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 31 毫秒
1.
E Hazum 《Biochemistry》1987,26(22):7011-7014
The interaction of 125I-buserelin, a superactive agonist of gonadotropin-releasing hormone (GnRH), with solubilized GnRH receptor was studied. The highest specific binding of 125I-buserelin to solubilized GnRH receptor is evident at 4 degrees C, and equilibrium is reached after 2 h of incubation. The soluble receptor retained 100% of the original binding activity when kept at 4 or 22 degrees C for 60 min. Mono- and divalent cations inhibited, in a concentration-dependent manner, the binding of 125I-buserelin to solubilized GnRH receptor. Monovalent cations require higher concentrations than divalent cations to inhibit the binding. Since the order of potency within the divalent cations was identical with that of their association constants to dicarboxylic compounds, it is suggested that there are at least two carboxylic groups of the receptor that participate in the binding of the hormone. The carboxyl groups of sialic acid residues are not absolutely required for GnRH binding since the binding of 125I-buserelin to solubilized GnRH receptor was only slightly affected by pretreatment with neuraminidase and wheat germ agglutinin. The finding that polylysines stimulate luteinizing hormone (LH) release from pituitary cell cultures with the same efficacy as GnRH suggests that simple charge interactions can induce LH release. According to these results, we propose that the driving force for the formation of the hormone-receptor complex is an ionic interaction between the positively charged amino acid arginine in position 8 and the carboxyl groups in the binding site.  相似文献   

2.
The actions of divalent cations on voltage-gated ion channels suggest that these cations bind to specific sites and directly influence gating kinetics. We have examined some chemical properties of the external divalent cation binding sites on neuronal potassium channels. Patch clamp techniques were used to measure the electrophysiological properties of these channels and Zn ions were used to probe the divalent cation binding site. The channel activation kinetics were greatly (three- to fourfold) slowed by low (2-5 mM) concentrations of Zn; deactivation kinetics were only slightly affected. These effects of Zn were inhibited by low solution pH in a manner consistent with competition between Zn and H ions for a single site. The apparent inhibitory pK for this site was near 7.2. Treatment of the neurons with specific amino acid reagents implicated amino, but no histidyl or sulfhydryl, residues in divalent cation binding.  相似文献   

3.
藏红花凝集素分子化学修饰与其活性的关系   总被引:1,自引:0,他引:1  
对甘露糖专一性结合藏红花凝集素 (Crocussativuslectin ,CSL)分子进行化学修饰 ,测定酿酒酵母 (S .cerevisiae)凝集活性和寡糖专一性结合活性的变化 .实验结果表明 ,Cys的修饰与活性无关 ,Arg、Tyr和His的修饰降低了CSL分子的酵母凝集活性和寡糖结合活性 ,但对CSL的CD光谱无显著影响 ,表明其为凝集素的活性氨基酸残基 .Glu和Asp的化学修饰可使CSL的凝集活性大幅度降低 ,与特异性寡糖的亲和力增大 ,CD光谱变化明显 ,提示CSL分子中的Glu和Asp对其空间结构影响较大 ,氨基酸羧基的修饰导致CSL构象改变 ,蛋白与寡糖的结合位点暴露 ,可有效结合的位点数增加  相似文献   

4.
Several observations have already suggested that the carboxyl groups are involved in the association of divalent cations with bacteriorhodopsin (Chang et al., 1985). Here we show that at least part of the protons released from deionized purple membrane (`blue membrane') samples when salt is added are from carboxyl groups. We find that the apparent pK of magnesium binding to purple membrane in the presence of 0.5 mM buffer is 5.85. We suggest this is the pK of the carboxyl groups shifted from their usual pK because of the proton concentrating effect of the large negative surface potential of the purple membrane. Divalent cations may interact with negatively charged sites on the surface of purple membrane through the surface potential and/or through binding either by individual ligands or by conformation-dependent chelation. We find that divalent cations can be released from purple membrane by raising the temperature. Moreover, purple membrane binds only about half as many divalent cations after bleaching. Neither of these operations is expected to decrease the surface potential and thus these experiments suggest that some specific conformation in purple membrane is essential for the binding of a substantial fraction of the divalent cations. Divalent cations in purple membrane can be replaced by monovalent, (Na+ and K+), or trivalent, (La+++) cations. Flash photolysis measurements show that the amplitude of the photointermediate, O, is affected by the replacement of the divalent cations by other ions, especially by La+++. The kinetics of the M photointermediate and light-induced H+ uptake are not affected by Na+ and K+, but they are drastically lengthened by La+++ substitution, especially at alkaline pHs. We suggest that the surface charge density and thus the surface potential is controlled by divalent cation binding. Removal of the cations (to make deionized blue membrane) or replacement of them (e.g. La+++-purple membrane) changes the surface potential and hence the proton concentration near the membrane surface. An increase in local proton concentration could cause the protonation of critical carboxyl groups, for example the counter-ion to the protonated Schiff's base, causing the red shift associated with the formation of both deionized and acid blue membrane. Similar explanations based on regulation of the surface proton concentration can explain many other effects associated with the association of different cations with bacteriorhodopsin.  相似文献   

5.
Acyl carrier protein (ACP), a small protein essential for bacterial growth and pathogenesis, interacts with diverse enzymes during the biosynthesis of fatty acids, phospholipids, and other specialized products such as lipid A. NMR and hydrodynamic studies have previously shown that divalent cations stabilize native helical ACP conformation by binding to conserved acidic residues at two sites (A and B) at either end of the "recognition" helix II. To examine the roles of these amino acids in ACP structure and function, site-directed mutagenesis was used to replace individual site A (Asp-30, Asp-35, Asp-38) and site B (Glu-47, Glu-53, Asp-56) residues in recombinant Vibrio harveyi ACP with the corresponding amides, along with combined mutations at each site (SA, SB) or both sites (SA/SB). Like native V. harveyi ACP, all individual mutants were unfolded at neutral pH but adopted a helical conformation in the presence of millimolar Mg(2+) or upon fatty acylation. Mg(2+) binding to sites A or B independently stabilized native ACP conformation, whereas mutant SA/SB was folded in the absence of Mg(2+), suggesting that charge neutralization is largely responsible for ACP stabilization by divalent cations. Asp-35 in site A was critical for holo-ACP synthase activity, while acyl-ACP synthetase and UDP-N-acetylglucosamine acyltransferase (LpxA) activities were more affected by mutations in site B. Both sites were required for fatty acid synthase activity. Overall, our results indicate that divalent cation binding site mutations have predicted effects on ACP conformation but unpredicted and variable consequences on ACP function with different enzymes.  相似文献   

6.
The use of equilibrium dialysis techniques established that isolated cell walls of Bacillus subtilis possess selective affinities for several cations. The binding of these cations to the cell wall was influenced by the presence of various functional groups in the peptidoglycan matrix. Selective chemical modification of the free carboxyl and amino groups showed that when amino groups were replaced by neutral, bulky, or negatively charged groups, the sites available for cation complexing generally increased. Introduction of positive charges into the wall resulted in a marked decrease in the numbers of metal binding sites and usually a decrease in the apparent association constants. Both teichoic acid and peptidoglycan contribute to the sites available for interaction with metals. Hill plots of equilibrium dialysis data suggest that metal binding to cell walls involves negative cooperativity. Competition between various metals for binding sites suggested that the cations complex with identical sites on the cell walls. When the hydrogen ion concentration was increased, the affinity of the walls for metals decreased, but the numbers of metal binding sites remained constant, suggesting that cations and protons also compete for the same sites.  相似文献   

7.
Developing new biopolymer-based materials with bio-identical properties is a significant challenge in modern science. One interesting route to this goal involves the biomineralization of collagen, a pre-structured and widely available protein, into a material with interesting properties. A prerequisite for biomineralization is the ability of cations (e.g., calcium) to bind to the protein and to result in concert with appropriate anions (e.g., phosphate) in composite material with e.g., bone-like properties. In order to increase the number of binding sites it is necessary to modify the protein prior to mineralization. For this glucuronic acid (GA) was used due to its carbonyl and carboxyl groups to derivatize proteinogenic amino groups transferring them into negatively charged carboxyl groups. Our experiments showed for the first time, that Nepsilon-carboxymethyllysine is the major product of in vitro non-enzymatic glycosylation of collagen by glucuronic acid. For an unequivocal determination of the reaction products, the lysine residues of collagen and of the model peptide were carboxymethylated through a reductive alkylation with glyoxalic acid and compared to the glucuronic acid derivatives. Beside their identical mass spectra the common structure elements could be confirmed with FTIR. Thus, in the context of matrix engineering, by producing Nepsilon-carboxymethyllysine, glucuronic acid offers a convenient way of introducing additional stable acidic groups into protein matrices.  相似文献   

8.
To determine the contribution of charged amino acids to binding with the photosystem II complex (PSII), the amino or carboxyl groups of the extrinsic 18 kDa protein were modified with N- succinimidyl propionate (NSP) or glycine methyl ester (GME) in the presence of a water-soluble carbodiimide, respectively. Based on isoelectric point shift, 4-10 and 10-14 amino groups were modified in the presence of 2 and 4 mM NSP, respectively. Similarly, 3-4 carboxyl groups were modified by reaction with 100 mM GME. Neutralization of negatively charged carboxyl groups with GME did not alter the binding activity of the extrinsic 18 kDa protein. However, the NSP-modified 18 kDa protein, in which the positively charged amino groups had been modified to uncharged methyl esters, failed to bind with the PSII membrane in the presence of the extrinsic 23 kDa protein. This defect can not be attributed to structural or conformational alterations imposed by chemical modification, as the fluorescence and circular dichroism spectra among native, GME- and NSP-modified extrinsic 18 kDa proteins were similar. Thus, we have concluded that the positive charges of lysyl residues in the extrinsic 18 kDa protein are important for its interaction with PSII membranes in the presence of the extrinsic 23 kDa protein. Furthermore, it was found that the negative charges of carboxyl groups of this protein did not participate in binding with the extrinsic 23 kDa protein associated with PSII membranes.  相似文献   

9.
用各种化学试剂修饰红花菜豆(Phaseoluscoccineusvarrubronanus,Berry)凝集素(简称PCL)分子,测定与其活性相关的氨基酸残基.经NBS修饰表明PCL具有8个Trp残基,其中4个暴露于分子表面,此4个Trp残基被修饰后,PCL的凝血活性完全丧失.比较PCL修饰前后的CD光谱表明修饰不改变其二级结构。修饰Tyr,Arg,His残基和游离氨基及羧基不影响PCL的血凝活性.巯基也不是血凝活性所必需,但是PCL分子中的二硫键被还原,或被CNBr分解为两个片断则使蛋白质丧失血凝活性,提示分子的完整结构对PCL的血凝活力是重要的  相似文献   

10.
Stone EM  Chantranupong L  Georgiou G 《Biochemistry》2010,49(49):10582-10588
The active sites of eukaryotic arginase enzymes are strictly conserved, especially the first- and second-shell ligands that coordinate the two divalent metal cations that generate a hydroxide molecule for nucleophilic attack on the guanidinium carbon of l-arginine and the subsequent production of urea and l-ornithine. Here by using comprehensive pairwise saturation mutagenesis of the first- and second-shell metal ligands in human arginase I, we demonstrate that several metal binding ligands are actually quite tolerant to amino acid substitutions. Of >2800 double mutants of first- and second-shell residues analyzed, we found more than 80 unique amino acid substitutions, of which four were in first-shell residues. Remarkably, certain second-shell mutations could modulate the binding of both the nucleophilic water/hydroxide molecule and substrate or product ligands, resulting in activity greater than that of the wild-type enzyme. The data presented here constitute the first comprehensive saturation mutagenesis analysis of a metallohydrolase active site and reveal that the strict conservation of the second-shell metal binding residues in eukaryotic arginases does not reflect kinetic optimization of the enzyme during the course of evolution.  相似文献   

11.
The binding isotherms of Mn2+ to bovine plasma protein C (PC), des(1-41)-light chain protein C (GDPC), and activated GDPC (GDAPC) have been measured. PC contains 14-16 total Mn2+ binding sites, a value that is reduced to approximately 7-8 in the presence of NaCl. The average Kd of the latter sites is 230 +/- 30 microM. Upon removal of a 41-residue peptide from the amino terminus of the light chain of PC, and, concomitantly, all of the gamma-carboxyglutamic acid residues, the resulting protein, GDPC, possesses a single Mn2+ site of Kd = 120 +/- 20 microM. Activation of GDPC to GDAPC results in a slight lowering of the Kd for the single Mn2+ binding site to 53 +/- 8 microM, a value that is essentially unchanged in the presence of monovalent cations, a competitive inhibitor of the enzyme, or an active site directed affinity label. The Mn2+ on GDAPC is displaced by Ca2+, suggesting that the protein binding site for these two divalent cations is the same. These studies establish that Mn2+ is a suitable spectroscopic probe for the Ca2+ binding site of GDAPC, and that the divalent cation site is separate from the monovalent cation site(s) and the active site of the enzyme.  相似文献   

12.
The crystal structure of the binary complex of non-activated ribulose-1,5-bisphosphate carboxylase/oxygenase from Rhodospirillum rubrum and its product 3-phospho-D-glycerate has been determined to 2.9-A resolution. This structure determination confirms the proposed location of the active site (Schneider, G., Lindqvist, Y., Br?ndén, C.-I., and Lorimer, G. (1986) EMBO J. 5, 3409-3415) at the carboxyl end of the beta-strands of the alpha/beta-barrel in the carboxyl-terminal domain. One molecule of 3-phosphoglycerate is bound per active site. All oxygen atoms of 3-phosphoglycerate form hydrogen bonds to groups of the enzyme. The phosphate group interacts with the sidechains of residues Arg-288, His-321, and Ser-368, which are conserved between enzymes from different species as well as with the main chain nitrogens from residues Thr-322 and Gly-323. These amino acid residues constitute one of the two phosphate binding sites of the active site. The carboxyl group interacts with the side chains of His-287, Lys-191, and Asn-111. Implications of the activation process for the binding of 3-phosphoglycerate are discussed.  相似文献   

13.
The modulation of pentameric ligand-gated ion channels (pLGICs) by divalent cations is believed to play an important role in their regulation in a physiological context. Ions such as calcium or zinc influence the activity of pLGIC neurotransmitter receptors by binding to their extracellular domain and either potentiate or inhibit channel activation. Here we have investigated by electrophysiology and X-ray crystallography the effect of divalent ions on ELIC, a close prokaryotic pLGIC homologue of known structure. We found that divalent cations inhibit the activation of ELIC by the agonist cysteamine, reducing both its potency and, at higher concentrations, its maximum response. Crystal structures of the channel in complex with barium reveal the presence of several distinct binding sites. By mutagenesis we confirmed that the site responsible for divalent inhibition is located at the outer rim of the extracellular domain, at the interface between adjacent subunits but at some distance from the agonist binding region. Here, divalent cations interact with the protein via carboxylate side-chains, and the site is similar in structure to calcium binding sites described in other proteins. There is evidence that other pLGICs may be regulated by divalent ions binding to a similar region, even though the interacting residues are not conserved within the family. Our study provides structural and functional insight into the allosteric regulation of ELIC and is of potential relevance for the entire family.  相似文献   

14.
The effects of modifying individual lysyl, aspartyl, or glutamyl residues in calmodulin on its ability to bind to the neural phosphatase calcineurin have been investigated using a competitive binding method termed "label selection." Samples of calmodulin were radiochemically labeled at a low level (0.03-0.6 group/molecule) by acetylation of amino groups or coupling carboxyl groups with ethanolamine to produce preparations containing predominantly single-site modified and unmodified molecules. These preparations were incubated in a 5-10-fold molar excess with bovine calcineurin under conditions appropriate for complex formation. The bound population was isolated, and the level of modification of each reactive residue was compared with the level in the corresponding group in the intial unselected preparation to determine if molecules modified at specific sites had been selected for or against during the competition for complex formation. Significant selection was observed against molecules modified at Lys21, Asp64, Glu67, Lys75, Glu84, Glu114, Asp118, or Lys148, whereas modification of Glu83 increased binding. The modification of other groups, including components of the four Ca2+-binding sites, had no effect on the interaction. Glu67, a Ca2+-liganding residue in Ca2+-binding site II that may regulate the orientation of this site in relation to the central helix, had the strongest influence on complex formation. Most of the residues identified form a nearly linear array in the three-dimensional structure of calmodulin and indicate the location of an extended surface for interaction with calcineurin and other enzymes.  相似文献   

15.
Glycoprotein (GP) IIb is the alpha subunit of platelet integrin GPIIb-IIIa. Analysis of the primary structure of this subunit has indicated the presence of four stretches of amino acid residues that are highly conserved among various integrin alpha subunits and that have been suggested to be putative calcium-binding sites. To verify the Ca(2+)-binding capacity of these conserved domains and their implication in integrin adhesive functions, a fragment corresponding to the amino acid sequence of GPIIb from positions 171 to 464 was expressed. The nucleotide sequence coding for this GPIIb domain was generated by polymerase chain reaction, cloned into the pTG1924 expression vector, and expressed in Escherichia coli strain TGE901. The recombinant protein was purified by gel exclusion chromatography and used in equilibrium dialysis experiments. The results demonstrate that the four binding sites can be occupied by Ca2+. Two classes of binding sites can be detected, including two sites with a Kd of 30 microns and two sites of lower affinity with a Kd of 120 microns. Interaction of Ca2+ with these two classes of sites was inhibited by a large excess of Mg2+ or Mn2+, suggesting that these cations are competitive for the same sites on GPIIb. Thus, the four Ca(2+)-binding sites of GPIIb are not similar and exhibit different affinities for divalent ions. To verify the functional implication of these Ca(2+)-binding sites, the effect of Ca2+ on the binding of fibrinogen to the recombinant protein was analyzed using a solid-phase assay. The results indicate that optimal fibrinogen binding occurs when the four calcium-binding sites are occupied and establish the functional importance of this Ca(2+)-binding domain in the ligand-binding activity of GPIIb.  相似文献   

16.
1. The role of the ionized carboxyl groups of proteins of the erythrocyte membrane as Ca(2+) receptor sites was investigated. A water-soluble carbodi-imide [1-cyclohexyl-3-(2-morpholinoethyl)carbodi-imide methotoluene-p-sulphonate], referred to as carbodi-imide reagent, and glycine methyl ester were used to modify the free carboxyl groups of the membrane. The degree of modification was estimated from amino acid analyses, which showed the amount of glycine incorporated. As the concentration of carbodi-imide reagent was raised (0.1-0.4m) incorporation of glycine increased and Ca(2+) binding decreased by about 77%. At 0.4m-carbodi-imide reagent all of the binding of Ca(2+) to protein was abolished and it was estimated that about 37% of the side-chain carboxyl groups of aspartic acid plus glutamic acid had been blocked by glycine. 2. Acetylation of all of the free amino groups was achieved by incubating the erythrocyte ;ghosts' at pH10.3 with acetic anhydride (10-15mg/10mg of ;ghost' protein). Acetylation increased by 1.5-fold the capacity of the ;ghost' to bind Ca(2+), indicating that the remaining carboxyl groups of aspartic acid and glutamic acid were made available for Ca(2+) binding by this procedure. These findings support the concept that in normal ;ghosts', at pH7.4, Ca(2+) binding to free carboxyl groups is partially hindered by the presence of charged amino groups. 3. Treatment of ;ghosts' with N-acetylneuraminidase, which removed 94% of sialic acid residues, and treatment with 1mm-p-chloromercuribenzoate did not alter Ca(2+) binding. The major effect of 5.8mm-p-chloromercuribenzoate upon ;ghosts' was to cause a solubilization of a calcium-membrane complex, which included about one-third of the ;ghost' protein. The molar ratio of Ca(2+): protein in the solubilized material was the same as that in the intact (untreated) ;ghosts'.  相似文献   

17.
In this study, the interaction of human serum low-density lipoprotein (LDL) with heparin immobilized on Sepharose was reinvestigated. Binding of isolated LDL (stabilized with human serum albumin (HSA] was compared with that of LDL in full serum. (1) Binding of isolated LDL was slightly decreased by CaCl2 and was not affected by MgCl2. In contrast, with full serum LDL binding was increased by these divalent cations. (2) In both situations, binding of LDL was saturable, but the maximum degree of binding that could be reached was much higher with isolated LDL than with LDL in full serum. This could be ascribed to an inhibitory action of a factor found in the d greater than 1.24 fraction of serum. (3) The effect of this factor was diminished in the presence of CaCl2 or MgCl2, which suggests that the stimulation of LDL binding by these cations in full serum is due to suppression of the inhibitory activity of this factor. (4) The inhibitory factor in the d greater than 1.24 fraction can be partially purified by absorption to heparin-Sepharose, followed by elution with 6 M guanidine chloride. The resulting preparation had a 30- to 50-fold higher specific activity. Attempts to purify the factor further resulted in loss of activity. (5) The activity is decreased upon treatment with trypsin and also upon acetylation or reduction with dithiothreitol, indicating that free amino groups and S-S bridges are essential.  相似文献   

18.
We report the discovery and characterization of three conantokin peptides from the venom of Conus parius. Each peptide (conantokin-Pr1, -Pr2, and -Pr3) contains 19 amino acids with three gamma-carboxyglutamate (Gla) residues, a post-translationally modified amino acid characteristic of conantokins. The new peptides contain several amino acid residues that differ from previous conantokin consensus sequences. Notably, the new conantokins lack Gla at the 3rd position from the N terminus, where the Gla residue is replaced by either aspartate or by another post-translationally modified residue, 4-trans-hydroxyproline. Conantokin-Pr3 is the first conantokin peptide to have three different post-translational modifications. Conantokins-Pr1 and -Pr2 adopt alpha-helical conformations in the presence of divalent cations (Mg2+ and Ca2+) but are generally unstructured in the absence of divalent cations. Conantokin-Pr3 adopts an alpha-helical conformation even in the absence of divalent cations. Like other conantokins, the new peptides induced sleep in young mice and hyperactivity in older mice upon intracranial injection. Electrophysiological assays confirmed that conantokins-Pr1, -Pr2, and -Pr3 are N-methyl-d-aspartate (NMDA) receptor antagonists, with highest potency for NR2B-containing NMDA receptors. Conantokin-Pr3 demonstrated approximately 10-fold selectivity for NR2B-containing NMDA receptors. However, conantokin-Pr2 showed minimal differences in potency between NR2B and NR2D. Conantokins-Pr1, -Pr2, and -Pr3 all demonstrated high specificity of block for NMDA receptors, when tested against various ligand-gated ion channels. Conus parius conantokins allow for a better definition of structural and functional features of conantokins as ligands targeting NMDA receptors.  相似文献   

19.
The effect of divalent cations on the near ultraviolet circular dichroism (CD) spectrum of yeast enolase showed that calcium, magnesium, and nickel ions produced identical changes. This was interpreted as indicating that the cations bound to the same sites on the enzyme and produced identical changes in tertiary structure. There was no effect of magnesium ion on the far ultraviolet spectrum. Evidently magnesium ion has no effect on the secondary structure. Substrate bound to the enzyme when the above cations were present although calcium permits no enzymatic activity. The CD spectral difference produced by the substrate was nearly the reverse of that produced by the metal ions. Glycolic acid phosphate, a competitive inhibitor lacking carbon-3, produced no effect, indicating carbon-3 was necessary for the CD spectral changes. The CD and visible absorption spectra of nickel and cobalt bound to various sites on the enzyme showed that the binding sites were octahedral or distorted octahedral in coordination and that the ligands appeared to be oxyligands: water molecules, hydroxyl or carboxyl groups. Examination of the effects of substrate and two compounds thought to be "transition state analogues" showed that these perturbed the "conformational" sites of the enzyme. The "catalytic" and "inhibitory" sites did not appear to be very CD active.  相似文献   

20.
Herpes simplex virus 1 (HSV-1) ICP8 is a single-stranded DNA-binding protein that is necessary for viral DNA replication and exhibits recombinase activity in vitro. Alignment of the HSV-1 ICP8 amino acid sequence with ICP8 homologs from other herpesviruses revealed conserved aspartic acid (D) and glutamic acid (E) residues. Amino acid residue D1087 was conserved in every ICP8 homolog analyzed, indicating that it is likely critical for ICP8 function. We took a genetic approach to investigate the functions of the conserved ICP8 D and E residues in HSV-1 replication. The E1086A D1087A mutant form of ICP8 failed to support the replication of an ICP8 mutant virus in a complementation assay. E1086A D1087A mutant ICP8 bound DNA, albeit with reduced affinity, demonstrating that the protein is not globally misfolded. This mutant form of ICP8 was also recognized by a conformation-specific antibody, further indicating that its overall structure was intact. A recombinant virus expressing E1086A D1087A mutant ICP8 was defective in viral replication, viral DNA synthesis, and late gene expression in Vero cells. A class of enzymes called DDE recombinases utilize conserved D and E residues to coordinate divalent metal cations in their active sites. We investigated whether the conserved D and E residues in ICP8 were also required for binding metal cations and found that the E1086A D1087A mutant form of ICP8 exhibited altered divalent metal binding in an in vitro iron-induced cleavage assay. These results identify a novel divalent metal cation-binding site in ICP8 that is required for ICP8 functions during viral replication.  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