Cu. The E. coli protein CueO, which has previously been only identified as a multicopper oxidase following homologous overexpression, was now directly detected as a copper protein against a wild-type background at an expression level of 0.007% of total soluble protein. The retention of the radioisotope by the copper proteins throughout the separation process corroborates the method to be genuinely native. The procedure developed here can be applied to cells of any origin, and to any metal having suitable radioisotopes. The finding that the periplasmic protein CueO is the only major form of soluble protein bound copper in E. coli strengthens the view that the bacterial periplasm contains only a few periplasmic copper proteins, and that the cytosol is devoid of copper proteins. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   
14.
The tungsten metallome of the hyperthermophilic archaeon Pyrococcus furiosus has been investigated using electroanalytical metal analysis and native-native 2D-PAGE with the radioactive tungsten isotope (187)W (t(1/2) = 23.9 h). P. furiosus cells have an intracellular tungsten concentration of 29 μM, of which ca. 30% appears to be free tungsten, probably in the form of tungstate or polytungstates. The remaining 70% is bound by five different tungsten enzymes: formaldehyde ferredoxin oxidoreductase, aldehyde ferredoxin oxidoreductase, glyceraldehyde-3-phosphate ferredoxin oxidoreductase and the tungsten-containing oxidoreductases WOR4 and WOR5. The membrane proteome of P. furiosus is devoid of tungsten. The differential expression, as measured by the tungsten level, of the five soluble tungsten enzymes when the cells are subjected to a cold-shock shows a strong correlation with previously published DNA microarray analyses.  相似文献   
15.
Considerable interest has been focused on the role of myosin light chain LC(2) in the contraction of vertebrate striated muscle. A study was undertaken to further our investigations (Moss, R.L., G.G. Giulian, and M.L. Greaser, 1981, J. Biol. Chem., 257:8588-8591) of the effects of LC(2) removal upon contraction in skinned fibers from rabbit psoas muscles. Isometric tension and maximum velocity of shortening, V(max), were measured in fiber segments prior to LC(2) removal. The segments were then bathed at 30 degrees C for up to 240 min in a buffer solution containing 20 mM EDTA in order to extract up to 60 percent of the LC(2). Troponin C (TnC) was also partially removed by this procedure. Mechanical measurements were done following the EDTA extraction and the readditions of first TnC and then LC(2) to the segments. The protein subunit compositions of the same fiber segments were determined following each of these procedures by SDS PAGE of small pieces of the fiber. V(max) was found to decrease as the LC(2) content of the fiber segments was reduced by increasing the duration of extraction. EDTA treatment also resulted in substantial reductions in tension due mainly to the loss of TnC, though smaller reductions due to the extraction of LC(2) were also observed. Reversal of the order of recombination of LC(2) and TnC indicated that the reduction in V(max) following EDTA treatment was a specific effect of LC(2) removal. These results strongly suggest that LC(2) may have roles in determining the kinetics and extent of interaction between myosin and actin.  相似文献   
16.
The inhibitor peptide DT-2 (YGRKKRRQRRRPPLRKKKKKH) is the most potent and selective inhibitor of the cGMP-dependent protein kinase (PKG) known today. DT-2 is a construct of a PKG tight binding sequence (W45, LRKKKKKH, KI = 0.8 μm) and a membrane translocating sequence (DT-6, YGRKKRRQRRRPP, KI = 1.1 μm), that combined strongly inhibits PKG catalyzed phosphorylation (KI = 12.5 nm) with ∼1000-fold selectivity toward PKG over protein kinase A, the closest relative of PKG. However, the molecular mechanism behind this inhibition is not entirely understood. Using a combination of photoaffinity labeling, stable isotope labeling, and mass spectrometry, we have located the binding sites of PKG-specific substrate and inhibitor peptides. Covalent linkage of a PKG-specific substrate analogue was localized in the catalytic core on residues 356–372, also known as the glycine-rich loop, essential for ATP binding. By analogy, the individual inhibitor peptides W45 and DT-6 were also found to cross-link near the glycine-rich loop, suggesting these are both substrate competitive inhibitors. A bifunctional photoreactive analogue of DT-2 was found to generate dimers of PKG. This cross-linking induced covalent PKG dimerization was not observed for an N-terminal deletion mutant of PKG, which lacks the dimerization domain. In addition, non-covalent mass spectrometry was used to determine binding stoichiometry and binding order of the inhibitor peptides. Dimeric PKG binds two W45 and DT-6 peptides, whereas only one DT-2 molecule was observed to bind to the dimeric PKG. Taken together, these findings imply that (i) the two individual components making up DT-2 are both targeted against the substrate-binding site and (ii) binding of a single DT-2 molecule inactivates both PKG monomers simultaneously, which is an indication that (iii) in cGMP-activated PKG the catalytic centers of both subunits may be in each other''s proximity.Among the superfamily of protein kinases the two cyclic nucleotide-regulated protein kinases, cAMP-dependent protein kinase and cGMP-dependent protein kinase, form a closely related subfamily of serine/threonine protein kinases (14). Both proteins share several structural elements, such as the N-terminal dimerization domain, an autoinhibition site, two in-tandem cyclic nucleotide-binding sites, and a highly conserved catalytic core (Fig. 1, A and B). Despite these similarities, these two enzymes display differences, which account for their unique properties. Whereas PKA2 is nearly ubiquitous, PKG is primarily found in the lung, cerebellum, and smooth muscles (5, 6). From a structural point of view these cyclic nucleotide-dependent protein kinases differ as well. The holoenzyme of PKA is a tetramer composed of two regulatory and two catalytic subunits. The catalytic subunits are non-covalently attached to the regulatory subunit dimer. Upon interaction with cAMP, the catalytic subunits dissociate from the holoenzyme and are free to catalyze heterophosphorylation (Fig. 1C). The mammalian type I PKGs are homodimeric cytosolic proteins containing two identical polypeptides of ∼76 kDa. Alternative mRNA splicing produces type Iα and type Iβ PKG, which are identical proteins apart from their first ∼100 N-terminal residues (7). Each PKG subunit is composed of a regulatory and a catalytic domain on a single polypeptide chain. Consequently, when cGMP activates PKG, the catalytic and regulatory components remain physically attached (Fig. 1D). Within the catalytic domain PKA and PKG share a strong primary sequence homology (8). Not surprisingly, these enzymes also exhibit overlapping substrate specificities, a feature that often interferes with efforts to elucidate their distinct biological pathways. Peptide substrates with a primary amino acid sequence motif RRX(S/T)X are in general recognized by both PKA and PKG (9). Besides this strong overlapping substrate specificity, several studies report on subtle differences in determinants that discriminate for PKA and PKG substrate specificity (1016). To specifically discriminate between PKG and PKA activity in biological assays a highly specific PKG peptide inhibitor was developed (17). This peptide, YGRKKRRQRRRPPLRKKKKKH (DT-2), is the most potent and selective PKG inhibitor known today. Recently, the validity of DT-2 as a superior inhibitor of PKG in terms of potency, selectivity, and membrane permeability has been demonstrated (1824). The inhibitor is a construct of a substrate competitive sequence, LRKKKKKH (W45), derived from a library screen that selected for tight PKG binding sequences, with a significant specificity toward PKG over PKA, and a membrane translocating signal peptide, YGRKKRRQRRRPP (DT-6). DT-2 strongly inhibits PKG-catalyzed phosphorylation (Ki = 12.5 nm), however, the molecular nature of DT-2 inhibition is not entirely understood (25). Because high resolution structural data are not available for PKG, one of our goals is to elucidate binding sites for PKG-specific substrates and inhibitors in more detail using a combination of mass spectrometric techniques and photoaffinity labeling. To further delineate the nature of inhibition we have developed photoaffinity analogues of DT-2 and related inhibitory peptides, as well as a high affinity peptide substrate. The method of photoaffinity labeling enables the direct probing of target proteins through a covalent bond, which is photochemically introduced between a ligand and its specific receptor (26). In combination with modern mass spectrometric techniques this is a powerful approach for the characterization of peptide-protein interactions (27). Substrate and inhibitor peptides containing photoactivatable analogues of phenylalanine, 4-benzoyl-l-phenylalanine (Phe(Bz)) or 4′-(3-(trifluoromethyl)-3H-diazirin-3-yl)-l-phenylalanine (Phe(Tmd)) were synthesized and used to locate their substrate/inhibitor-binding sites on PKG. These measurements indicate that the substrate peptide resides near the glycine-rich loop within the catalytic domain and that the inhibitor peptides are directed similarly toward this substrate-binding site, thereby acting as competitive inhibitors. In addition, nanoflow electrospray ionization time of flight mass spectrometry (ESI-TOF-MS) was performed to study the interaction between DT-2 and PKG in more detail. ESI-MS has proven to be a useful tool to analyze the non-covalent interaction of proteins with ligands, oligonucleotides, peptides, or other proteins (2831). Using this technique, important information on conformational changes (3235), measurement of relative dissociation constants (36, 37), and sequential binding order and cooperativity (38, 39) can be obtained. ESI-MS confirms that PKG is primarily a homodimer and is able to bind four cGMP molecules. Binding of DT-2 was strongly enhanced in the presence of cGMP. Surprising is the observation that only one DT-2 molecule binds to dimeric PKG. The information derived from these measurements allows for molecular modeling and structural refinements of the next generation of PKG-selective inhibitors.Open in a separate windowFIGURE 1.Linear arrangement of the functional domains of the regulatory and catalytic subunit of PKA (A) and PKG (B) type I and schematic representation of the current working models of the activation process of PKA (C) and PKG (D) type 1. Binding of cAMP to the PKA induces a conformational change that results in the dissociation of the catalytic subunits. Binding of cGMP to PKG also induces a conformational change, which exposes the catalytic domains, but both catalytic domains remain near each other via the N-terminal dimerization domain. (Images adapted from Scholten et al. (4).)

TABLE 1

Inhibition contants (KI) of PKA- or PKG-specific peptide inhibitors and the PKA/PKG specificity index
PeptideSequencePKGKiPKAKiSpecificity index (PKA/PKG)Ref.
μmμm
PKI(5–24)TTYDFIASGRTGRRNAIHD-NH21500.0030.0002(11)
WW21TQAKRKKALAMA-NH27.5750100(11)
W45LRKKKKKH0.82 ± 0.33559680(17)
DT-6YGRGGRRQRRRPP1.1 ± 0.2226 ± 423.6(17)
DT-2YGRKKRRQRRRPPLRKKKKKH0.0125 ± 0.00316.5 ± 3.81320(17)
Open in a separate window  相似文献   
17.
18.
The number of pore complexes per nucleus was determined for a wide variety of cultured cells selected for their variable DNA content over a range of 1-5,6000. The pore number was compared to DNA content, nuclear surface area, and nuclear volume. Values for pore frequency (pores/square micrometer) were relatively constant in the species studied. When the pore to DNA ratio was plotted against the DNA content, there was a remarkable correlation which decreased exponentially for the cells of vertebrae origin. Exceptions were the heteroploid mammalian cells which had the same ratio as the diploid mammalian cells despite higher DNA content. The results are interpreted to mean that neither the nuclear surface, the nuclear volume, nor the DNA content alone determines the pore number of the nucleus, but rather an as yet undetermined combination of different factors. The surface and volume of vertebrate nuclei do not decrease with decreasing DNA content below a given value. The following speculation is suggested to account for the anomalous size changes of the nucleus relative to DNA content in vertebrates. Species with small DNA complements have a relatively large proportion of active chromatin which determines the limits of the physical parameters of the nucleus. The amount of active chromatin maybe the same for at least the vertebrates with low DNA content, At high DNA content, the nuclear parameters may be determined by the relatively high proportion of inactive condensed chromatin which increases the nuclear surface and volume.  相似文献   
19.
A high‐throughput sample preparation protocol based on the use of 96‐well molecular weight cutoff (MWCO) filter plates was developed for shotgun proteomics of cell lysates. All sample preparation steps, including cell lysis, buffer exchange, protein denaturation, reduction, alkylation and proteolytic digestion are performed in a 96‐well plate format, making the platform extremely well suited for processing large numbers of samples and directly compatible with functional assays for cellular proteomics. In addition, the usage of a single plate for all sample preparation steps following cell lysis reduces potential samples losses and allows for automation. The MWCO filter also enables sample concentration, thereby increasing the overall sensitivity, and implementation of washing steps involving organic solvents, for example, to remove cell membranes constituents. The optimized protocol allowed for higher throughput with improved sensitivity in terms of the number of identified cellular proteins when compared to an established protocol employing gel‐filtration columns.  相似文献   
20.
In the developing embryo, as in many other biological processes, complex signaling pathways are under tight control of reversible phosphorylation, guiding cell proliferation, differentiation, and growth. Therefore the large-scale identification of signaling proteins and their post-translational modifications is crucial to understand the proteome biology of the developing zebrafish embryo. Here, we used an automated, robust, and sensitive online TiO 2-based LC-MS/MS setup to enrich for phosphorylated peptides from 1 day old zebrafish embryos. We identified, with high confidence, 1067 endogenous phosphorylation sites in a sample taken from 60 embryos (approximately 180 microg), 321 from 10 embryos, and 47 phosphorylation sites from a single embryo, illustrating the sensitivity of the method. This data set, representing by far the largest for zebrafish, was further exploited by searching for serine/threonine or tyrosine kinase motifs using Scansite. For one-third of the identified phosphopeptides a potential kinase motif could be predicted, where it appeared that Cdk5 kinase, p38MAPK, PKA, and Casein Kinase 2 substrates were the most predominant motifs present, underpinning the importance of these kinases in signaling pathways in embryonic development. The phosphopeptide data set was further interrogated using alignments with phosphopeptides identified in recent large-scale phosphoproteomics screens in human and mouse samples. These alignments revealed conservation of phosphorylation sites in several proteins suggesting preserved function in embryonic development.  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