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 15 毫秒
1.
MICA is a major histocompatibility complex-like protein that undergoes a structural transition from disorder to order upon binding its immunoreceptor, NKG2D. We redesigned the disordered region of MICA with RosettaDesign to increase NKG2D binding. Mutations that stabilize this region were expected to increase association kinetics without changing dissociation kinetics, increase affinity of interaction, and reduce entropy loss upon binding. MICA mutants were stable in solution, and they were amenable to surface plasmon resonance evaluation of NKG2D binding kinetics and thermodynamics. Several MICA mutants bound NKG2D with enhanced affinity, kinetic changes were primarily observed during association, and thermodynamic changes in entropy were as expected. However, none of the 15 combinations of mutations predicted to stabilize the receptor-bound MICA conformation enhanced NKG2D affinity, whereas all 10 mutants predicted to be destabilized bound NKG2D with increased on-rates. Five of these had affinities enhanced by 0.9-1.8 kcal/mol over wild type by one to three non-contacting substitutions. Therefore, in this case, mutations designed to mildly destabilize a protein enhanced association and affinity.  相似文献   

2.
Marble brain disease (MBD) also known as Guibaud-Vainsel syndrome is caused by autosomal recessive mutations in the human carbonic anhydrase II (HCA II) gene. HCA II is a 259 amino acid single domain enzyme and is dominated by a 10-stranded beta-sheet. One mutation associated with MBD entails the H107Y substitution where H107 is a highly conserved residue in the carbonic anhydrase protein family. We have previously demonstrated that the H107Y mutation is a remarkably destabilizing folding mutation [Almstedt et al. (2004) J. Mol. Biol. 342, 619-633]. Here, the exceptional destabilization by the H107Y mutation has been further investigated. A mutational survey of position H107 and a neighboring conserved position E117 has been performed entailing the mutants H107A, H107F, H107N, E117A and the double mutants H107A/E117A and H107N/E117A. All mutants were severely destabilized versus GuHCl and heat denaturation. Thermal denaturation and GuHCl phase diagram and ANS analyses showed that the mutants shifted HCA II toward populating ensembles of intermediates of molten globule type under physiological conditions. The native state stability of the mutants was in the following order: wt > H107N > E117A > H107A > H107F > H107Y > H107N/E117A > H107A/E117A. In conclusion: (i) H107N is least destabilizing likely due to compensatory H-bonding ability of the introduced Asn residue. (ii) Double mutant cycles surprisingly reveal additive destabilization of H107N and E117A showing that H107 and E117 are independently stabilizing the folded protein. (iii) H107Y and H107F are exceptionally destabilizing due to bulkiness of the side chains whereas H107A is more accommodating, indicating long-range destabilizing effects of the natural pathogenic H107Y mutation.  相似文献   

3.
Nine single-cysteine mutants were labeled with 5-(2-iodoacetylaminoethylamino)naphthalene-1-sulfonic acid, an efficient acceptor of Trp fluorescence in fluorescence resonance energy transfer. The ratio between the fluorescence intensity of the 5-(2-acetylaminoethylamino)naphthalene-1-sulfonic acid (AEDANS) moiety excited at 295 nm (Trp absorption) and 350 nm (direct AEDANS absorption) was used to estimate the average distances between the seven Trp residues in human carbonic anhydrase II (HCA II) and the AEDANS label. Guanidine HCl denaturation of the HCA II variants was also performed to obtain a curve that reflected the compactness of the protein at various stages of the unfolding, which could serve as a scale of the expansion of the protein. This approach was developed in this study and was used to estimate the compactness of HCA II during heat denaturation and interaction with GroEL. It was shown that thermally induced unfolding of HCA II proceeded only to the molten globule state. Reaching this state was sufficient to allow HCA II to bind to GroEL, and the volume of the molten globule intermediate increased approximately 2.2-fold compared with that of the native state. GroEL-bound HCA II expands to a volume three to four times that of the native state (to approximately 117,000 A(3)), which correlates well with a stretched and loosened-up HCA II molecule in an enlarged GroEL cavity. Recently, we found that HCA II binding causes such an inflation of the GroEL molecule, and this probably represents the mechanism by which GroEL actively stretches its protein substrates apart (Hammarstr?m, P., Persson, M., Owenius, R., Lindgren, M., and Carlsson, U. (2000) J. Biol. Chem. 275, 22832-22838), thereby facilitating rearrangement of misfolded structure.  相似文献   

4.
Cancer-associated mutations in the BRCT domain of BRCA1 (BRCA1-BRCT) abolish its tumor suppressor function by disrupting interactions with other proteins such as BACH1. Many cancer-related mutations do not cause sufficient destabilization to lead to global unfolding under physiological conditions, and thus abrogation of function probably is due to localized structural changes. To explore the reasons for mutation-induced loss of function, the authors performed molecular dynamics simulations on three cancer-associated mutants, A1708E, M1775R, and Y1853ter, and on the wild type and benign M1652I mutant, and compared the structures and fluctuations. Only the cancer-associated mutants exhibited significant backbone structure differences from the wild-type crystal structure in BACH1-binding regions, some of which are far from the mutation sites. Backbone differences of the A1708E mutant from the liganded wild type structure in these regions are much larger than those of the unliganded wild type X-ray or molecular dynamics structures. These BACH1-binding regions of the cancer-associated mutants also exhibited increases in their fluctuation magnitudes compared with the same regions in the wild type and M1562I mutant, as quantified by quasiharmonic analysis. Several of the regions of increased fluctuation magnitude correspond to correlated motions of residues in contact that provide a continuous path of fluctuating amino acids in contact from the A1708E and Y1853ter mutation sites to the BACH1-binding sites with altered structure and dynamics. The increased fluctuations in the disease-related mutants suggest an increase in vibrational entropy in the unliganded state that could result in a larger entropy loss in the disease-related mutants upon binding BACH1 than in the wild type. To investigate this possibility, vibrational entropies of the A1708E and wild type in the free state and bound to a BACH1-derived phosphopeptide were calculated using quasiharmonic analysis, to determine the binding entropy difference DeltaDeltaS between the A1708E mutant and the wild type. DeltaDeltaS was determined to be -4.0 cal mol(-1) K(-1), with an uncertainty of 2 cal mol(-1) K(-1); that is, the entropy loss upon binding the peptide is 4.0 cal mol(-1) K(-1) greater for the A1708E mutant, corresponding to an entropic contribution to the DeltaDeltaG of binding (-TDeltaDeltaS) 1.1 kcal mol(-1) more positive for the mutant. The observed differences in structure, flexibility, and entropy of binding likely are responsible for abolition of BACH1 binding, and illustrate that many disease- related mutations could have very long-range effects. The methods described here have potential for identifying correlated motions responsible for other long-range effects of deleterious mutations.  相似文献   

5.
T7 phage DNA polymerase is a tight 1:1 complex of the gene 5 protein (g5p) (80 kDa) of phage T7 and thioredoxin (12 kDa) from the Escherichia coli host. The holoenzyme is essential for the replication of the phage. We estimated the real-time kinetics and thermodynamics of the interaction of g5p with thioredoxin (wild type and mutants) using surface plasmon resonance. Thioredoxin was immobilized on a CM5 sensor chip through a six-carbon spacer (6-amino-n-hexanoic acid) using standard amine coupling. Reduced thioredoxin bound g5p but oxidized thioredoxin did not. The association and dissociation phases of the complex fit a two-exponential model with an apparent equilibrium dissociation constant (KD) of 2.2 nm for thioredoxin with 4.7 x 104.M-1.s-1 and 10.5 x 10-5.s-1 as the corresponding association (ka) and dissociation (kd) rate constants. The strong binding of g5p to thioredoxin is therefore due to fast association and very slow dissociation, a situation similar to antigen-antibody interactions. Thioredoxin mutants P34S, D26A, K57M, D26A/K57M, W31F, W31Y, K36A, K36E, and Y49F had KD values in the range of 1 to 8 nm, whereas mutant W28A had a KD of 12.5 nm. No detectable interaction was observed for mutants P40G, W31H, W31A, and C35A. The effect of temperature on KD and the changes in enthalpy (-DeltaH = 20.2 kcal.m-1) and entropy (TDeltaS =-8.4 kcal.m-1) upon formation of the complex suggested that the interaction is driven by an increase in enthalpy and opposed by a decrease in entropy.  相似文献   

6.
Mutation of the ubiquitous cytosolic enzyme Cu/Zn superoxide dismutase (SOD1) is hypothesized to cause familial amyotrophic lateral sclerosis (FALS) through structural destabilization leading to misfolding and aggregation. Considering the late onset of symptoms as well as the phenotypic variability among patients with identical SOD1 mutations, it is clear that nongenetic factor(s) impact ALS etiology and disease progression. Here we examine the effect of Cys-111 glutathionylation, a physiologically prevalent post-translational oxidative modification, on the stabilities of wild type SOD1 and two phenotypically diverse FALS mutants, A4V and I112T. Glutathionylation results in profound destabilization of SOD1(WT) dimers, increasing the equilibrium dissociation constant K(d) to ~10-20 μM, comparable to that of the aggressive A4V mutant. SOD1(A4V) is further destabilized by glutathionylation, experiencing an ~30-fold increase in K(d). Dissociation kinetics of glutathionylated SOD1(WT) and SOD1(A4V) are unchanged, as measured by surface plasmon resonance, indicating that glutathionylation destabilizes these variants by decreasing association rate. In contrast, SOD1(I112T) has a modestly increased dissociation rate but no change in K(d) when glutathionylated. Using computational structural modeling, we show that the distinct effects of glutathionylation on different SOD1 variants correspond to changes in composition of the dimer interface. Our experimental and computational results show that Cys-111 glutathionylation induces structural rearrangements that modulate stability of both wild type and FALS mutant SOD1. The distinct sensitivities of SOD1 variants to glutathionylation, a modification that acts in part as a coping mechanism for oxidative stress, suggest a novel mode by which redox regulation and aggregation propensity interact in ALS.  相似文献   

7.
Clausen JD  Andersen JP 《Biochemistry》2003,42(9):2585-2594
Point mutants with alterations to Leu249, Lys252, Leu253, Asp254, and Glu255 in membrane segment M3, and Pro824, Lys825, and Glu826 in loop L6-7, of the sarcoplasmic reticulum Ca2+-ATPase were analyzed functionally by steady-state and transient kinetic methods. In mutants Leu249Ala, Lys252Glu, and Leu253Ala, the rate of Ca2+ dissociation from the cytoplasmically facing high-affinity Ca2+ sites was increased 4- to 7-fold relative to wild type, and in Leu249Ala and Lys252Glu the rate of Ca2+ binding was increased as well. Substitution of Lys252 with arginine, alanine, glutamine, or methionine affected Ca2+ interaction much less, indicating that the negative charge of the glutamate is particularly disturbing. These findings may be understood on the basis of the hypothesis that a water-accessible channel leading between membrane segments M1 and M3 in the thapsigargin-bound Ca2+-free structure [Toyoshima, C., and Nomura, H. (2002) Nature 418, 605-611] is closely related to the migration pathway for Ca2+. The effects of alanine mutations to Leu249 and Leu253 on Ca2+ dissociation may arise from destabilization of the hydrophobic wall lining the pathway. In mutant Lys252Glu, unfavorable interaction between the glutamate and L6-7 may open the pathway. In addition, Leu253Ala, and to a lesser extent some of the other mutations, reduced the rate of the E1PCa2 to E2P transition of the phosphoenzyme, enhanced the rate of dephosphorylation of E2P, and reduced the apparent affinity for vanadate, suggesting interference with the conformational change of the phosphoenzyme and the function of the catalytic site in E2 and E2P.  相似文献   

8.
Nitric oxide (NO) release from nitric oxide synthases (NOSs) is largely dependent on the dissociation of an enzyme ferric heme-NO product complex (Fe(III)NO). Although the NOS-like protein from Bacillus subtilis (bsNOS) generates Fe(III)NO from the reaction intermediate N-hydroxy-l-arginine (NOHA), its NO dissociation is about 20-fold slower than in mammalian NOSs. Crystal structures suggest that a conserved Val to Ile switch near the heme pocket of bsNOS might determine its kinetic profile. To test this we generated complementary mutations in the mouse inducible NOS oxygenase domain (iNOSoxy, V346I) and in bsNOS (I224V) and characterized the kinetics and extent of their NO synthesis from NOHA and their NO-binding kinetics. The mutations did not greatly alter binding of Arg, (6R)-tetrahydrobiopterin, or alter the electronic properties of the heme or various heme-ligand complexes. Stopped-flow spectroscopy was used to study heme transitions during single turnover NOHA reactions. I224V bsNOS displayed three heme transitions involving four species as typically occurs in wild-type NOS, the beginning ferrous enzyme, a ferrous-dioxy (Fe(II)O(2)) intermediate, Fe(III)NO, and an ending ferric enzyme. The rate of each transition was increased relative to wild-type bsNOS, with Fe(III)NO dissociation being 3.6 times faster. In V346I iNOSoxy we consecutively observed the beginning ferrous, Fe(II)O(2), a mixture of Fe(III)NO and ferric heme species, and ending ferric enzyme. The rate of each transition was decreased relative to wild-type iNOSoxy, with the Fe(III)NO dissociation being 3 times slower. An independent measure of NO binding kinetics confirmed that V346I iNOSoxy has slower NO binding and dissociation than wild-type. Citrulline production by both mutants was only slightly lower than wild-type enzymes, indicating good coupling. Our data suggest that a greater shielding of the heme pocket caused by the Val/Ile switch slows down NO synthesis and NO release in NOS, and thus identifies a structural basis for regulating these kinetic variables.  相似文献   

9.
We measured the kinetics of calcium dissociation from calsequestrin in solution or forming part of isolated junctional sarcoplasmic reticulum membranes by mixing calsequestrin equilibrated with calcium with calcium-free solutions in a stopped-flow system. In parallel, we measured the kinetics of the intrinsic fluorescence changes that take place following calcium dissociation from calsequestrin. We found that at 25 degrees C calcium dissociation was 10-fold faster for calsequestrin attached to junctional membranes (k = 109 s(-1)) than in solution. These results imply that calcium dissociation from calsequestrin in vivo is not rate limiting during excitation-contraction coupling. In addition, we found that the intrinsic fluorescence decrease for calsequestrin in solution or forming part of junctional membranes was significantly slower than the rates of calcium dissociation. The kinetics of intrinsic fluorescence changes had two components for calsequestrin associated to junctional membranes and only one for calsequestrin in solution; the faster component was 8-fold faster (k = 54.1 s(-1)) than the slower component (k = 6.9 s(-1)), which had the same k value as for calsequestrin in solution. These combined results suggest that the presence of calsequestrin at high concentrations in a restricted space, such as when bound to the junctional membrane, accelerates calcium dissociation and the resulting structural changes, presumably as a result of cooperative molecular interactions.  相似文献   

10.
We characterized electron transfer (ET) from putidaredoxin (Pdx) to the mutants of cytochrome P450(cam) (P450(cam)), in which one of the residues located on the putative binding site to Pdx, Gln360, was replaced with Glu, Lys, and Leu. The kinetic analysis of the ET reactions from reduced Pdx to ferric P450(cam) (the first ET) and to ferrous oxygenated P450(cam) (the second ET) showed the dissociation constants (K(m)) that were moderately perturbed for the Lys and Leu mutants and the distinctly increased for the Glu mutant. Although the alterations in K(m) indicate that Gln360 is located at the Pdx binding site, the effects of the Gln360 mutations (0.66-20-fold of that of wild type) are smaller than those of the Arg112 mutants (25-2500-fold of that of wild type) [Unno, M., et al. (1996) J. Biol. Chem. 271, 17869-17874], allowing us to conclude that Gln360 much less contributes to the complexation with Pdx than Arg112. The first ET rate (35 s(-1) for wild-type P450(cam)) was substantially reduced in the Glu mutant (5.4 s(-1)), while less perturbation was observed for the Lys (53 s(-1)) and Leu (23 s(-1)) mutants. In the second ET reaction, the retarded ET rate was detected only in the Glu mutant but not in the Lys and Leu mutants. These results showed the smaller mutational effects of Gln360 on the ET reactions than those of the Arg112 mutants. In contrast to the moderate perturbations in the kinetic parameters, the mutations at Gln360 significantly affected both the standard enthalpy and entropy of the redox reaction of P450(cam), which cause the negative shift of the redox potentials for the Fe(3+)/Fe(2+) couple by 20-70 mV. Since the amide group of Gln360 is located near the carbonyl oxygen of the amide group of the axial cysteine, it is plausible that the mutation at Gln360 perturbs the electronic interaction of the axial ligand with heme iron, resulting in the reduction of the redox potentials. We, therefore, conclude that Gln360 primarily regulates the ET reaction of P450(cam) by modulating the redox potential of the heme iron and not by the specific interaction with Pdx or the formation of the ET pathway that are proposed as the regulation mechanism of Arg112.  相似文献   

11.
The Hsp60-type chaperonin GroEL assists in the folding of the enzyme human carbonic anhydrase II (HCA II) and protects it from aggregation. This study was aimed to monitor conformational rearrangement of the substrate protein during the initial GroEL capture (in the absence of ATP) of the thermally unfolded HCA II molten-globule. Single- and double-cysteine mutants were specifically spin-labeled at a topological breakpoint in the β-sheet rich core of HCA II, where the dominating antiparallel β-sheet is broken and β-strands 6 and 7 are parallel. Electron paramagnetic resonance (EPR) was used to monitor the GroEL-induced structural changes in this region of HCA II during thermal denaturation. Both qualitative analysis of the EPR spectra and refined inter-residue distance calculations based on magnetic dipolar interaction show that the spin-labeled positions F147C and K213C are in proximity in the native state of HCA II at 20 °C (as close as ~8 Å), and that this local structure is virtually intact in the thermally induced molten-globule state that binds to GroEL. In the absence of GroEL, the molten globule of HCA II irreversibly aggregates. In contrast, a substantial increase in spin–spin distance (up to >20 Å) was observed within minutes, upon interaction with GroEL (at 50 and 60 °C), which demonstrates a GroEL-induced conformational change in HCA II. The GroEL binding-induced disentanglement of the substrate protein core at the topological break-point is likely a key event for rearrangement of this potent aggregation initiation site, and hence, this conformational change averts HCA II misfolding.  相似文献   

12.
The first stage in hemostasis is the binding of the platelet membrane receptor glycoprotein (GP) Ib-IX complex to the A1 domain of von Willebrand factor in the subendothelium. A bleeding disorder associated with this interaction is platelet-type von Willebrand disease, which results from gain-of-function (GOF) mutations in amino acid residues 233 or 239 of the GP Ibalpha subunit of GP Ib-IX. Using optical tweezers and a quadrant photodetector, we investigated the binding of A1 to GOF and loss-of-function mutants of GP Ibalpha with mutations in the region containing the two known naturally occurring mutations. By dynamically measuring unbinding force profiles at loading rates ranging from 200-20,000 pN/s, we found that the bond strengths between A1 and GP Ibalpha GOF mutants (233, 235, 237, and 239) were significantly greater than the A1/wild-type GP Ib-IX bond at all loading rates examined (p < 0.05). In addition, mutants 231 and 232 exhibited significantly lower bond strengths with A1 than the wild-type receptors (p < 0.05). We computed unloaded dissociation rate constant (k(off)(0)) values for interactions involving mutant and wild-type GP Ib-IX receptors with A1 and found the A1/wild-type GP Ib-IX k(off)(0) value of 5.47 +/- 0.25 s(-1) to be significantly greater than the GOF k(off)(0) values and significantly less than the loss-of-function k(off)(0) values. Our data illustrate the importance of the bond kinetics associated with the VWF/GP Ib-IX interaction in hemostasis and also demonstrate the drastic changes in binding that can occur when only a single amino acid of GP Ibalpha is altered.  相似文献   

13.
In the catalysis of the hydration of carbon dioxide and dehydration of bicarbonate by human carbonic anhydrase II (HCA II), a histidine residue (His64) shuttles protons between the zinc-bound solvent molecule and the bulk solution. To evaluate the effect of the position of the shuttle histidine and pH on proton shuttling, we have examined the catalysis and crystal structures of wild-type HCA II and two double mutants: H64A/N62H and H64A/N67H HCA II. His62 and His67 both have their side chains extending into the active-site cavity with distances from the zinc approximately equivalent to that of His64. Crystal structures were determined at pH 5.1-10.0, and the catalysis of the exchange of (18)O between CO(2) and water was assessed by mass spectrometry. Efficient proton shuttle exceeding a rate of 10(5) s(-)(1) was observed for histidine at positions 64 and 67; in contrast, relatively inefficient proton transfer at a rate near 10(3) s(-)(1) was observed for His62. The observation, in the crystal structures, of a completed hydrogen-bonded water chain between the histidine shuttle residue and the zinc-bound solvent does not appear to be required for efficient proton transfer. The data suggest that the number of intervening water molecules between the donor and acceptor supporting efficient proton transfer in HCA II is important, and furthermore suggest that a water bridge consisting of two intervening water molecules is consistent with efficient proton transfer.  相似文献   

14.
Cardiac troponin C belongs to the EF-hand superfamily of calcium-binding proteins and plays an essential role in the regulation of muscle contraction and relaxation. To follow calcium binding and exchange with the regulatory N-terminal domain (N-domain) of human cardiac troponin C, we substituted Phe at position 27 with Trp, making a fluorescent cardiac troponin C(F27W). Trp(27) accurately reported the kinetics of calcium association and dissociation of the N-domain of cardiac troponin C(F27W). To sensitize the N-domain of cardiac troponin C(F27W) to calcium, we individually substituted the hydrophobic residues Phe(20), Val(44), Met(45), Leu(48), and Met(81) with polar Gln. These mutations were designed to increase the calcium affinity of the N-domain of cardiac troponin C by facilitating the movement of helices B and C (BC unit) away from helices N, A, and D (NAD unit). As anticipated, these selected hydrophobic residue substitutions increased the calcium affinity of the regulatory domain of cardiac troponin C(F27W) approximately 2.1-15.2-fold. Surprisingly, the increased calcium affinity caused by the hydrophobic residue substitutions was largely due to faster calcium association rates (2.6-8.7-fold faster) rather than to slower calcium dissociation rates (1.2-2.9-fold slower). The regulatory N-domains of cardiac troponin C(F27W) and its mutants were also able to bind magnesium competitively and with physiologically relevant affinities (1.2-2.7 mm). The design of calcium-sensitizing cardiac troponin C mutants presented in this work enhances the understanding of how to control cation binding properties of EF-hand proteins and ultimately their structure and physiological function.  相似文献   

15.
Zheng J  Avvaru BS  Tu C  McKenna R  Silverman DN 《Biochemistry》2008,47(46):12028-12036
Catalysis by the zinc metalloenzyme human carbonic anhydrase II (HCA II) is limited in maximal velocity by proton transfer between His64 and the zinc-bound solvent molecule. Asn62 extends into the active site cavity of HCA II adjacent to His64 and has been shown to be one of several hydrophilic residues participating in a hydrogen-bonded solvent network within the active site. We compared several site-specific mutants of HCA II with replacements at position 62 (Ala, Val, Leu, Thr, and Asp). The efficiency of catalysis in the hydration of CO 2 for the resulting mutants has been characterized by (18)O exchange, and the structures of the mutants have been determined by X-ray crystallography to 1.5-1.7 A resolution. Each of these mutants maintained the ordered water structure observed by X-ray crystallography in the active site cavity of wild-type HCA II; hence, this water structure was not a variable in comparing with wild type the activities of mutants at residue 62. Crystal structures of wild-type and N62T HCA II showed both an inward and outward orientation of the side chain of His64; however, other mutants in this study showed predominantly inward (N62A, N62V, N62L) or predominantly outward (N62D) orientations of His64. A significant role of Asn62 in HCA II is to permit two conformations of the side chain of His64, the inward and outward, that contributes to maximal efficiency of proton transfer between the active site and solution. The site-specific mutant N62D had a mainly outward orientation of His64, yet the difference in p K a between the proton donor His64 and zinc-bound hydroxide was near zero, as in wild-type HCA II. The rate of proton transfer in catalysis by N62D HCA II was 5% that of wild type, showing that His64 mainly in the outward orientation is associated with inefficient proton transfer compared with His64 in wild type which shows both inward and outward orientations. These results emphasize the roles of the residues of the hydrophilic side of the active site cavity in maintaining efficient catalysis by carbonic anhydrase.  相似文献   

16.
The maximal velocity of catalysis of CO(2) hydration by human carbonic anhydrase II (HCA II) requires proton transfer from zinc-bound water to solution assisted by His 64. The catalytic activity of a site-specific mutant of HCA II in which His 64 is replaced with Ala (H64A HCA II) can be rescued by exogenous proton donors/acceptors, usually derivatives of imidazole and pyridine. X-ray crystallography has identified Trp 5 as a binding site of the rescue agent 4-methylimidazole (4-MI) on H64A HCA II. This binding site overlaps with the "out" position in which His 64 in wild-type HCA II points away from the zinc. Activation by 4-MI as proton donor/acceptor in catalysis was determined in the dehydration direction using (18)O exchange between CO(2) and water and in the hydration direction by stopped-flow spectrophotometry. Replacement of Trp 5 by Ala, Leu, or Phe in H64A HCA II had no significant effect on enhancement by 4-MI of maximal rate constants for proton transfer in catalysis to levels near 10(5) s(-1). This high activity for chemical rescue indicates that the binding site of 4-MI at Trp 5 in H64A HCA II appears to be a nonproductive binding site, although it is possible that a similarly effective pathway for proton transfer exists in the mutants lacking Trp 5. Moreover, the data suggest that the out position of His 64 considered alone is not active in proton transfer in HCA II. In contrast to isozyme II, the replacement of Trp 5 by Ala in HCA III abolished chemical rescue of k(cat) by imidazole but left k(cat)/K(m) for hydration unchanged. This demonstrates that Trp 5 contributes to the predominant productive binding site for imidazole, with a maximal level for the rate constant of proton transfer near 10(4) s(-1). This difference in the susceptibility of CA II and III to chemical rescue may be related to the more sterically constrained and electrostatically positive nature of the active site cavity of CA III compared with CA II. The possibility of nonproductive binding sites for exogenous proton donors offers an explanation for the unusually low value of the intrinsic kinetic barrier obtained by application of Marcus theory to chemical rescue of H64A HCA II.  相似文献   

17.
Mei H  Geren L  Miller MA  Durham B  Millett F 《Biochemistry》2002,41(12):3968-3976
The interaction of yeast iso-1-cytochrome c (yCc) with the high- and low-affinity binding sites on cytochrome c peroxidase compound I (CMPI) was studied by stopped-flow spectroscopy. When 3 microM reduced yCc(II) was mixed with 0.5 microM CMPI at 10 mM ionic strength, the Trp-191 radical cation was reduced from the high-affinity site with an apparent rate constant >3000 s(-1), followed by slow reduction of the oxyferryl heme with a rate constant of only 10 s(-1). In contrast, mixing 3 microM reduced yCc(II) with 0.5 microM preformed CMPI *yCc(III) complex led to reduction of the radical cation with a rate constant of 10 s(-1), followed by reduction of the oxyferryl heme in compound II with the same rate constant. The rate constants for reduction of the radical cation and the oxyferryl heme both increased with increasing concentrations of yCc(II) and remained equal to each other. These results are consistent with a mechanism in which both the Trp-191 radical cation and the oxyferryl heme are reduced by yCc(II) in the high-affinity binding site, and the reaction is rate-limited by product dissociation of yCc(III) from the high-affinity site with apparent rate constant k(d). Binding yCc(II) to the low-affinity site is proposed to increase the rate constant for dissociation of yCc(III) from the high-affinity site in a substrate-assisted product dissociation mechanism. The value of k(d) is <5 s(-1) for the 1:1 complex and >2000 s(-1) for the 2:1 complex at 10 mM ionic strength. The reaction of horse Cc(II) with CMPI was greatly inhibited by binding 1 equiv of yCc(III) to the high-affinity site, providing evidence that reduction of the oxyferryl heme involves electron transfer from the high-affinity binding site rather than the low-affinity site. The effects of CcP surface mutations on the dissociation rate constant indicate that the high-affinity binding site used for the reaction in solution is the same as the one identified in the yCc*CcP crystal structure.  相似文献   

18.
《MABS-AUSTIN》2013,5(6):662-669
Recombinant antibodies may be engineered to obtain improved functional properties. Functional mapping of the residues in the binding surfaces is of importance for predicting alterations needed to yield the desired properties. In this investigation, 17 single mutation mutant single-chain variable fragments (scFvs) of the anti-idiotypic antibody anti-TS1 were generated in order to functionally map amino acid residues important for the interaction with its idiotype TS1. Residues in anti-TS1 determined to be very important for the interaction were identified, Y32L, K50L, K33H, and Y52H, and they were distributed adjacent to a centrally located hydrophobic area, and contributed extensively to the interaction energy (≥2.5 kcal/mol) in the interaction. Quantitative ELISA assays, BIAcore technologies and three-dimensional surface analysis by modeling were employed to visualize the consequences of the mutations. The expression levels varied between 2 - 1,800 nM as determined by ELISA. All the 17 scFvs displayed higher dissociation rates (60 - 1,300 times) and all but two of them also faster association rates (1.3 - 56 times). The decrease in affinity was determined to be 1.6 - 12,200 times. Two of the mutants displayed almost identical affinity with the wild type anti-TS1, but with a change in both association and dissociation rates. The present investigation demonstrates that it is possible to generate a large panorama of anti-idiotypic antibodies, and single out a few that might be of potential use for future clearing and pre-targeting purposes of idiotypic-anti-idiotypic interactions.  相似文献   

19.
R E Whitwam  R S Koduri  M Natan  M Tien 《Biochemistry》1999,38(30):9608-9616
Site-directed mutagenesis was performed on Mn peroxidase (MnP) from the white-rot fungus Phanerochaete chrysosporium to investigate the role of the axial ligand hydrogen-bonding network on heme reactivity. D242 is hydrogen bonded to the proximal His of MnP; in other peroxidases, this conserved Asp, in turn, is hydrogen bonded to a Trp. In MnP and other fungal peroxidases, the Trp is replaced by a Phe (F190). Both residues are thought to have a direct influence on the electronic environment of the catalytic center. To study only the active mutants at D242 and F190, we used degenerate oligonucleotides allowing us to screen all 19 possible amino acid mutants at these positions. Two mutants at D242 passed our screen, D242E and D242S. Both mutations impaired only the functioning of compound II. The reactions of the ferric enzyme with H(2)O(2) were unaffected by the mutations, as were the reactions of compound I with reducing substrates. The D242S and D242E mutations reduced the first-order rate constant for the reaction of MnP compound II with chelated Mn(2+) from 233 s(-1) (wild type) to 154 s(-1) and 107 s(-1), respectively. Three F190 mutants passed our screen, F190V, F190L, and F190W. Similar to mutants at D242, these mutants largely affected the function of compound II. The F190V mutation increased the first-order rate constant for the reduction of compound II by chelated Mn(2+) to 320 s(-1). The F190L mutation decreased this rate to 137 s(-1). The F190W mutant was not very stable, but at pH 6.0, this mutation decreased the rate of compound II reduction by Mn(2+) from 140 s(-1) in the wild type to 36 s(-1). There was no indication that the F190W mutant was capable of forming a protein-centered Trp cation radical. All the mutations altered the midpoint potential of the Fe(3+)/Fe(2+) couple of the enzyme, as calculated from cyclic voltammagrams of the proteins. The values were shifted from -96 mV in the wild-type enzyme to -123 mV in D242S, -162 mV in D242E, -82 mV in F190L, -173 mV in F190V, and -51 mV in F190W. Collectively, these results demonstrate that D242 and F190 in MnP influence the electronic environment around the heme and that the reactions of compound II are far more sensitive to this influence than the reduction of compound I.  相似文献   

20.
S K Singh  F Guo  M R Maurizi 《Biochemistry》1999,38(45):14906-14915
The Escherichia coli ClpA and ClpP proteins form a complex, ClpAP, that catalyzes ATP-dependent degradation of proteins. Formation of stable ClpA hexamers and stable ClpAP complexes requires binding of ATP or nonhydrolyzable ATP analogues to ClpA. To understand the order of events during substrate binding, unfolding, and degradation by ClpAP, it is essential to know the oligomeric state of the enzyme during multiple catalytic cycles. Using inactive forms of ClpA or ClpP as traps for dissociated species, we measured the rates of dissociation of ClpA hexamers or ClpAP complexes. When ATP was saturating, the rate constant for dissociation of ClpA hexamers was 0.032 min(-1) (t(1/2) of 22 min) at 37 degrees C, and dissociation of ClpP from the ClpAP complexes occurred with a rate constant of 0. 092 min(-1) (t(1/2) of 7.5 min). Because the k(cat) for casein degradation is approximately 10 min(-1), these results indicate that tens of molecules of casein can be turned over by the ClpAP complex before significant dissociation occurs. Mutations in the N-terminal ATP binding site led to faster rates of ClpA and ClpAP dissociation, whereas mutations in the C-terminal ATP binding site, which cause significant decreases in ATPase activity, led to lower rates of dissociation of ClpA and ClpAP complexes. Dissociation rates for wild-type and first domain mutants of ClpA were faster at low nucleotide concentrations. The t(1/2) for dissociation of ClpAP complexes in the presence of nonhydrolyzable analogues was >/=30 min. Thus, ATP binding stabilizes the oligomeric state of ClpA, and cycles of ATP hydrolysis affect the dynamics of oligomer interaction. However, since the k(cat) for ATP hydrolysis is approximately 140 min(-1), ClpA and the ClpAP complex remain associated during hundreds of rounds of ATP hydrolysis. Our results indicate that the ClpAP complex is the functional form of the protease and as such engages in multiple rounds of interaction with substrate proteins, degradation, and release of peptide products without dissociation.  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