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 complex of ribonuclease A (RNase A) with uridine vanadate (U-V), a transition-state analogue, has been studied with 51V and proton NMR spectroscopy in solution and by neutron diffraction in the crystalline state. Upon the addition of aliquots of U-V at pH 6.6, the C epsilon-H resonances of the two active-site histidine residues 119 and 12 decrease in intensity while four new resonances appear. Above pH 8 and below pH 5, these four resonances decrease in intensity as the complex dissociates. These four resonances are assigned to His-119 and His-12 in protonated and unprotonated forms in the RNase-U-V complex. These resonances do not titrate or change in relative area in the pH range 5-8, indicating a slow protonation process, and the extent of protonation remains constant with ca. 58% of His-12 and ca. 26% of His-119 being protonated. The results of diffraction studies show that both His-12 and His-119 occupy well-defined positions in the RNase-U-V complex and that both are protonated. However, while the classic interpretation of the mechanism of action of RNase based on the proposal of Findlay et al. [Findlay, D., Herries, D. G., Mathias, A. P., Rabin, B. R., & Ross, C. A. (1962) Biochem. J. 85, 152-153] requires both His-12 and His-119 to be in axial positions relative to the pentacoordinate transition state, in the diffraction structure His-12 is found to be in an equatorial position, while Lys-41 is close to an axial position. Hydrogen exchange data show that the mobility and accessibility of amides in the RNase-U-V complex do not significantly differ from what was observed in the native enzyme. The results of both proton NMR in solution and neutron diffraction in the crystal are compared and interpreted in terms of the mechanism of action of RNase.  相似文献   

2.
Tables of protonic activity (paH) of a number of buffers, determined in mixed solvents and at subzero temperatures, are reported for the following media: water-1,2-propanediol, water-glycerol, and water-dimethylsulfoxide (50:50, in volume). These data with those previously reported allowed us to study enzymic reactions under these conditions. The paH dependence of the tryptic hydrolysis of benzoyl-L-arginine ethyl ester has been studied in the presence of organic solvents (methanol, ethylene glycol, 1,2-propanediol, glycerol, and dimethylsulfoxide, all 50% by volume) between 20 and -20 degrees. The results have allowed us to show the validity of our paH scales in mixed solvents. The paH profiles obtained under these conditions are similar to those observed in pure water at 20 degrees. They are shifted nevertheless by both solvent and temperature. Such shifts are interpreted in terms of the effects of solvents and temperature on pKES on the basis of the conclusions drawn from a study of the effect of these variables on small dissociable molecules. The results obtained under these conditions of solvents and temperature are consistent with the presence at the active site of the enzyme of a histidine residue, and thus provide, concerning the solvent effect, a direct verification of the method of Findlay et al. (Findlay, D., Mathias, A. P., and Rabin, B. R. (1962) Biochem. J. 85, 139-144). On the other hand, the large temperature interval provided by the low temperature procedure, allows us to vary significantly the pK of ionizable groups of the enzymes and thus makes possible their identification, on the basis of their enthalpy of ionization.  相似文献   

3.
Rab3A is a small, Ras-like GTPase expressed in neuroendocrine cells, in which it is associated with secretory vesicle membranes and regulates exocytosis. Using the yeast two-hybrid system, we have identified a rat brain cDNA encoding a novel 50-kDa protein, which we have named Rabin3, that interacts with Rab3A and Rab3D but not with other small GTPases (Rab3C, Rab2, Ran, or Ras). Several independent point mutations in the effector domain of Rab3A (F51L, V55E, and G56D) which do not alter nucleotide binding by the GTPase abolish the interaction with Rabin3, while another mutation (V52A) appears to increase the interaction. These results demonstrate that the interaction is highly specific. However, a glutathione S-transferase-Rabin3 fusion protein associates only weakly in vitro with recombinant Rab3A and possesses no detectable GTPase-activating protein or nucleotide exchange activity, and Rabin3 overexpressed in adrenal chromaffin cells has no observable effect on secretion. The protein possess a sequence characteristic of coiled-coil domains and a second small region with sequence similarity to a Saccharomyces cerevisiae protein, Sec2p, Sec2p is essential for constitutive secretion in yeast cells and interacts with Sec4p, a close relative of the Rab3A GTPase. Rabin3 mRNA and protein are widely expressed but are particularly abundant in testes.  相似文献   

4.
The mechanisms mediating polarized delivery of vesicles to cell surface domains are poorly understood in animal cells. We have previously shown that expression of Rab8 promotes the formation of new cell surface domains through reorganization of actin and microtubules. To unravel the function of Rab8, we used the yeast two-hybrid system to search for potential Rab8-specific activators. We identified a coil-coiled protein (Rabin8), homologous to the rat Rabin3 that stimulated nucleotide exchange on Rab8 but not on Rab3A and Rab5. Furthermore, we show that rat Rabin3 has exchange activity on Rab8 but not on Rab3A, supporting the view that rat Rabin3 is the rat equivalent of human Rabin8. Rabin8 localized to the cortical actin and expression of Rabin8 resulted in remodeling of actin and the formation of polarized cell surface domains. Activation of PKC by phorbol esters enhanced translocation of both Rabin8 and Rab8-specific vesicles to the outer edge of lamellipodial structures. Moreover, coexpression of Rabin8 with dominant negative Rab8 (T22N) redistributes Rabin8 from cortical actin to Rab8-specific vesicles and promotes their polarized transport to cell protrusions. The C-terminal region of Rabin8 plays an essential role in this transport. We propose that Rabin8 is a Rab8-specific activator that is connected to processes that mediate polarized membrane traffic to dynamic cell surface structures.  相似文献   

5.
beta-Ketoacyl-[acyl carrier protein (ACP)] synthase forms new carbon-carbon bonds in three steps: transfer of an acyl primer from ACP to the enzyme, decarboxylation of the elongating substrate and its condensation with the acyl primer substrate. Six residues of Escherichia coli beta-ketoacyl-ACP synthase I (KAS I) implicated in these reactions were subjected to site-directed mutagenesis. Analyses of the abilities of C163A, C163S, H298A, D306A, E309A, K328A, and H333A to carry out the three reactions lead to the following conclusions. The active site Cys-163 is not required for decarboxylation, whereas His-298 and His-333 are indispensable. Neither of the histidines is essential for increasing the nucleophilicity of Cys-163 to enable transfer of the acyl primer substrate. Maintenance of the structural integrity of the active site by Asp-306 and Glu-309 is required for decarboxylation but not for transfer. One function of Lys-328 occurs very early in catalysis, potentially before transfer. These results in conjunction with structural analyses of substrate complexes have led to a model for KAS I catalysis [Olsen, J. G., Kadziola, A., von Wettstein-Knowles, P., Siggaard-Andersen, M., and Larsen, S. (2001) Structure 9, 233-243]. Another facet of catalysis revealed by the mutant analyses is that the acyl primer transfer activity of beta-ketoacyl-ACP synthase I is inhibited by free ACP at physiological concentrations. Differences in the inhibitory response by individual mutant proteins indicate that interaction of free ACP with Cys-163, Asp-306, Glu-309, Lys-328, and His-333 might form a sensitive regulatory mechanism for the transfer of acyl primers.  相似文献   

6.
Han J  Burke JM 《Biochemistry》2005,44(21):7864-7870
We have used nucleobase substitution and kinetic analysis to test the hypothesis that hammerhead catalysis occurs by a general acid-base mechanism, in which nucleobases are directly involved in deprotonation of the attacking 2'-hydroxyl group and protonation of the 5'-oxygen that serves as the leaving group in the cleavage reaction. We demonstrate that simultaneous substitution of two important nucleobases, G8 and G12, with 2,6-diaminopurine shifts the pH optimum of the cleavage reaction from greater than 9.5 to approximately 6.8 in two different hammerhead constructs. Controls involving substitution with other nucleobases and combinations of nucleobases at G5, G8, and/or G12 do not show this behavior. The observed changes in the pH-rate behavior are consistent with a mechanism in which N1 protonation-deprotonation events of guanine or 2,6-diaminopurine at positions 8 and 12 are essential for catalysis. Further support for the participation of G8 and G12 comes from photochemical cross-linking experiments, which show that G8 and G12 can stack upon the two substrate nucleobases at the reactive linkage, G(or U)1.1 and C17 (Heckman, J. E., Lambert, D., and Burke, J. M. (2005) Photocrosslinking detects a compact active structure of the hammerhead ribozyme, Biochemistry 44, 4148-4156). Together, these results support a model in which the hammerhead undergoes a transient conformational change into a catalytically active structure, in which stacking of G8 and G12 upon the nucleobases spanning the cleavage site provides an appropriate architecture for general acid-base catalysis. The hammerhead and hairpin ribozymes may share similarities in the organization of their active sites and their catalytic mechanism.  相似文献   

7.
We have explored the substrate protonation mechanism of Escherichia coli dihydrofolate reductase by changing the location of the proton donor. A double mutant was constructed in which the proton donor of the wild-type enzyme, aspartic acid-27, has been changed to serine and simultaneously an alternative proton donor, glutamic acid, has replaced threonine at position 113. The active site of the resulting variant enzyme molecule should therefore somewhat resemble that proposed for the R67 plasmid-encoded dihydrofolate reductase [Matthews, D. A., Smith, S. L., Baccanari, D. P., Burchall, J. J., Oatley, S. J., & Kraut, J. (1986) Biochemistry 25, 4194]. At pH 7, the double-mutant enzyme has a 3-fold greater kcat and an unchanged Km(dihydrofolate) as compared with the single-mutant Asp-27----Ser enzyme described previously [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science (Washington, D.C.) 231, 1123]. Additionally, its activity vs pH profiles together with observed deuterium isotope effects, suggest that catalysis depends on an acidic group with a pKa of 8. It is concluded that the dihydropteridine ring of a bound substrate molecule can indeed be protonated by a glutamic acid side chain at position 113 (instead of an aspartic acid side chain at position 27), but with greatly decreased efficiency: at pH 7, the double mutant still has a 25-fold lower kcat (1.2 s-1) and a 2900-fold lower kcat/km(dihydrofolate) (8.6 X 10(3) s-1 M-1) than the wild-type enzyme.  相似文献   

8.
Aeromonas hydrophila releases a protein which is a member of the lipase superfamily, similar in reaction mechanism to the important mammalian plasma enzyme lecithin-cholesterol acyltransferase. We have used chemical modification and site-directed mutagenesis of the protein to identify amino acids which may be involved in catalysis. The enzyme was unaffected by phenylmethylsulfonyl fluoride, but it was almost completely inhibited by another serine-reactive compound, diethyl p-nitrophenyl phosphate. A serine selectively modified by this reagent was identified by sequencing the amino-terminal region of the protein. It was located at position 16, in the short consensus sequence shared by the enzyme with other lipases. When this serine was changed to asparagine the product was an enzymatically inert protein which nevertheless retained the surface activity of the wild-type enzyme, suggesting its ability to bind to substrate was not changed. Diethylpyrocarbonate treatment drastically reduced the rate of acyl transfer by the native enzyme, but this did not appear to be due to modification of an essential histidine, since inhibition was not reversed by addition of hydroxylamine. We have shown that only two of the histidines in the enzyme can be involved in catalysis (Hilton, S., McCubbin, W. D., Kay, C.M., and Buckley, J. T. (1990) Biochemistry, 29, 9072-9078). Replacing both of these with asparagines had little or no effect on enzyme activity. These results indicate that, in apparent contrast to other lipases, histidine does not participate in the reaction catalyzed by the microbial enzyme. Since catalysis was not inhibited by sulfhydryl reagents, we conclude that a free cysteine is also not required for activity. This may distinguish the microbial enzyme from the mammalian acyltransferase.  相似文献   

9.
A F Monzingo  B W Matthews 《Biochemistry》1984,23(24):5724-5729
The mode of binding of the specific thermolysin inhibitor N-(1-carboxy-3-phenylpropyl)-L-leucyl-L-tryptophan (KI approximately 5 X 10(-8) M) [Maycock, A. L., DeSousa, D. M., Payne, L. G., ten Broeke, J., Wu, M. T., & Patchett, A. A. (1981) Biochem. Biophys. Res. Commun. 102, 963-969] has been determined by X-ray crystallography and refined to an R value of 17.1% at 1.9-A resolution. The inhibitor binds to thermolysin with both oxygens of the N-carboxymethyl group liganded to the zinc to give overall pentacoordination of the metal. The bidentate ligation of the inhibitor differs from the monodentate binding seen previously for carboxylate-zinc interactions in thermolysin and is closer to the bidentate geometry observed for the binding of hydroxamates [Holmes, M. A., & Matthews, B. W. (1981) Biochemistry 20, 6912-6920]. The geometry of the inhibitor and its interactions with the protein have a number of elements in common with the presumed transition state formed during peptide hydrolysis. The observed zinc ligation supports the previous suggestion that a pentacoordinate intermediate participates in the mechanism of catalysis. However, the alpha-amino nitrogen of the inhibitor is close to Glu-143, suggesting that this residue might accept a proton from an attacking water molecule (as proposed before) and subsequently donate this proton to the leaving nitrogen. By analogy with thermolysin, it is proposed that a related mechanism should be considered for peptide cleavage by carboxypeptidase A.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Choi G  Ha NC  Kim SW  Kim DH  Park S  Oh BH  Choi KY 《Biochemistry》2000,39(5):903-909
Delta 5-3-ketosteroid isomerase (KSI) catalyzes the allylic isomerization of Delta 5-3-ketosteroids at a rate approaching the diffusion limit by an intramolecular transfer of a proton. Despite the extensive studies on the catalytic mechanism, it still remains controversial whether the catalytic residue Asp-99 donates a hydrogen bond to the steroid or to Tyr-14. To clarify the role of Asp-99 in the catalysis, two single mutants of D99E and D99L and three double mutants of Y14F/D99E, Y14F/D99N, and Y14F/D99L have been prepared by site-directed mutagenesis. The D99E mutant whose side chain at position 99 is longer by an additional methylene group exhibits nearly the same kcat as the wild-type while the D99L mutant exhibits ca. 125-fold lower kcat than that of the wild-type. The mutations made at positions 14 and 99 exert synergistic or partially additive effect on kcat in the double mutants, which is inconsistent with the mechanism based on the hydrogen-bonded catalytic dyad, Asp-99 COOH...Tyr-14 OH...C3-O of the steroid. The crystal structure of D99E/D38N complexed with equilenin, an intermediate analogue, at 1.9 A resolution reveals that the distance between Tyr-14 O eta and Glu-99 O epsilon is ca. 4.2 A, which is beyond the range for a hydrogen bond, and that the distance between Glu-99 O epsilon and C3-O of the steroid is maintained to be ca. 2.4 A, short enough for a hydrogen bond to be formed. Taken together, these results strongly support the idea that Asp-99 contributes to the catalysis by donating a hydrogen bond directly to the intermediate.  相似文献   

11.
Arginine kinase is a member of the phosphagen kinase family that includes creatine kinase and likely shares a common reaction mechanism in catalyzing the buffering of cellular ATP energy levels. Abstraction of a proton from the substrate guanidinium by a catalytic base has long been thought to be an early mechanistic step. The structure of arginine kinase as a transition state analog complex (Zhou, G., Somasundaram, T., Blanc, E., Parthasarathy, G., Ellington, W. R., and Chapman, M. S. (1998) Proc. Natl. Acad. Sci. U. S. A. 95, 8449-8454) showed that Glu-225 and Glu-314 were the only potential catalytic residues contacting the phosphorylated nitrogen. In the present study, these residues were changed to Asp, Gln, and Val or Ala in several single and multisite mutant enzymes. These mutations had little impact on the substrate binding constants. The effect upon activity varied with reductions in kcat between 3000-fold and less than 2-fold. The retention of significant activity in some mutants contrasts with published studies of homologues and suggests that acid-base catalysis by these residues may enhance the rate but is not absolutely essential. Crystal structures of mutant enzymes E314D at 1.9 A and E225Q at 2.8 A resolution showed that the precise alignment of substrates is subtly distorted. Thus, pre-ordering of substrates might be just as important as acid-base chemistry, electrostatics, or other potential effects in the modest impact of these residues upon catalysis.  相似文献   

12.
13.
The crystal structures of rhizopuspepsin complexed with two oligopeptide inhibitors have been determined. CP-69,799, an azahomostatine dipeptide isostere, had previously been associated with a displacement of the C-terminal subdomain of endothiapepsin [Sali, A., Veerapandian, B., Cooper, J. B., Foundling, S. I., Hoover, D. J., & Blundell, T. L. (1989) EMBO J. 8, 2179-2188]. Here, we report the measurement of two data sets, one from crystals soaked in the inhibitor and the other from protein crystallized in the presence of excess inhibitor. In neither case is there any significant movement of the C-terminal subdomain of the rhizopuspepsin. The data suggest that the energy associated with any conformational change is small and is overcome by the crystal packing forces. The second inhibitor, a hydrated difluorostatone, was examined in a search for transition-state analogs that could cast further light on the mechanism of action [Suguna, K., Padlan, E. A., Smith, C. W., Carlson, W. D., & Davies, D. R. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 7009-7013]. The gem-diol provides a set of contact distances with the enzyme that mimic the interactions with the tetrahedral intermediate of the substrate during catalysis. These data provide support for the suggestion that the polarization of the keto group of the peptide substrate is enhanced by a hydrogen bond from the OD1 of Asp 35 (Suguna et al., 1987).  相似文献   

14.
Murine antibody 1D4 selectively catalyzes a highly disfavored beta-elimination reaction. Crystal structures of unliganded 1D4 and 1D4 in complex with a transition-state analog (TSA) have elucidated a possible general base mode of catalysis. The structures of the unliganded and liganded Fabs were determined to 1.80 and 1.85 A resolution, respectively. The structure of the complex reveals a binding pocket with high shape complementarity to the TSA, which is recruited to coerce the substrate into the sterically demanding, eclipsed conformation that is required for catalysis. A histidine residue and two water molecules are likely involved in the catalysis. The structure supports either a concerted E2 or stepwise E1cB-like mechanism for elimination. Finally, the liganded 1D4 structure shows minor conformational rearrangements in CDR H2, indicative of induced-fit binding of the hapten. 1D4 has pushed the boundaries of antibody-mediated catalysis into the realm of disfavored reactions and, hence, represents an important milestone in the development of this technology.  相似文献   

15.
Despite their clinical importance, the mechanism of action of the class C beta-lactamases is poorly understood. In contrast to the class A and class D beta-lactamases, which contain a glutamate residue and a carbamylated lysine in their respective active sites that are thought to serve as general base catalysts for beta-lactam hydrolysis, the mechanism of activation of the serine and water nucleophiles in the class C enzymes is unclear. To probe for residues involved in catalysis, the class C beta-lactamase from Enterobacter cloacae P99 was studied by combinatorial scanning mutagenesis at 122 positions in and around the active site. Over 1000 P99 variants were screened for activity in a high-throughput in vivo antibiotic resistance assay and sequenced by 96-capillary electrophoresis to identify residues that are important for catalysis. P99 mutants showing reduced capability to convey antibiotic resistance were purified and characterized in vitro. The screen identified an active-site hydrogen-bonding network that is key to catalysis. A second cluster of residues was identified that likely plays a structural role in the enzyme. Otherwise, residues not directly contacting the substrate showed tolerance to substitution. The study lends support to the notion that the class C beta-lactamases do not have a single residue that acts as the catalytic general base. Rather, catalysis is affected by a hydrogen-bonding network in the active site, suggesting a possible charge relay system.  相似文献   

16.
Acetate kinase catalyzes transfer of the gamma-phosphate of ATP to acetate. The only crystal structure reported for acetate kinase is the homodimeric enzyme from Methanosarcina thermophila containing ADP and sulfate in the active site (Buss, K. A., Cooper, D. C., Ingram-Smith, C., Ferry, J. G., Sanders, D. A., and Hasson, M. S. (2001) J. Bacteriol. 193, 680-686). Here we report two new crystal structure of the M. thermophila enzyme in the presence of substrate and transition state analogs. The enzyme co-crystallized with the ATP analog adenosine 5'-[gamma-thio]triphosphate contained AMP adjacent to thiopyrophosphate in the active site cleft of monomer B. The enzyme co-crystallized with ADP, acetate, Al(3+), and F(-) contained a linear array of ADP-AlF(3)-acetate in the active site cleft of monomer B. Together, the structures clarify the substrate binding sites and support a direct in-line transfer mechanism in which AlF(3) mimics the meta-phosphate transition state. Monomers A of both structures contained ADP and sulfate, and the active site clefts were closed less than in monomers B, suggesting that domain movement contributes to catalysis. The finding that His(180) was in close proximity to AlF(3) is consistent with a role for stabilization of the meta-phosphate that is in agreement with a previous report indicating that this residue is essential for catalysis. Residue Arg(241) was also found adjacent to AlF(3), consistent with a role for stabilization of the transition state. Kinetic analyses of Arg(241) and Arg(91) replacement variants indicated that these residues are essential for catalysis and also indicated a role in binding acetate.  相似文献   

17.
J H Shim  S J Benkovic 《Biochemistry》1999,38(31):10024-10031
Site-directed mutagenesis followed by studies of the pH dependence of the kinetic parameters of the mutants has been used to probe the role of the active site residues and loops in catalysis by glycinamide ribonucleotide transformylase (EC 2.1.2.2). The analysis of the mutants of the strictly conserved active site residues, His108 and Asp144, revealed that His108 acts in a salt bridge with Asp144 as a general acid catalyst with a pK(a) value of 9.7. Asp144 also plays a key role in the preparation of the active site geometry for catalysis. The rate-limiting step in the pH range of 6-10 appears to be the catalytic steps involving tetrahedral intermediates, supported by the observation of a pL (L being H or D)-independent solvent deuterium isotope effect of 2. The ionization of the amino group of glycinamide ribonucleotide both as a free and as a bound form dominates the kinetic behavior at low pH. The analysis of a mutation, H121Q, within the loop spanning amino acids 111-131 suggests the closure of the loop is involved in the binding of the substrate. The kinetic behavior parallels pH effects revealed by a series of X-ray crystallographic structures of the apoenzyme and inhibitor-bound enzyme [Su, Y., Yamashita, M. M., Greasley, S. E. , Mullen, C. A., Shim, J. H., Jennings, P. A., Benkovic, S. J., and Wilson, I. A. (1998) J. Mol. Biol. 281, 485-499], permitting a more exact formulation of the probable catalytic mechanism.  相似文献   

18.
P Carter  L Abrahmsén  J A Wells 《Biochemistry》1991,30(25):6142-6148
A mutant of the serine protease, subtilisin BPN', in which the catalytic His64 is replaced by Ala (H64A), is very specific for substrates containing a histidine, presumably by the substrate-bound histidine assisting in catalysis [Carter, P., & Wells, J.A. (1987) Science (Washington, D.C.) 237, 394-399]. Here we probe the catalytic mechanism of H64A subtilisin for cleaving His and non-His substrates. We show that the ratio of aminolysis to hydrolysis is the same for ester and amide substrates as catalyzed by the H64A subtilisin. This is consistent with formation of a common acyl-enzyme intermediate for H64A subtilisin, analogous to the mechanism of the wild-type enzyme. However, the catalytic efficiencies (kcat/KM) for amidase and esterase activities with His-containing substrates are reduced by 5000-fold and 14-fold, respectively, relative to wild-type subtilisin BPN, suggesting that acylation is more compromised than deacylation in the H64A mutant. High concentrations of imidazole are much less effective than His substrates in promoting hydrolysis by the H64A variant, suggesting that the His residue on the bound (not free) substrate is involved in catalysis. The reduction in catalytic efficiency kcat/KM for hydrolysis of the amide substrate upon replacement of the oxyanion stabilizing asparagine (N155G) is only 7-fold greater for wild-type than H64A subtilisin. In contrast, the reductions in kcat/KM upon replacement of the catalytic serine (S221A) or aspartate (D32A) are about 3000-fold greater for wild-type than H64A subtilisin, suggesting that the functional interactions between the Asp32 and Ser221 with the substrate histidine are more compromised in substrate-assisted catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Primary cilia are antenna‐like sensory organelles protruding from the plasma membrane. Defects in ciliogenesis cause diverse genetic disorders. NDR2 was identified as the causal gene for a canine ciliopathy, early retinal degeneration, but its role in ciliogenesis remains unknown. Ciliary membranes are generated by transport and fusion of Golgi‐derived vesicles to the pericentrosome, a process requiring Rab11‐mediated recruitment of Rabin8, a GDP–GTP exchange factor (GEF) for Rab8, and subsequent Rab8 activation and Rabin8 binding to Sec15, a component of the exocyst that mediates vesicle tethering. This study shows that NDR2 phosphorylates Rabin8 at Ser‐272 and defects in this phosphorylation impair preciliary membrane assembly and ciliogenesis, resulting in accumulation of Rabin8‐/Rab11‐containing vesicles at the pericentrosome. Rabin8 binds to and colocalizes with GTP‐bound Rab11 and phosphatidylserine (PS) on pericentrosomal vesicles. The phospho‐mimetic S272E mutation of Rabin8 decreases affinity for PS but increases affinity for Sec15. These results suggest that NDR2‐mediated Rabin8 phosphorylation is crucial for ciliogenesis by triggering the switch in binding specificity of Rabin8 from PS to Sec15, thereby promoting local activation of Rab8 and ciliary membrane formation.  相似文献   

20.
p-Hydroxybenzoate hydroxylase is extensively studied as a model for single-component flavoprotein monooxygenases. It catalyzes a reaction in two parts: (1) reduction of the FAD in the enzyme by NADPH in response to binding of p-hydroxybenzoate to the enzyme and (2) oxidation of reduced FAD with oxygen in an environment free from solvent to form a hydroperoxide, which then reacts with p-hydroxybenzoate to form an oxygenated product. These different reactions are coordinated through conformational rearrangements of the protein and the isoalloxazine ring during catalysis. Until recently, it has not been clear how p-hydroxybenzoate gains access to the buried active site. In 2002, a structure of a mutant form of the enzyme without substrate was published that showed an open conformation with solvent access to the active site [Wang, J., Ortiz-Maldonado, M., Entsch, B., Massey, V., Ballou, D., and Gatti, D. L. (2002) Proc. Natl. Acad. Sci. U.S.A. 99, 608-613]. The wild-type enzyme does not form high-resolution crystals without substrate. We hypothesized that the wild-type enzyme without substrate also forms an open conformation for binding p-hydroxybenzoate, but only transiently. To test this idea, we have studied the properties of two different mutant forms of the enzyme that are stabilized in the open conformation. These mutant enzymes bind p-hydroxybenzoate very fast, but with very low affinity, as expected from the open structure. The mutant enzymes are extremely inactive, but are capable of slowly forming small amounts of product by the normal catalytic pathway. The lack of activity results from the failure of the mutants to readily form the out conformation required for flavin reduction by NADPH. The mutants form a large fraction of an abnormal conformation of the reduced enzyme with p-hydroxybenzoate bound. This conformation of the enzyme is unreactive with oxygen. We conclude that transient formation of this open conformation is the mechanism for sequestering p-hydroxybenzoate to initiate catalysis. This overall study emphasizes the role that protein dynamics can play in enzymatic catalysis.  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