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 328 毫秒
1.
Hearing requires proper function of the auditory hair cell, which is critically dependent upon its actin-based cytoskeletal structure. Currently, ten point mutations in nonmuscle γ-actin have been identified as causing progressive autosomal dominant nonsyndromic hearing loss (DFNA20/26), highlighting these ten residues as functionally important to actin structure and/or regulation. Two of the mutations, K118M and K118N, are located near the putative binding site for the ubiquitously expressed Arp2/3 complex. We therefore hypothesized that these mutations may affect Arp2/3-dependent regulation of the actin cytoskeleton. Using in vitro bulk polymerization assays, we show that the Lys-118 mutations notably reduce actin + Arp2/3 polymerization rates compared with WT. Further in vitro analysis of the K118M mutant using TIRF microscopy indicates the actual number of branches formed per filament is reduced compared with WT and, surprisingly, branch location is altered such that the majority of K118M branches form near the pointed end of the filament. These results highlight a previously unknown role for the Lys-118 residue in the actin-Arp2/3 interaction and also further suggest that Lys-118 may play a more significant role in intra- and intermonomer interactions than was initially hypothesized.  相似文献   

2.
J A Barden 《Biochemistry》1987,26(19):6023-6030
The actin peptide 106-124 exists in a completely conserved region of the sequence and binds strongly to both ATP and tripolyphosphate. Binding particularly affects residues 116 and 118 and generally affects the two segments 115-118 and 121-124 [Barden, J. A., & Kemp, B. E. (1987) Biochemistry 26, 1471-1478]. One-dimensional nuclear Overhauser enhancement difference spectroscopy was used to detect molecular interactions between both adjacent and nonadjacent residues. The N-terminal segment 106-112 was found to be largely extended. A sharp bend was detected between Pro-112 and Lys-113. The triphosphate moiety binds to the strongly hydrophilic central segment of the peptide. Evidence was obtained for a reverse turn involving residues 121-124. Amide proton temperature coefficients and coupling constants provide evidence for a type I beta-turn. A model of the ATP binding site is proposed together with its relationship to other parts of the actin structure and to the phalloidin binding site.  相似文献   

3.
Activated platelets and phospholipid vesicles promote assembly of the intrinsic factor X (FX) activating complex by presenting high-affinity binding sites for blood coagulation FIXa, FVIIIa, and FX. Previous reports suggest that the second epidermal growth factor (EGF)-like domain of FIXa mediates assembly of the FX activating complex (Ahmad, S. S., Rawala, R., Cheung, W. F., Stafford, D. W., and Walsh, P. N. (1995) Biochem. J. 310, 427-431; Wong, M. Y., Gurr, J. A., and Walsh, P. N. (1999) Biochemistry 38, 8948-8960). To identify important residues, we prepared several chimeric FIXa proteins using homologous sequences from FVII: FIXa(FVIIEGF2) (FIX Delta 88-124,inverted Delta FVII91-127), FIXa(loop1) (FIX Delta 88-99,inverted Delta FVII91-102), FIXa(loop2) (FIX Delta 95-109,inverted Delta FVII98-112), FIXa(loop3) (FIX Delta 111-124,inverted Delta FVII114-127), and point mutants (FIXaR94D and FIXa(loop1)G94R). In the presence and absence of FVIIIa, a 2- to 10-fold reduced V(max) of FX activation (nm FXa min(-1)) was observed for FIXa(FVIIEGF2), FIXa(loop1), FIXa(loop2), and FIXa(loop1)G94R, whereas FIXa(loop3) and FIXaR94D were normal. For all of the FIXa proteins, K(m)((app)) values were normal as were EC(50) values for interactions with FVIIIa. However, K(d)((app)) (in nm) for the FX activating complex assembled on phospholipid vesicles was increased for FIXa(FVIIEGF2) (43.3 +/- 2.70), FIXa(loop1)(10.9 +/- 2.8), FIXa(loop2) (70.5 +/- 1.60), and FIXa(loop1)G94R (17.1 +/- 2.90) relative to FIXa(N) (3.9 +/- 0.11), FIXa(WT) (4.6 +/- 0.17), FIXa(loop3) (4.5 +/- 0.20), and FIXaR94D (2.2 +/- 0.09) suggesting that reduced V(max) is a result of impaired complex assembly. These data indicate that residues 88-109 (but not Arg(94)) are important for normal assembly of the FX activating complex on phospholipid vesicles.  相似文献   

4.
The N-terminus of all actins so far studied is acetylated. Although the pathways of acetylation have been well studied, its functional importance has been unclear. A negative charge cluster in the actin N-terminal region is shown to be important for the function of actomyosin. Acetylation at the N-terminus removes a positive charge and increases the amount of net negative charges in the N-terminal region. This may augment the role of the negative charge cluster. To examine this possibility, actin with a nonacetylated N-terminus (nonacetylated actin) was produced. The nonacetylated actin polymerized and depolymerized normally. In actin-activated heavy meromyosin ATPase assays, the nonacetylated actin showed higher K(app) without significantly changing V(max), compared with those of wild-type actin. This is in contrast to the effect of the N-terminal negative charge cluster, which increases V(max) without changing K(app). These results indicate that the acetylation at the N-terminus of actin strengthens weak actomyosin interaction.  相似文献   

5.
Human monoamine (M) form and simple phenol (P) form phenol sulfotransferases (PSTs) are greater than 93% identical in their primary sequences and yet display distinct substrate specificities and other enzymatic properties. Through the generation and characterization of a series of chimeric PSTs, we have previously demonstrated two highly variable regions within their sequences to be responsible for determining their substrate phenotypes (Sakakibara, Y., Takami, Y., Nakayama, T., Suiko, M., and Liu, M.-C. (1998) J. Biol. Chem. 273, 6242-6247). By employing the site-directed mutagenesis technique, the present study aims to identify and quantitatively evaluate the specific amino acid residues critical to the substrate binding and catalysis in these two enzymes. Twelve mutated M-PSTs and seven mutated P-PSTs were generated, expressed, and purified. Enzymatic characterization showed that, of the twelve mutated M-PSTs, mutations at residues Asp-86, Glu-89, and Glu-146 resulted in a dramatic decrease in V(max)/K(m) with dopamine as substrate, being greater than 450 times for the D86A/E89I/E146A mutated M-PST. With p-nitrophenol as substrate, the V(max)/K(m) determined for the D86A/E89I/E146A-mutated M-PST increased more than 25 times and approached that determined for the wild-type P-PST. These results indicated that the concerted action of the three mutated residues (D86A, E89I, and E146A) is sufficient for the conversion of the substrate phenotype of M-PST to that of P-PST. Among the mutated P-PSTs, the I89E- and A146E-mutated P-PSTs displayed considerable deviations in V(max)/K(m) with dopamine or p-nitrophenol as substrate. No corresponding changes, however, were detected with the opposite compound as substrate. These results indicated that, in contrast to M-PST, mutations at Ala-86, Ile-89, and Ala-146 to the corresponding residues in M-PST are not sufficient for rendering the change of P-PST substrate phenotype to that of M-PST. For both M-PSTs and P-PSTs, mutations at Lys-48 or His-108 led to the loss of sulfotransferase activities, indicating their importance in the catalytic mechanism.  相似文献   

6.
Inositol 1,4,5-trisphosphate (IP(3)) receptors (IP(3)Rs) are large, ubiquitously expressed, endoplasmic reticulum membrane proteins that form tetrameric IP(3) and Ca(2+)-gated Ca(2+) channels. Endogenous IP(3)Rs provide very appealing tools for studying the ubiquitin-proteasome pathway in intact mammalian cells because, upon activation, they are rapidly ubiquitinated and degraded. Using mass spectrometry, we previously examined the ubiquitination of IP(3)R1 in αT3-1 pituitary gonadotrophs and found that IP(3)R1 ubiquitination is highly complex, with receptors being modified at multiple sites by monoubiquitin and polyubiquitin chains formed through both Lys-48 and Lys-63 linkages (Sliter, D. A., Kubota, K., Kirkpatrick, D. S., Alzayady, K. J., Gygi, S. P., and Wojcikiewicz, R. J. H. (2008) J. Biol. Chem. 283, 35319-35328). Here, we have extended these studies to determine whether IP(3)R2 and IP(3)R3 are similarly modified and if ubiquitination is cell type-dependent. Using mass spectrometry and linkage-specific ubiquitin antibodies, we found that all IP(3)R types are subject to ubiquitination at approximately the same locations and that, independent of cell type, IP(3)Rs are modified by monoubiquitin and Lys-48- and Lys-63-linked ubiquitin chains, although in differing proportions. Remarkably, the attached Lys-48- and Lys-63-linked ubiquitin chains are homogeneous and are segregated to separate IP(3)R subunits, and Lys-48-linked ubiquitin chains, but not Lys-63-linked chains, are required for IP(3)R degradation. Together, these data provide unique insight into the complexities of ubiquitination of an endogenous ubiquitin-proteasome pathway substrate in unperturbed mammalian cells. Importantly, although Lys-48-linked ubiquitin chains appear to trigger proteasomal degradation, the presence of Lys-63-linked ubiquitin chains suggests that ubiquitination of IP(3)Rs may have physiological consequences beyond signaling for degradation.  相似文献   

7.
8.
Huang X  Raushel FM 《Biochemistry》1999,38(48):15909-15914
Carbamoyl phosphate synthetase from Escherichia coli catalyzes the formation of carbamoyl phosphate from bicarbonate, glutamine, and two molecules of ATP. The enzyme consists of a large synthetase subunit, and a small amidotransferase subunit, which belongs to the Triad family of glutamine amidotransferases. Previous studies have established that the reaction mechanism of the small subunit proceeds through the formation of a gamma-glutamyl thioester with Cys-269. The roles in the hydrolysis of glutamine played by the conserved residues, Glu-355, Ser-47, Lys-202, and Gln-273, were determined by mutagenesis. In the X-ray crystal structure of the H353N mutant, Ser-47 and Gln-273 interact with the gamma-glutamyl thioester intermediate [Thoden, J. B., Miran, S. G., Phillips, J. C., Howard, A. J., Raushel, F. M., and Holden, H. M. (1998) Biochemistry 37, 8825-8831]. The mutants E355D and E355A have elevated values of K(m) for glutamine, but the overall carbamoyl phosphate synthesis reaction is unperturbed. E355Q does not significantly affect the bicarbonate-dependent ATPase or glutaminase partial reactions. However, this mutation almost completely uncouples the two partial reactions such that no carbamoyl phosphate is produced. The partial recovery of carbamoyl phosphate synthesis activity in the double mutant E355Q/K202M argues that the loss of activity in E355Q is at least partly due to additional interactions between Gln-355 and Lys-202 in E355Q. The mutants S47A and Q273A have elevated K(m) values for glutamine while the V(max) values are comparable to that of the wild-type enzyme. It is concluded that contrary to the original proposal for the catalytic triad, Glu-355 is not an essential residue for catalysis. The results are consistent with Ser-47 and Gln-273 playing significant roles in the binding of glutamine.  相似文献   

9.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

10.
R67 dihydrofolate reductase (R67 DHFR) is a novel protein encoded by an R-plasmid that confers resistance to the antibiotic, trimethoprim. This homotetrameric enzyme possesses 222 symmetry, which imposes numerous constraints on the single active site pore, including a "one-site-fits-both" strategy for binding its ligands, dihydrofolate (DHF) and NADPH. Previous studies uncovered salt effects on binding and catalysis (Hicks, S. N., Smiley, R. D., Hamilton, J. B., and Howell, E. E. (2003) Biochemistry 42, 10569-10578), however the one or more residues that participate in ionic contacts with the negatively charged tail of DHF as well as the phosphate groups in NADPH were not identified. Several studies predict that Lys-32 residues were involved, however mutations at this residue destabilize the R67 DHFR homotetramer. To study the role of Lys-32 in binding and catalysis, asymmetric K32M mutations have been utilized. To create asymmetry, individual mutations were added to a tandem array of four in-frame gene copies. These studies show one K32M mutation is tolerated quite well, whereas addition of two mutations has variable effects. Two double mutants, K32M:1+2 and K32M: 1+4, which place the mutations on opposite sides of the pore, reduce kcat. However a third double mutant, K32M: 1+3, that places two mutations on the same half pore, enhances kcat 4- to 5-fold compared with the parent enzyme, albeit at the expense of weaker binding of ligands. Because the kcat/Km values for this double mutant series are similar, these mutations appear to have uncovered some degree of non-productive binding. This non-productive binding mode likely arises from formation of an ionic interaction that must be broken to allow access to the transition state. The K32M:1+3 mutant data suggest this interaction is an ionic interaction between Lys-32 and the charged tail of dihydrofolate. This unusual catalytic scenario arises from the 222 symmetry imposed on the single active site pore.  相似文献   

11.
Actobindin is an 88-amino acid polypeptide, containing two almost identical repeated domains of 33 and 34 residues. Depending on the molar ratios in which they are mixed, actobindin binds either one or two actin molecules. We cross-linked actobindin and actin in the 1:1 complex, using the zero-length cross-linker 1-ethyl-3(3-dimethylaminopropyl)carbodiimide. The cross-linked peptides were purified after consecutive CNBr cleavage and trypsin and Staphylococcus protease V8 digestions, and the cross-linked side chains were identified by amino acid sequencing. Isopeptide linkages were formed between residues Glu-100 of actin and Lys-16 of actobindin. In addition, we found a connection between one or more of the acidic residues 1,2, or 3 of actin and Lys-16 and Lys-52 of actobindin. The cross-linked regions in actobindin contain Leu-Lys-His-Ala-Glu-Thr motifs, similar to sequences observed in several other actin-binding proteins.  相似文献   

12.
F Simeoni  L Masotti  P Neyroz 《Biochemistry》2001,40(27):8030-8042
Site-directed mutagenesis, gel filtration, and fluorescence spectroscopy approaches were used to study the molecular hinge mechanism involved in the beta-strand-exchanged dimer formation of the cyclin-dependent protein kinase regulatory subunit p13(suc1) from Schizosaccharomyces pombe. Single and double mutants of residues Pro-90 and Pro-92 (P90V, P92V, and P90V/P92V) were prepared and assayed. Substitution of Pro-90 prevented dimer formation by arm exchange. However, single point mutations did not affect the two-state unfolding transition of wild-type p13(suc1) at equilibrium (i.e., wild type, DeltaG degrees (0,un) = 7.38 +/- 0.35 kcal mol(-1), vs P90V, DeltaG degrees (0,un) = 6.71 +/- 0.18 kcal mol(-1)). On the contrary, the double mutant unfolded with a complex transition, and the reaction was best described by a three-state model (N <==> I <==> U). Resolution of the state-dependent (native vs denatured) intrinsic fluorescence decay amplitudes of p13(suc1) showed that with P90V/P92V these parameters were affected at [GuHCl] significantly less than with wild-type and single mutant proteins. Moreover, with the latter products, fluorescence quenching measurements at 1 M GuHCl revealed linear Stern-Volmer plots with quenching constants typical of tryptophan residues located in a native environment (1.6 M(-1) < K(SV) < 2.3 M(-1)). Dissimilarly, with P90V/P92V a significant deviation from linearity of the Stern-Volmer plot was obtained. Nonlinear least-squares analysis of these data resolved the significant contribution of highly solvent-accessible emitting species (K(SV) = 26 M(-1)) consistent with large exposure of the tryptophan residues. These results are compatible with the existence of an intermediate unfolding state of the double mutation product. Thus, while single residue substitution studies give support to the primary role of Pro-90 in the p13(suc1) dimer formation by domain swapping, double residue substitution studies indicate the important role of the conserved repeat, Pro-x-Pro, for the proper beta-strand spatial organization and stability.  相似文献   

13.
Site-directed mutagenesis studies of bovine pancreatic phospholipase A2 (PLA2, overproduced in Escherichia coli) showed that replacement of surface residue Lys-56 by a neutral or hydrophobic amino acid residue resulted in an unexpected and significant change in the function of the enzyme. The kcat for phosphatidylcholine micelles increases 3-4-fold for K56M, K56I, and K56F and ca. 2-fold for K56N and K56T but does not change for K56R. These results suggest that the side chain of residue 56 has significant influence on the activity of PLA2. In order to probe the structural basis for the enhanced activity, the crystal structures of wild-type and K56M PLA2 were determined by X-ray crystallography to a resolution of 1.8 A. The results suggest that the mutation has not only perturbed the conformation of the side chain of Met-56 locally but also caused conformational changes in the neighboring loop (residues 60-70), resulting in the formation of a hydrophobic pocket by residues Met-56, Tyr-52, and Tyr-69. Docking of a phosphatidylcholine inhibitor analogue into the active site of K56M, according to the structure of the complex of cobra venom PLA2-phosphatidylethanolamine inhibitor analogue [White, S.P., Scott, D. L., Otwinowski, Z., Gleb, M. H., & Sigler, P. (1990) Science 250, 1560-1563], showed that the choline moiety [N(CH3)3]+ is readily accommodated into the newly formed hydrophobic pocket with a high degree of surface complementarity. This suggests a possible interaction between residue 56 and the head group of the phospholipid, explaining the enhanced activities observed when the positively charged Lys-56 is substituted by apolar residues, viz., K56M, K56I, and K56F. Further support for this interpretation comes from the 5-fold enhancement in kcat for the mutant K56E with a negatively charged side chain, where there would be an attractive electrostatic interaction between the side chain of Glu-56 and the positively charged choline moiety. Our results also refute a recent report [Tomasselli, A. G., Hui, J., Fisher, J., Zürcher-Neely, H., Reardon, I.M., Oriaku, E., Kézdy, F.J., & Heinrikson, R.L. (1989) J. Biol. Chem. 264, 10041-10047] that substrate-level acylation of Lys-56 is an obligatory step in the catalysis by PLA2.  相似文献   

14.
Amino acids in the phosphate binding loop of adenylate kinase of Escherichia coli were mutated by site-directed mutagenesis. The mutant proteins with a Pro-9----Gly (P9G) and with a Lys-13----Gln (K13Q) exchange were overexpressed and purified. They were characterized by steady-state kinetics, fluorescence binding, and structural studies, together with the phosphate binding loop mutants P9L and G10V prepared earlier [Reinstein, J., Brune, M., & Wittinghofer, A. (1988) Biochemistry 27, 4712-4720]. The results obtained show that all these mutations change the structure of the protein as evidenced by NMR spectroscopy and temperature-stability studies. All the mutant proteins have increased dissociation constants for substrates and inhibitors, but their catalytic activity, except for K13Q, is not reduced. The results obtained with K13Q suggest that this lysine residue, which is conserved in all guanine and many adenine nucleotide proteins, might have an important role in catalysis.  相似文献   

15.
According to the Lorenz et al. (Lorenz, M., Poole, K. J., Popp, D., Rosenbaum, G., and Holmes, K. C. (1995) J. Mol. Biol. 246, 108-119) atomic model of the actin-tropomyosin complex, actin residue Asp-311 (Glu-311 in yeast) is predicted to have a high binding energy contribution to actin-tropomyosin binding. Using the yeast actin mutant E311A/R312A in the in vitro motility assays, we have investigated the role of these residues in such interactions. Wild type (wt) yeast actin, like skeletal alpha-actin, is fully regulated when complexed with tropomyosin (Tm) and troponin (Tn). Structure-function comparisons of the wt and E311A/R312A actins show no significant differences between them, and the unregulated F-actins slide at similar speeds in the in vitro motility assay. However, in the presence of Tm and Tn, the mutation increases both the sliding speed and the number of moving filaments at high pCa values, shifting the speed-pCa curve nearly 0.5 pCa units to the left. Tm alone (no Tn) inhibits the motilities of both actins at low heavy meromyosin densities but potentiates only the motility of the mutant actin at high heavy meromyosin densities. Actin-Tm binding measurements indicate no significant difference between wt and E311A/R312A actin in Tm binding. These results implicate allosteric effects in the regulation of actomyosin function by tropomyosin.  相似文献   

16.
Long chain curarimimetic toxins from snake venom bind with high affinities to both muscular type nicotinic acetylcholine receptors (AChRs) (K(d) in the pm range) and neuronal alpha 7-AChRs (K(d) in the nm range). To understand the molecular basis of this dual function, we submitted alpha-cobratoxin (alpha-Cbtx), a typical long chain curarimimetic toxin, to an extensive mutational analysis. By exploring 36 toxin mutants, we found that Trp-25, Asp-27, Phe-29, Arg-33, Arg-36, and Phe-65 are involved in binding to both neuronal and Torpedo (Antil, S., Servent, D., and Ménez, A. (1999) J. Biol. Chem. 274, 34851-34858) AChRs and that some of them (Trp-25, Asp-27, and Arg-33) have similar binding energy contributions for the two receptors. In contrast, Ala-28, Lys-35, and Cys-26-Cys-30 selectively bind to the alpha 7-AChR, whereas Lys-23 and Lys-49 bind solely to the Torpedo AChR. Therefore, alpha-Cbtx binds to two AChR subtypes using both common and specific residues. Double mutant cycle analyses suggested that Arg-33 in alpha-Cbtx is close to Tyr-187 and Pro-193 in the alpha 7 receptor. Since Arg-33 of another curarimimetic toxin is close to the homologous alpha Tyr-190 of the muscular receptor (Ackermann, E. J., Ang, E. T. H., Kanter, J. R., Tsigelny, I., and Taylor, P. (1998) J. Biol. Chem. 273, 10958-10964), toxin binding probably occurs in homologous regions of neuronal and muscular AChRs. However, no coupling was seen between alpha-Cbtx Arg-33 and alpha 7 receptor Trp-54, Leu-118, and Asp-163, in contrast to what was observed in a homologous situation involving another toxin and a muscular receptor (Osaka, H., Malany, S., Molles, B. E., Sine, S. M., and Taylor, P. (2000) J. Biol. Chem. 275, 5478-5484). Therefore, although occurring in homologous regions, the detailed modes of toxin binding to alpha 7 and muscular receptors are likely to be different. These data offer a molecular basis for the design of toxins with predetermined specificities for various members of the AChR family.  相似文献   

17.
Helical complexes formed between fd DNA and reductively methylated fd gene 5 protein were indistinguishable by electron microscopy from complexes formed with the nonmethylated protein. 13C NMR spectroscopy of 13C-enriched N epsilon, N epsilon-dimethyllsyl residues of the protein showed that three of these residues (Lys-24, Lys-46, and Lys-69) were selectively perturbed by binding of the oligomer d(pA)7. These were the same lysyl residues that we previously found to be most protected from methylation by binding of the protein to poly[r(U)] [Dick, L. R., Sherry, A. D., Newkirk, M. M., & Gray D. M. (1988) J. Biol. Chem. 263, 18864-18872]. Thus, these lysines are probably directly involved in the nucleic acid binding function of the protein. Negatively charged chelates of lanthanide ions were used to perturb the 13C NMR resonances of labeled lysyl and amino-terminal residues of the gene 5 protein. The terbium chelate was found to bind tightly (Ka approximately 10(5) M-1) to the protein with a stoichiometry of 1 chelate molecule per protein dimer. 13C resonances of Lys-24, Lys-46, and Lys-69 were maximally shifted by the terbium chelate and were maximally relaxed by the gadolinium chelate. Also, the terbium chelate was excluded by the oligomer d(pA)7. Computer fits of the induced chemical shifts of 13C resonances with those expected for various positions of the terbium chelate failed to yield a possible chelate binding site unless the chemical shift for Lys-24 was excluded from the fitting process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The investigation goal was to identify mechanisms for reversal of actin S-nitrosylation in neutrophils after exposure to high oxygen partial pressures. Prior work has shown that hyperoxia causes S-nitrosylated actin (SNO-actin) formation, which mediates β(2) integrin dysfunction, and these changes can be reversed by formylmethionylleucylphenylalanine or 8-bromo-cyclic GMP. Herein we show that thioredoxin reductase (TrxR) is responsible for actin denitrosylation. Approximately 80% of cellular TrxR is localized to the cytosol, divided between the G-actin and short filamentous actin (sF-actin) fractions based on Triton solubility of cell lysates. TrxR linkage to sF-actin requires focal adhesion kinase (FAK) based on immunoprecipitation studies. S-Nitrosylation accelerates actin filament turnover (by mechanisms described previously (Thom, S. R., Bhopale, V. M., Yang, M., Bogush, M., Huang, S., and Milovanova, T. (2011) Neutrophil β(2) integrin inhibition by enhanced interactions of vasodilator stimulated phosphoprotein with S-nitrosylated actin. J. Biol. Chem. 286, 32854-32865), which causes FAK to disassociate from sF-actin. TrxR subsequently dissociates from FAK, and the physical separation from actin impedes denitrosylation. If SNO-actin is photochemically reduced with UV light or if actin filament turnover is impeded by incubations with cytochalasin D, latrunculin B, 8-bromo-cGMP, or formylmethionylleucylphenylalanine, FAK and TrxR reassociate with sF-actin and cause SNO-actin removal. FAK-TrxR association can also be demonstrated using isolated enzymes in ex vivo preparations. Uniquely, the FAK kinase domain is the site of TrxR linkage. We conclude that through its scaffold function, FAK influences TrxR activity and actin S-nitrosylation.  相似文献   

19.
Green fluorescent proteins (GFP) are widely used in biology for tracking purposes. Although expression of GFP is considered to be innocuous for the cells, deleterious effects have been reported. We recently demonstrated that expression of eGFP in muscle impairs its contractile properties (Agbulut, O., Coirault, C., Niederlander, N., Huet, A., Vicart, P., Hagege, A., Puceat, M., and Menasche, P. (2006) Nat. Meth. 3, 331). This prompted us to identify the molecular mechanisms linking eGFP expression to contractile dysfunction and, particularly, to test the hypothesis that eGFP could inhibit actin-myosin interactions. Therefore, we assessed the cellular, mechanical, enzymatic, biochemical, and structural properties of myosin in the presence of eGFP and F-actin. In vitro motility assays, the maximum actin-activated ATPase rate (V(max)) and the associated constant of myosin for actin (K(m)) were determined at 1:0.5, 1:1, and 1:3 myosin:eGFP molar ratios. At a myosin:eGFP ratio of 1:0.5, there was a nearly 10-fold elevation of K(m). As eGFP concentration increased relative to myosin, the percentage of moving filaments, the myosin-based velocity, and V(max) significantly decreased compared with controls. Moreover, myosin co-precipitated with eGFP. Crystal structures of myosin, actin, and GFP indicated that GFP and actin exhibited similar electrostatic surface patterns and the ClusPro docking model showed that GFP bound preferentially to the myosin head and especially to the actin-binding site. In conclusion, our data demonstrate that expression of eGFP in muscle resulted in the binding of eGFP to myosin, thereby disturbing the actin-myosin interaction and in turn the contractile function of the transduced cells. This potential adverse effect of eGFP should be kept in mind when using this marker to track cells following transplantation.  相似文献   

20.
Modification of lysine residues with 4-chloro-3,5-dinitrobenzoate results in the loss of the binding capacity of K99 fibrillae to horse erythrocytes (Jacobs, A.A.C., van Mechelen, J.R. and de Graaf, F.K. (1985) Biochim. Biophys. Acta 832, 148-155). In the present study we used dinitrobenzoate as a spectral probe to map the modified residues. After the incorporation of 0.7 mol CDNB per mol subunit, 90% of the binding activity disappeared and the lysine residues at positions 87, 132 and 133 incorporated 20%, 27.5% and 52.2% of the totally incorporated label, respectively. In the presence of the glycolipid receptor, Lys-132 and Lys-133 were partially protected against modification, while Lys-87 was not protected. The results suggest that Lys-132 and Lys-133 are part of the receptor-binding domain of the K99 fibrillar subunit and that the positive charges on these residues are important for the interaction of the fibrillae with the negatively charged sialic acid residue of the glycolipid receptor. A striking homology was found between a six-amino-acid residue segment of K99, containing Lys-132 and Lys-133, and segments of three other sialic-acid-specific lectins; cholera toxin B subunit, heat-labile toxin B subunit of Escherichia coli and CFA1 fimbrial subunit, suggesting that these segments might also be part of the receptor-binding domain in these three proteins.  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