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.
The influence of the epsilon-subunit on the nucleotide binding affinities of the three catalytic sites of Escherichia coli F1-ATPase was investigated, using a genetically engineered Trp probe in the adenine-binding subdomain (beta-Trp-331). The interaction between epsilon and F1 was not affected by the mutation. Kd for binding of epsilon to betaY331W mutant F1 was approximately 1 nM, and epsilon inhibited ATPase activity by 90%. The only nucleotide binding affinities that showed significant differences in the epsilon-depleted and epsilon-replete forms of the enzyme were those for MgATP and MgADP at the high-affinity catalytic site 1. Kd1(MgATP) and Kd1(MgADP) were an order of magnitude higher in the absence of epsilon than in its presence. In contrast, the binding affinities for MgATP and MgADP at sites 2 and 3 were similar in the epsilon-depleted and epsilon-replete enzymes, as were the affinities at all three sites for free ATP and ADP. Comparison of MgATP binding and hydrolysis parameters showed that in the presence as well as the absence of epsilon, Km equals Kd3. Thus, in both cases, all three catalytic binding sites have to be occupied to obtain rapid (Vmax) MgATP hydrolysis rates.  相似文献   

2.
Binding of the transition state analog MgADP-fluoroaluminate to F1-ATPase   总被引:1,自引:0,他引:1  
Escherichia coli F1-ATPase from mutant betaY331W was potently inhibited by fluoroaluminate plus MgADP but not by MgADP alone. beta-Trp-331 fluorescence was used to measure MgADP binding to catalytic sites. Fluoroaluminate induced a very large increase in MgADP binding affinity at catalytic site one, a smaller increase at site two, and no effect at site three. Mutation of either of the critical catalytic site residues beta-Lys-155 or beta-Glu-181 to Gln abolished the effects of fluoroaluminate on MgADP binding. The results indicate that the MgADP-fluoroaluminate complex is a transition state analog and independently demonstrate that residues beta-Lys-155 and (particularly) beta-Glu-181 are important for generation and stabilization of the catalytic transition state. Dicyclohexylcarbodiimide-inhibited enzyme, with 1% residual steady-state ATPase, showed normal transition state formation as judged by fluoroaluminate-induced MgADP binding affinity changes, consistent with a proposed mechanism by which dicyclohexylcarbodiimide prevents a conformational interaction between catalytic sites but does not affect the catalytic step per se. The fluorescence technique should prove valuable for future transition state studies of F1-ATPase.  相似文献   

3.
J G Wise  B J Hicke  P D Boyer 《FEBS letters》1987,223(2):395-401
Under appropriate conditions tight, noncovalent binding of 2-azido-adenine nucleotides to either catalytic or noncatalytic binding sites on the E. coli F1-ATPase occurs. After removal of unbound ligands, UV-irradiation results primarily in the covalent incorporation of nucleotide moieties into the beta-subunit in both catalytic and noncatalytic site labeling experiments. Minor labeling of the alpha-subunit was also observed. After trypsin digestion and purification of the labeled peptides, microsequencing studies identified two adjacent beta-subunit tryptic peptides labeled by 2-azido-ADP or -ATP. These beta-subunit peptides were labeled on tyrosine-331 (catalytic sites) and tyrosine-354 (noncatalytic sites) in homology with the labeling patterns of the mitochondrial and chloroplast enzymes.  相似文献   

4.
We searched for new fluorescent probes of catalytic-site nucleotide binding in F(1)F(0)-ATP synthase by introducing Cys mutations at positions in or close to catalytic sites and then reacting Cys-mutant F(1) with thiol-reactive fluorescent probes. Four suitable mutant/probe combinations were identified. beta F410C labeled by 7-fluorobenz-2-oxa-1,3-diazole-4-sulfonamide (ABD-F) gave very large signal changes in response to nucleotide, allowing facile measurement of fluorescence and nucleotide-binding parameters, not only in F(1) but also in F(1)F(0). The results are consistent with the presence of three asymmetric catalytic sites of widely different affinities, with similar properties in both enzymes, and revealed a unique probe environment at the high-affinity site 1. beta Y331C F(1) labeled by ABD-F gave a large signal which monitored catalytic site polarity changes that occur along the ATP hydrolysis pathway. Two other mutant/probe combinations with significant nucleotide-responsive signals were beta Y331C labeled by 5-((((2-iodoacetyl)amino)ethyl)amino)naphthaline-1-sulfonic acid and alpha F291C labeled by 2-4'-(iodoacetamido)anilino)naphthalene-6-sulfonic acid. The signal of the latter responds differentially to nucleoside diphosphate versus triphosphate bound in catalytic sites.  相似文献   

5.
Mitochondrial F1 from the yeast Schizosaccharomyces pombe exhibits an intrinsic tryptophan fluorescence sensitive to adenine nucleotides and inorganic phosphate [Divita, G., Di Pietro, A., Deléage, G., Roux, B., & Gautheron, D.C. (1991) Biochemistry 30, 3256-3262]. The present results indicate that the intrinsic fluorescence is differentially modified by nucleotide binding to either catalytic or noncatalytic sites. Guanine or hypoxanthine nucleotides, which selectively bind to the catalytic site, produce a hyperbolic saturation monitored by fluorescence quenching at 332 nm, the maximal emission wavelength. On the contrary, adenine nucleotides, which bind to both catalytic and noncatalytic sites, exhibit a biphasic saturation. High-affinity ATP binding produces a marked quenching as opposed to the lower-affinity one. In contrast, ADP exhibits a sigmoidal saturation, with high-affinity binding producing no quenching but responsible for positive cooperativity of binding to the lower-affinity site. The catalytic-site affinity for GDP is almost 20-fold higher at pH 5.0 as compared to pH 9.0, and the high sensitivity of the method allows detection of the 10-fold lower-affinity GMP binding. In contrast, high-affinity binding of ADP, or AMP, is not pH-dependent. The selective catalytic-site saturation induces a F1 conformational change decreasing the Stern-Volmer constant for acrylamide and the tryptophan fraction accessible to iodide. ATP saturation of both catalytic and noncatalytic sites produces an additional reduction of the accessible fraction to acrylamide.  相似文献   

6.
Nadanaciva S  Weber J  Senior AE 《Biochemistry》2000,39(31):9583-9590
MgADP in combination with fluoroscandium (ScFx) is shown to form a potently inhibitory, tightly bound, noncovalent complex at the catalytic sites of F(1)-ATPase. The F(1).MgADP.ScFx complex mimics a catalytic transition state. Notably, ScFx caused large enhancement of MgADP binding affinity at both catalytic sites 1 and 2, with little effect at site 3. These results indicate that sites 1 and 2 may form a transition state conformation. A new direct optical probe of F(1)-ATPase catalytic transition state conformation is also reported, namely, substantial enhancement of fluorescence emission of residue beta-Trp-148 observed upon binding of MgADP.ScFx or MgIDP. ScFx. Using this fluorescence signal, titrations were performed with MgIDP.ScFx which demonstrated that catalytic sites 1 and 2 can both form a transition state conformation but site 3 cannot. Supporting data were obtained using MgIDP-fluoroaluminate. Current models of the MgATP hydrolysis mechanism uniformly make the assumption that only one catalytic site hydrolyzes MgATP at any one time. The fluorometal analogues demonstrate that two sites have the capability to form the transition state simultaneously.  相似文献   

7.
A structural feature shared by the metallo-beta-lactamases is a flexible loop of amino acids that extends over their active sites and that has been proposed to move during the catalytic cycle of the enzymes, clamping down on substrate. To probe the movement of this loop (residues 152-164), a site-directed mutant of metallo-beta-lactamase L1 was engineered that contained a Trp residue on the loop to serve as a fluorescent probe. It was necessary first, however, to evaluate the contribution of each native Trp residue to the fluorescence changes observed during the catalytic cycle of wild-type L1. Five site-directed mutants of L1 (W39F, W53F, W204F, W206F, and W269F) were prepared and characterized using metal analyses, CD spectroscopy, steady-state kinetics, stopped-flow fluorescence, and fluorescence titrations. All mutants retained the wild-type tertiary structure and bound Zn(II) at levels comparable with wild type and exhibited only slight (<10-fold) decreases in k(cat) values as compared with wild-type L1 for all substrates tested. Fluorescence studies revealed a single mutant, W39F, to be void of the fluorescence changes observed with wild-type L1 during substrate binding and catalysis. Using W39F as a template, a Trp residue was added to the flexile loop over the active site of L1, to generate the double mutant, W39F/D160W. This double mutant retained all the structural and kinetic characteristics of wild-type L1. Stopped-flow fluorescence and rapid-scanning UV-visible studies revealed the motion of the loop (k(obs) = 27 +/- 2 s(-1)) to be similar to the formation rate of a reaction intermediate (k(obs) = 25 +/- 2 s(-1)).  相似文献   

8.
Tryptophan (Trp) fluorescence of two phospholipases A2 (PLA2) from Naja naja atra and Naja nigricollis snake venoms was quenched by acrylamide and iodide. Trp residues in N. naja atra PLA2 were equally accessible to acrylamide and iodide. Iodide quenching studies indicate that there are two classes of Trp fluorophores in N. nigricollis CMS-9. The accessible class consists of Trp-18 and Trp-19. Removal of the N-terminal octapeptide caused a perturbation of the micro-environment of the Trp residues in the PLA2 enzymes. The presence of a substrate lowers the susceptibility of the Trp residues to iodide quenching in N. naja atra PLA2, suggesting that all three Trp residues are at the substrate binding site, but in N. nigricollis CMS-9 Trp-18 and Trp-19 are related to substrate binding.  相似文献   

9.
Kersten MV  Dunn SD  Wise JG  Vogel PD 《Biochemistry》2000,39(13):3856-3860
Electron spin resonance (ESR) spectroscopy using site-specific cysteine spin-labeling of the catalytic nucleotide binding sites of F(1)-ATPase was employed to investigate conformational changes within the nucleotide binding sites of the enzyme. Mutant Escherichia coli F(1) that had been modified at position beta-Y331C with a spin label showed almost normal catalytic activity and enabled us to study the effects of binding of different nucleotides and of the F(o) subunit b on the conformation of the catalytic binding sites. The ESR spectra of the spin-labeled, nucleotide-depleted F(1) indicate asymmetry within the sites as is expected from the structural models of the enzyme. Nucleotide binding to the enzyme clearly affects the conformation of the sites; the most pronounced feature upon nucleotide binding is the formation of catalytic site(s) in a very open conformation. Using the same beta-331 spin-labeled F(1) and a truncated form of F(o) subunit b, b(24)(-)(156), we found that binding of b(24)(-)(156) to spin-labeled F(1) significantly changes the conformation of the catalytic sites. In this paper we present data that for the first time directly show that a conformational binding change takes place upon binding of nucleotides to the nucleotide binding sites and that also show that binding of b(24)(-)(156) strongly affects the conformation of the catalytic sites, most likely by increasing the population of binding sites that are in the open conformation.  相似文献   

10.
1) Using a combination of site-directed mutagenesis and fluorescence spectroscopy we have studied the location and function of residue beta Y331 in the catalytic site of Escherichia coli F1-ATPase. The fluorescent analog lin-benzo-ADP was used as a catalytic-site probe, and was found to bind to three sites in normal F1, with Kd1 = 0.20 microM and Kd2,3 = 5.5 microM. lin-Benzo-ATP was a good substrate for hydrolysis. 2) The mutants investigated were beta Y331F, L, A and E. kcat/KM for ATP hydrolysis in purified F1 was reduced according to the series Y greater than or equal to F greater than L greater than A greater than E, with E being severely impaired; concomitant decreases in binding affinity for lin-benzo-ADP were seen. 3) Fluorescence properties of lin-benzo-ADP bound to F1 differed widely, depending on the residue present at position beta 331. Red shifts of excitation and emission spectra occurred with F and L residues, but not with Y, A, or E. There was strong quenching of fluorescence with wild-type (Y), partial quenching with A, and no quenching with F, L, or E. 4) We conclude that (a) the environment around the bound adenine moiety in the catalytic site is nonpolar, (b) residue beta 331 is part of the adenine-binding subdomain and when tyrosine is the residue, the phenolic hydroxyl makes direct interaction with the fluorophore, (c) an aromatic residue is not absolutely required at position beta 331 for catalytic function, but an increase in polarity leads to functional impairment, and (d) in terms of fluorescence response of bound lin-benzo-ADP all three catalytic sites behaved the same. 5) F1 from mutant beta Y297F bound lin-benzo-ADP with the same fluorescence and binding characteristics as normal F1, and catalytic properties were similar to normal. Therefore, there was no reason to conclude that residue beta Y297 is involved in binding the adenine moiety of ATP.  相似文献   

11.
Le NP  Omote H  Wada Y  Al-Shawi MK  Nakamoto RK  Futai M 《Biochemistry》2000,39(10):2778-2783
The three catalytic sites of the F(O)F(1) ATP synthase interact through a cooperative mechanism that is required for the promotion of catalysis. Replacement of the conserved alpha subunit Arg-376 in the Escherichia coli F(1) catalytic site with Ala or Lys resulted in turnover rates of ATP hydrolysis that were 2 x 10(3)-fold lower than that of the wild type. Mutant enzymes catalyzed hydrolysis at a single site with kinetics similar to that of the wild type; however, addition of excess ATP did not chase bound ATP, ADP, or Pi from the catalytic site, indicating that binding of ATP to the second and third sites failed to promote release of products from the first site. Direct monitoring of nucleotide binding in the alphaR376A and alphaR376K mutant F(1) by a tryptophan in place of betaTyr-331 (Weber et al. (1993) J. Biol. Chem. 268, 20126-20133) showed that the catalytic sites of the mutant enzymes, like the wild type, have different affinities and therefore, are structurally asymmetric. These results indicate that alphaArg-376, which is close to the beta- or gamma-phosphate group of bound ADP or ATP, respectively, does not make a significant contribution to the catalytic reaction, but coordination of the arginine to nucleotide filling the low-affinity sites is essential for promotion of rotational catalysis to steady-state turnover.  相似文献   

12.
The interactions between the pyrophosphate (PPi) binding sites and the nucleotide binding sites on mitochondrial F1-ATPase have been investigated, using F1 preparations containing different numbers of catalytic and noncatalytic nucleotide-binding sites occupied by ligands. In all cases, the total number of moles of bound nucleotides and PPi per mole of F1 was less than or equal to six. F1 preparations containing either three or two filled noncatalytic sites and no filled catalytic sites (referred as F1[3,0] and F1[2,0]) were found to bind 3 mol of PPi/mol of F1. Tight binding of ADP-fluoroberyllate complexes to two of the catalytic sites of F1 converted the three heterogeneous PPi-binding sites into three homogeneous binding sites, each exhibiting the same affinity for PPi. The addition of PPi at saturating concentrations to F1 containing GDP bound to two catalytic sites (F1[2,2]) resulted in the release of 1 mol of GDP. Furthermore, the addition of PPi to F1 filled with ADP-fluoroberyllate at the catalytic sites resulted in the release of 1 mol of tightly bound ADP/mol of F1. Taken together, these results indicate that PPi binds to specific sites that interact with both the catalytic and the noncatalytic nucleotide-binding sites of F1.  相似文献   

13.
The subtilisin-like serine proteinases, VPR, from a psychrotrophic Vibrio species and aqualysin I (AQUI) from the thermophile Thermus aquaticus, are structural homologues, but differ significantly with respect to stability and catalytic properties. It has been postulated that the higher catalytic activity of cold adapted enzymes when compared to homologues from thermophiles, reflects their higher molecular flexibility. To assess a potential difference in molecular flexibility between the two homologous proteinases, we have measured their Trp fluorescence quenching by acrylamide at different temperatures. We also investigated protein dynamics of VPR and AQUI at an atomic level by molecular dynamics simulations. VPR contains four Trp residues, three of which are at corresponding sites in the structure of AQUI. To aid in the comparison, a Tyr at the fourth corresponding site in AQUI was mutated to Trp (Y191W). A lower quenching effect of acrylamide on the intrinsic fluorescence of the thermophilic AQUI_Y191W was observed at all temperatures measured (10–55 °C), suggesting that it possesses a more rigid structure than VPR. The MD analysis (Cα rmsf profiles) showed that even though VPR and AQUI have similar flexibility profiles, the cold adapted VPR displays higher flexibility in most regions of the protein structure. Some of these regions contain or are in proximity to some of the Trp residues (Trp6, Trp114 and Trp208) in the proteins. Thus, we observe an overall agreement between the fluorescence quenching data and the flexibility profiles obtained from the MD simulations to different flexibilities of specific regions in the proteins.  相似文献   

14.
A spin-labeled photoaffinity ATP analog, 2-N3-2',3'-SL-ATP (2-N3-SL-ATP) was specifically loaded at catalytic (exchangeable) or noncatalytic (nonexchangeable) nucleotide-binding sites on nucleotide-depleted mitochondrial F1-ATPase. Photolysis of the enzyme complexes resulted in the specific modification of beta-Tyr-345 when the catalytic sites were occupied and beta-Tyr-368 when noncatalytic sites were filled. These are the same amino acid assignments that were made previously using 2-N3ATP. The results demonstrate that the attachment of a spin label moiety to the ribose ring does not prevent proper binding of the analog at both types of nucleotide sites on F1-ATPase and suggest that the probe can be used for investigations of the nucleotide-binding sites using ESR spectroscopy. Enzyme that is in complex with the 2-N3-SL-ATP exhibits an ESR spectrum that is typical for highly immobilized nitroxyl radicals both in the dark or after photolysis. Additional peaks in the high- and low-field regions arise due to dipolar spin interactions most likely involving a pair of catalytic and noncatalytic sites. The two sites are calculated to be approximately 15 A apart. This distance, obtained through ESR spectroscopy, combined with the finding that the 2 labeled amino acids are only 23 residues apart from each other, further supports an adenylate kinase-like arrangement of nucleotide binding sites on F1-ATPase where catalytic and noncatalytic sites are in close proximity (Vogel, P. D., and Cross, R. L. (1991) J. Biol. Chem. 266, 6101-6105).  相似文献   

15.
J Pagan  A E Senior 《FEBS letters》1990,273(1-2):147-149
It is shown that ATP dissociates very slowly (koff less than 6.4 x 10(5) s-1, t1/2 greater than 3 h) from the three noncatalytic sites of E. coli F1-ATPase and that ADP dissociates from these three sites in a homogeneous fashion with koff = 1.5 x 10(-4) s-1 (t1/2 = 1.35 h). Mutagenesis of alpha-subunit residues R171 and Q172 in the 'glycine-rich loop' (Homology A) consensus region of the noncatalytic sites was carried out to test the hypothesis that unusually bulky residues at these positions are responsible wholly or partly for the observed tight binding of adenine nucleotides. The mutations alpha Q172G or alpha R171S,Q172G had no effects on ATP or ADP binding to or rates of dissociation from F1 noncatalytic sites. KdATP and KdADP of isolated alpha-subunit were weakened by approximately 1 order of magnitude in both mutants. The results suggest that neither residue alpha R171 nor alpha Q172 interacts directly with bound nucleotide, and show that the presence of bulky residues per se in the glycine-rich loop region of F1-alpha-subunit is not responsible for tight binding in the noncatalytic sites.  相似文献   

16.
Tryptophan fluorescence was used to study GK (glucokinase), an enzyme that plays a prominent role in glucose homoeostasis which, when inactivated or activated by mutations, causes diabetes mellitus or hypoglycaemia in humans. GK has three tryptophan residues, and binding of D-glucose increases their fluorescence. To assess the contribution of individual tryptophan residues to this effect, we generated GST-GK [GK conjugated to GST (glutathione transferase)] and also pure GK with one, two or three of the tryptophan residues of GK replaced with other amino acids (i.e. W99C, W99R, W167A, W167F, W257F, W99R/W167F, W99R/W257F, W167F/W257F and W99R/W167F/W257F). Enzyme kinetics, binding constants for glucose and several other sugars and fluorescence quantum yields (varphi) were determined and compared with those of wild-type GK retaining its three tryptophan residues. Replacement of all three tryptophan residues resulted in an enzyme that retained all characteristic features of GK, thereby demonstrating the unique usefulness of tryptophan fluorescence as an indicator of GK conformation. Curves of glucose binding to wild-type and mutant GK or GST-GK were hyperbolic, whereas catalysis of wild-type and most mutants exhibited co-operativity with D-glucose. Binding studies showed the following order of affinities for the enzyme variants: N-acetyl-D-glucosamine>D-glucose>D-mannose>D-mannoheptulose>2-deoxy-D-glucose>L-glucose. GK activators increased sugar binding of most enzymes, but not of the mutants Y214A/V452A and C252Y. Contributions to the fluorescence increase from Trp(99) and Trp(167) were large compared with that from Trp(257) and are probably based on distinct mechanisms. The average quantum efficiency of tryptophan fluorescence in the basal and glucose-bound state was modified by activating (Y214A/V452A) or inactivating (C213R and C252Y) mutations and was interpreted as a manifestation of distinct conformational states.  相似文献   

17.
In order to get insight into the origin of apparent negative cooperativity observed for F(1)-ATPase, we compared ATPase activity and ATPMg binding of mutant subcomplexes of thermophilic F(1)-ATPase, alpha((W463F)3)beta((Y341W)3)gamma and alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma. For alpha((W463F)3)beta((Y341W)3)gamma, apparent K(m)'s of ATPase kinetics (4.0 and 233 microM) did not agree with apparent K(m)'s deduced from fluorescence quenching of the introduced tryptophan residue (on the order of nM, 0.016 and 13 microM). On the other hand, in case of alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma, which lacks noncatalytic nucleotide binding sites, the apparent K(m) of ATPase activity (10 microM) roughly agreed with the highest K(m) of fluorescence measurements (27 microM). The results indicate that in case of alpha((W463F)3)beta((Y341W)3)gamma, the activating effect of ATP binding to noncatalytic sites dominates overall ATPase kinetics and the highest apparent K(m) of ATPase activity does not represent the ATP binding to a catalytic site. In case of alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma, the K(m) of ATPase activity reflects the ATP binding to a catalytic site due to the lack of noncatalytic sites. The Eadie-Hofstee plot of ATPase reaction by alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma was rather linear compared with that of alpha((W463F)3)beta((Y341W)3)gamma, if not perfectly straight, indicating that the apparent negative cooperativity observed for wild-type F(1)-ATPase is due to the ATP binding to catalytic sites and noncatalytic sites. Thus, the frequently observed K(m)'s of 100-300 microM and 1-30 microM range for wild-type F(1)-ATPase correspond to ATP binding to a noncatalytic site and catalytic site, respectively.  相似文献   

18.
Previously we have shown that beef heart mitochondrial F1 contains a total of six adenine nucleotide binding sites. Three "catalytic" sites exchange bound ligand rapidly during hydrolysis of MgATP, whereas three "noncatalytic" sites do not. The noncatalytic sites behave asymmetrically in that a single site releases bound ligand upon precipitation of F1 with ammonium sulfate. In the present study, we find this same site to be the only noncatalytic site that undergoes rapid exchange of bound ligand when F1 is incubated in the presence of EDTA at pH 8.0. Following 1000 catalytic turnovers/F1, the site retains the unique capacity for EDTA-induced exchange, indicating that the asymmetric determinants are permanent and that the three noncatalytic sites on soluble F1 do not pass through equivalent states during catalysis. Measurements of the rate of ligand binding at the unique noncatalytic site show that uncomplexed nucleotide binds preferentially. At pH 7.5, in the presence of Mg2+, the rate constant for ADP binding is 9 X 10(3) M-1 s-1 and for dissociation is 4 X 10(-4) s-1 to give a Kd = 50 nM. The rate of dissociation is 10 times faster in the presence of EDTA or during MgATP hydrolysis, and it increases rapidly at pH below 7. EDTA-induced exchange is inhibited by Mg2+, Mn2+, Co2+, and Zn2+ but not by Ca2+ and is unaffected by dicyclohexylcarbodiimide modification. The unique noncatalytic site binds 2-azido-ADP. Photolysis results in the labeling of the beta subunit. Photolabeling of a single high-affinity catalytic site under conditions for uni-site catalysis also results in the labeling of beta, but a different pattern of labeled peptides is obtained in proteolytic digests. The results demonstrate the presence of two different nucleotide binding domains on the beta subunit of mitochondrial F1.  相似文献   

19.
The access of three proteases to their sites of cleavage was used as a measure of regulatory interactions in the plasma membrane Ca2+ pump isoform 4b (PMCA4b). When the proteases could not cut at their sites in the C-terminal regulatory region, the interaction was judged to be tight. This was the case in the absence of Ca2+, when chymotrypsin and caspase cut PMCA only very slowly. Ca2+ accelerated the fragmentation, but the digestion remained incomplete. In the presence of Ca2+ plus calmodulin, the digestion became nearly complete in all cases, indicating a more flexible conformation of the carboxyl terminus in the fully activated state. The acceleration of proteolysis by Ca2+ or Ca2+ plus calmodulin occurred equally at the caspase site upstream of the calmodulin-binding domain and the chymotrypsin and calpain sites downstream of that domain. Replacing Trp1093 (a key residue within the calmodulin-binding domain) with alanine had a much more specific effect, because it exposed only proteolytic sites within the calmodulin-binding domain that had previously been shielded in the native protein. At these sites, both calpain and chymotrypsin cut the Trp1093 --> Ala mutant in the absence of calmodulin. These data indicate that, in the auto-inhibited conformation, the calmodulin-binding/auto-inhibitory sequence and the regions both upstream and downstream are in close contact with the catalytic core. Trp1093 plays an essential role not only in stabilizing the Ca2+-calmodulin/calmodulin-binding domain complex but also in the formation or stability of the inhibitory conformation of that domain when it interacts with the catalytic core of PMCA4b.  相似文献   

20.
The availability of a high-resolution structure of the Thermobifida fusca endocellulase Cel6A catalytic domain makes this enzyme ideal for structure-based efforts to engineer cellulases with high activity on native cellulose. In order to determine the role of conserved, noncatalytic residues in cellulose hydrolysis, 14 mutations of six conserved residues in or near the Cel6A active-site cleft were studied for their effects on catalytic activity, substrate specificity, processivity and ligand-binding affinity. Eleven mutations were generated by site-directed mutagenesis using PCR, while three were from previous studies. All the CD spectra of the mutant enzymes were indistinguishable from that of Cel6A indicating that the mutations did not dramatically change protein conformation. Seven mutations in four residues (H159, R237, K259 and E263) increased activity on carboxymethyl cellulose (CM-cellulose), with K259H (in glucosyl subsite -2) creating the highest activity (370%). Interestingly, the other mutations in these residues reduced CM-cellulose activity. Only the K259H enzyme retained more activity on acid-swollen cellulose than on filter paper, suggesting that this mutation affected the rate-limiting step in crystalline cellulose hydrolysis. All the mutations lowered activity on cellotriose and cellotetraose, but two mutations, both in subsite +1 (H159S and N190A), had higher kcat/Km values (6.6-fold and 5.0-fold, respectively) than Cel6A on 2,4-dinitrophenyl-beta-D-cellobioside. Measurement of enzyme : ligand dissociation constants for three methylumbelliferyl oligosaccharides and cellotriose showed that all mutant enzymes bound these ligands either to the same extent as or more weakly than Cel6A. These results show that conserved noncatalytic residues can profoundly affect Cel6A activity and specificity.  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