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 46 毫秒
1.
The ability of an arginine residue to function as the active site acid catalyst in the fumarate reductase family of enzymes is now well-established. Recently, a dual role for the arginine during fumarate reduction has been proposed [Mowat, C. G., Moysey, R., Miles, C. S., Leys, D., Doherty, M. K., Taylor, P., Walkinshaw, M. D., Reid, G. A., and Chapman, S. K. (2001) Biochemistry 40, 12292-12298] in which it acts both as a Lewis acid in transition-state stabilization and as a Br?nsted acid in proton delivery. This proposal has led to the prediction that, if appropriately positioned, a water molecule would be capable of functioning as the active site Br?nsted acid. In this paper, we describe the construction and kinetic and crystallographic analysis of the Q363F single mutant and Q363F/R402A double mutant forms of flavocytochrome c(3), the soluble fumarate reductase from Shewanella frigidimarina. Although replacement of the active site acid, Arg402, with alanine has been shown to eliminate fumarate reductase activity, this phenomenon is partially reversed by the additional substitution of Gln363 with phenylalanine. This Gln --> Phe substitution in the inactive R402A mutant enzyme was designed to "push" a water molecule close enough to the substrate C3 atom to allow it to act as a Br?nsted acid. The 2.0 A resolution crystal structure of the Q363F/R402A mutant enzyme does indeed reveal the introduction of a water molecule at the correct position in the active site to allow it to act as the catalytic proton donor. The 1.8 A resolution crystal structure of the Q363F mutant enzyme shows a water molecule similarly positioned, which can account for its measured fumarate reductase activity. However, in this mutant enzyme Michaelis complex formation is impaired due to significant and unpredicted structural changes at the active site.  相似文献   

2.
There is now overwhelming evidence supporting a common mechanism for fumarate reduction in the respiratory fumarate reductases. The X-ray structures of substrate-bound forms of these enzymes indicate that the substrate is well positioned to accept a hydride from FAD and a proton from an arginine side chain. Recent work on the enzyme from Shewanella frigidimarina [Doherty, M. K., Pealing, S. L., Miles, C. S., Moysey, R., Taylor, P., Walkinshaw, M. D., Reid, G. A., and Chapman, S. K. (2000) Biochemistry 39, 10695-10701] has strengthened the assignment of an arginine (Arg402) as the proton donor in fumarate reduction. Here we describe the crystallographic and kinetic analyses of the R402A, R402K, and R402Y mutant forms of the Shewanella enzyme. The crystal structure of the R402A mutant (2.0 A resolution) shows it to be virtually identical to the wild-type enzyme, apart from the fact that a water molecule occupies the position previously taken by part of the guanidine group of R402. Although structurally similar to the wild-type enzyme, the R402A mutant is inactive under all the conditions that were studied. This implies that a water molecule, in this position in the active site, cannot function as the proton donor for fumarate reduction. In contrast to the R402A mutation, both the R402K and R402Y mutant enzymes are active. Although this activity was at a very low level (at pH 7.2 some 10(4)-fold lower than that for the wild type), it does imply that both lysine and tyrosine can fulfill the role of an active site proton donor, albeit very poorly. The crystal structures of the R402K and R402Y mutant enzymes (2.0 A resolution) show that distances from the lysine and tyrosine side chains to the nearest carbon atom of fumarate are approximately 3.5 A, clearly permitting proton transfer. The combined results from mutagenesis, crystallographic, and kinetic studies provide formidable evidence that R402 acts as both a Lewis acid (stabilizing the build-up of negative charge upon hydride transfer from FAD to fumarate) and a Br?nsted acid (donating the proton to the substrate to complete the formation of succinate).  相似文献   

3.
Engineering disulfide bridges is a common technique to lock a protein movement in a defined conformational state. We have designed two double mutants of Escherichia coli 5'-nucleotidase to trap the enzyme in both an open (S228C, P513C) and a closed (P90C, L424C) conformation by the formation of disulfide bridges. The mutant proteins have been expressed, purified, and crystallized, to structurally characterize the designed variants. The S228C, P513C is a double mutant crystallized in two different crystal forms with three independent conformers, which differ from each other by a rotation of up to 12 degrees of the C-terminal domain with respect to the N-terminal domain. This finding, as well as an analysis of the domain motion in the crystal, indicates that the enzyme still exhibits considerable residual domain flexibility. In the double mutant that was designed to trap the enzyme in the closed conformation, the structure analysis reveals an unexpected intermediate conformation along the 96 degrees rotation trajectory between the open and closed enzyme forms. A comparison of the five independent conformers analyzed in this study shows that the domain movement of the variant enzymes is characterized by a sliding movement of the residues of the domain interface along the interface, which is in contrast to a classical closure motion where the residues of the domain interface move perpendicular to the interface.  相似文献   

4.
The human type 1 (placenta, breast tumors) and type 2 (gonads, adrenals) isoforms of 3beta-hydroxysteroid dehydrogenase/isomerase (3beta-HSD) are key enzymes in steroidogenic pathways leading to the production of all active steroid hormones. Kinetic analyses of purified 3beta-HSD1 show that the Michaelis-Menten constants (Km) for substrates and cofactor are decreased dramatically (three- to eight-fold) by the addition of beta-mercaptoethanol (BME), which suggest that a disulfide bond may be critical to ligand utilization. Western immunoblots and SDS-PAGE of purified 3beta-HSD1 in the presence or absence of BME showed a lack of intersubunit disulfide bonds in the dimeric enzyme. The Rossmann-fold domain of 3beta-HSD1 contains two Cys residues, Cys72 and Cys111, which are capable of forming an intrasubunit disulfide bond based on their proximity in our structural model. Our structural model also predicts that Cys83 may affect the orientation of substrate and cofactor. To test these predictions, the C72S, C72F, C111S, C111A, C83S and C83A mutants of 3beta-HSD1 were produced, expressed, and purified. BME failed to diminish the Km values of substrate and cofactor for C72S, C72F, C111S and C111A but produced a 2.5 decrease in Km values for C83A ligands similar to wild-type 3beta-HSD. Thus, our results support the presence of an intrasubunit disulfide bond between Cys72 and Cys111 that participates in the tertiary structure of the Rossmann-fold domain. Although C83S had no enzyme activity, the C83A mutant enzyme exhibited two- to five-fold higher Km values for substrate and cofactor but had similar K(cat) values compared to wild-type 3beta-HSD. These data characterize the roles of Cys residues in 3beta-HSD and validate the predictions of our structural model.  相似文献   

5.
A disulfide bond introduced between amino acid positions 9 and 164 in phage T4 lysozyme has been shown to significantly increase the stability of the enzyme toward thermal denaturation [Matsumura, M., Becktel, W.J., Levitt, M., & Matthews, B. W. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 6562-6566]. To elucidate the structural features of the engineered disulfide, the crystal structure of the disulfide mutant has been determined at 1.8-A resolution. Residue 9 lies in the N-terminal alpha-helix, while residue 164 is located at the extreme C terminus of T4 lysozyme, which is the most mobile part of the molecule. The refined structure shows that the formation of the disulfide bond is accompanied by relatively large (approximately 2.5 A) localized shifts in C-terminal main-chain atoms. Comparison of the geometry of the engineered disulfide with those of naturally observed disulfides in proteins shows that the engineered bridge adopts a left-handed spiral conformation with a typical set of dihedral angles and C alpha-C alpha distance. The geometry of the engineered disulfide suggests that it is slightly more strained than the disulfide of oxidized dithiothreitol but that the strain is within the range observed in naturally occurring disulfides. The wild-type and cross-linked lysozymes have very similar overall crystallographic temperature factors, indicating that the introduction of the disulfide bond does not impose rigidity on the folded protein structure. In particular, residues 162-164 retain high mobility in the mutant structure, consistent with the idea that stabilization of the protein is due to the effect of the disulfide cross-link on the unfolded rather than the folded state.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
In eukaryotes, DNA topoisomerase I (Top1) catalyzes the relaxation of supercoiled DNA by a conserved mechanism of transient DNA strand breakage, rotation, and religation. The unusual architecture of the monomeric human enzyme comprises a conserved protein clamp, which is tightly wrapped about duplex DNA, and an extended coiled-coil linker domain that appropriately positions the C-terminal active site tyrosine domain against the Top1 core to form the catalytic pocket. A structurally undefined N-terminal domain, dispensable for enzyme activity, mediates protein-protein interactions. Previously, reversible disulfide bonds were designed to assess whether locking the Top1 clamp around duplex DNA would restrict DNA strand rotation within the covalent Top1-DNA intermediate. The active site proximal disulfide bond in full-length Top1-clamp(534) restricted DNA rotation (Woo, M. H., Losasso, C., Guo, H., Pattarello, L., Benedetti, P., and Bjornsti, M. A. (2003) Proc. Natl. Acad. Sci. U. S. A. 100, 13767-13772), whereas the more distal disulfide bond of the N-terminally truncated Topo70-clamp(499) did not (Carey, J. F., Schultz, S. J., Sisson, L., Fazzio, T. G., and Champoux, J. J. (2003) Proc. Natl. Acad. Sci. U. S. A. 100, 5640-5645). To assess the contribution of the N-terminal domain to the dynamics of Top1 clamping of DNA, the same disulfide bonds were engineered into full-length Top1 and truncated Topo70, and the activities of these proteins were assessed in vitro and in yeast. Here we report that the N terminus impacts the opening and closing of the Top1 protein clamp. We also show that the architecture of yeast and human Top1 is conserved in so far as cysteine substitutions of the corresponding residues suffice to lock the Top1-clamp. However, the composition of the divergent N-terminal/linker domains impacts Top1-clamp activity and stability in vivo.  相似文献   

7.
Mirabilis antiviral protein (MAP) is a rigid, heat-stable protein composed of 250 amino acids with an intramolecular disulfide bond. MAP inhibits the in vitro protein synthesis of rabbit reticulocyte with approximately one-thirtieth the activity of the ricin A chain, a homologous protein with no such bond (Habuka, N., Murakami, Y., Noma, M., Kudo, T., and Horikoshi, K. (1989) J. Biol. Chem. 264, 6629-6637; Habuka, N., Akiyama, K., Tsuge, H., Miyano, M., Matsumoto, T., and Noma, M. (1990) J. Biol. Chem. 265, 10988-10992). The bond is presumed to induce some structural perturbation that alters the mode of interaction with the substrate ribosome and thus lowers the activity. To confirm this hypothesis, a mutant MAP gene in which the codons of both cysteines were replaced by those of serines was constructed and expressed in Escherichia coli, and its product (C36/22OS) was purified. In a sodium dodecyl sulfate-polyacrylamide gel electrophoresis, C36/220S showed the same mobility as that of MAP reduced by 2-mercaptoethanol, whereas nonreduced MAP showed faster migration. The inhibitory activity of C36/220S was approximately 22 times higher than that of native MAP, that is the mutant had an IC50 of 0.16 nM for the protein synthesis of the rabbit reticulocyte system, whereas the native MAP had an IC50 of 3.5 nM. The results indicate that the activity of MAP is increased by the elimination of the disulfide bond, and this supports the hypothesis.  相似文献   

8.
From the comparison of the three-dimensional structure of mesophilic pyroglutamyl peptidase from Bacillus amyloliquefaciens and the thermophilic enzyme from Thermococcus litoralis, the intersubunit disulfide bond was estimated to be one of the factors for thermal stability. Since Ser185 was corresponded to Cys190 of the thermophilic enzyme by sequence alignment, the Ser185 residue was replaced with cysteine by site-directed mutagenesis. The S185C mutant enzyme appeared to form a disulfide bond, which was confirmed by SDS-PAGE with and without 2-mercaptoethanol. The mutant enzyme showed a catalytic efficiency equivalent to that of the wild-type enzyme for hydrolysis of a synthetic peptide substrate. However, the thermal stability of the S185C mutant was found to be 30 degrees C higher than that of wild-type. Thus the introduction of a disulfide bond enhanced thermal stability without changing the catalytic efficiency of the enzyme.  相似文献   

9.
The active sites of respiratory fumarate reductases are highly conserved, indicating a common mechanism of action involving hydride and proton transfer. Evidence from the X-ray structures of substrate-bound fumarate reductases, including that for the enzyme from Shewanella frigidimarina [Taylor, P., Pealing, S. L., Reid, G. A., Chapman, S. K., and Walkinshaw, M. D. (1999) Nat. Struct. Biol. 6, 1108-1112], indicates that the substrate is well positioned to accept a hydride from N5 of the FAD. However, the identity of the proton donor has been the subject of recent debate and has been variously proposed to be (using numbering for the S. frigidimarina enzyme) His365, His504, and Arg402. We have used site-directed mutagenesis to examine the roles of these residues in the S. frigidimarina enzyme. The H365A and H504A mutant enzymes exhibited lower k(cat) values than the wild-type enzyme but only by factors of 3-15, depending on pH. This, coupled with the increase in K(m) observed for these enzymes, indicates that His365 and His504 are involved in Michaelis complex formation and are not essential catalytic residues. In fact, examination of the crystal structure of S. frigidimarina fumarate reductase has led to the proposal that Arg402 is the only plausible active site acid. Consistent with this proposal, we report that the R402A mutant enzyme has no detectable fumarate reductase activity. The crystal structure of the H365A mutant enzyme shows that, in addition to the replacement at position 365, there have been some adjustments in the positions of active site residues. In particular, the observed change in the orientation of the Arg402 side chain could account for the decrease in k(cat) seen with the H365A enzyme. These results demonstrate that an active site arginine and not a histidine residue is the proton donor for fumarate reduction.  相似文献   

10.
The engineered disulfide bridge between residues 21 and 142 of phage T4 lysozyme spans the active-site cleft and can be used as a switch to control the activity of the enzyme (Matsumura, M. & Matthews, B.W., 1989, Science 243, 792-794). In the oxidized form the disulfide increases the melting temperature of the protein by 11 degrees C at pH 2. The crystal structure of this mutant lysozyme has been determined in both the reduced and oxidized forms. In the reduced form, the crystal structure of the mutant is shown to be extremely similar to that of wild type. In the oxidized form, however, the formation of the disulfide bridge causes the alpha-carbons of Cys 21 and Cys 142, on opposite sides of the active-site cleft, to move toward each other by 2.5 A. In association with this movement, the amino-terminal domain of the protein undergoes a rigid-body rotation of 5.1 degrees relative to the carboxy-terminal domain. This rotation occurs about an axis passing through the junction of the amino-terminal and carboxy-terminal domains and is also close to the axis that best fits the apparent thermal motion of the amino-terminal domain seen previously in crystals of wild-type lysozyme. Even though the engineered Cys 21-Cys 142 disulfide links together the amino-terminal and carboxy-terminal domains of T4 lysozyme, it does not reduce the apparent mobility of the one domain relative to the other. The pronounced "hinge-bending" mobility of the amino-terminal domain that is suggested by the crystallographic thermal parameters of wild-type lysozyme persists in the oxidized (and reduced) mutant structures. In the immediate vicinity of the introduced disulfide bridge the mutant structure is more mobile (or disordered) than wild type, so much so that the exact conformation of Cys 21 remains obscure. As with the previously described disulfide bridge between residues 9 and 164 of T4 lysozyme (Pjura, P.E., Matsumura, M., Wozniak, J.A., & Matthews, B.W., 1990, Biochemistry 29, 2592-2598), the engineered cross-link substantially enhances the stability of the protein without making the folded structure more rigid.  相似文献   

11.
Molecular dynamics simulations as a tool for improving protein stability   总被引:1,自引:0,他引:1  
Haloalkane dehalogenase (DhlA) was used as a model protein to explore the possibility to use molecular dynamics (MD) simulations as a tool to identify flexible regions in proteins that can serve as a target for stability enhancement by introduction of a disulfide bond. DhlA consists of two domains: an alpha/beta-hydrolase fold main domain and a cap domain composed of five alpha-helices. MD simulations of DhlA showed high mobility in a helix-loop-helix region in the cap domain, involving residues 184-211. A disulfide cross-link was engineered between residue 201 of this flexible region and residue 16 of the main domain. The mutant enzyme showed substantial changes in both thermal and urea denaturation. The oxidized form of the mutant enzyme showed an increase of the apparent transition temperature from 47.5 to 52.5 degrees C, whereas the T(m,app) of the reduced mutant decreased by more than 8 degrees C compared to the wild-type enzyme. Urea denaturation results showed a similar trend. Measurement of the kinetic stability showed that the introduction of the disulfide bond caused a decrease in activation free energy of unfolding of 0.43 kcal mol(-1) compared to the wild-type enzyme and also indicated that the helix-loop-helix region was involved early in the unfolding process. The results show that MD simulations are capable of identifying mobile protein domains that can successfully be used as a target for stability enhancement by the introduction of a disulfide cross-link.  相似文献   

12.
The 1.8 A resolution crystal structure of the tetraheme flavocytochrome c3, Fcc3, provides the first mechanistic insight into respiratory fumarate reductases or succinate dehydrogenases. The multi-redox center, three-domain protein shows a 40 A long 'molecular wire' allowing rapid conduction of electrons through a new type of cytochrome domain onto the active site flavin, driving the reduction of fumarate to succinate. In this structure a malate-like molecule is trapped in the enzyme active site. The interactions between this molecule and the enzyme suggest a clear mechanism for fumarate reduction in which the substrate is polarized and twisted, facilitating hydride transfer from the reduced flavin and subsequent proton transfer. The enzyme active site in the oxidized form is completely buried at the interface between the flavin-binding and the clamp domains. Movement of the cytochrome and clamp domains is postulated to allow release of the product.  相似文献   

13.
To study the essentiality of head domain movement of the Rieske iron-sulfur protein (ISP) during bc(1) catalysis, Rhodobacter sphaeroides mutants expressing His-tagged cytochrome bc(1) complexes with three pairs of cysteines engineered (one cysteine each) on the interface between cytochrome b and ISP, A185C(cytb)/K70C(ISP), I326C(cytb)/G165C(ISP), and T386C(cytb)/K164C(ISP), were generated and characterized. Formation of an intersubunit disulfide bond between cytochrome b and ISP is detected in membrane (intracytoplasmic membrane and air-aged chromatophore), and purified bc(1) complex was prepared from the A185C(cytb)/K70C(ISP) mutant cells. Formation of the intersubunit disulfide bond in this cysteine pair mutant complex is concurrent with the loss of its bc(1) activity. Reduction of this disulfide bond by beta-mercaptoethanol restores activity, indicating that mobility of the head domain of ISP is functionally important in the cytochrome bc(1) complex. The rate of intramolecular electron transfer, between 2Fe2S and heme c(1), in the A185C(cytb)/K70C(ISP) mutant complex is much lower than that in the wild type or in their respective single cysteine mutant complexes, indicating that formation of an intersubunit disulfide bond between cytochrome b and ISP arrests the head domain of ISP in the "fixed state" position, which is too far for electron transfer to heme c(1).  相似文献   

14.
Flavocytochrome b(2) from Saccharomyces cerevisiae is a l-lactate/cytochrome c oxidoreductase belonging to a large family of 2-hydroxyacid-dependent flavoenzymes. The crystal structure of the enzyme, with pyruvate bound at the active site, has been determined [Xia, Z.-X., and Mathews, F. S. (1990) J. Mol. Biol. 212, 837-863]. The authors indicate that the methyl group of pyruvate is in close contact with Ala198 and Leu230. These two residues are not well-conserved throughout the family of (S)-2-hydroxy acid oxidases/dehydrogenases. Thus, to probe substrate specificity in flavocytochrome b(2), these residues have been substituted by glycine and alanine, respectively. Kinetic studies on the L230A mutant enzyme and the A198G/L230A double mutant enzyme indicate a change in substrate selectivity for the enzyme toward larger (S)-2-hydroxy acids. In particular, the L230A enzyme is more efficient at utilizing (S)-2-hydroxyoctanoate by a factor of 40 as compared to the wild-type enzyme [Daff, S., Manson, F. D. C., Reid, G. A., and Chapman, S. K. (1994) Biochem. J. 301, 829-834], and the A198G/L230A double mutant enzyme is 6-fold more efficient with the aromatic substrate l-mandelate than it is with l-lactate [Sinclair, R., Reid, G. A., and Chapman, S. K. (1998) Biochem. J. 333, 117-120]. To complement these solution studies, we have solved the structure of the A198G/L230A enzyme in complex with pyruvate and as the FMN-sulfite adduct (both to 2.7 A resolution). We have also obtained the structure of the L230A mutant enzyme in complex with phenylglyoxylate (the product of mandelate oxidation) to 3.0 A resolution. These structures reveal the increased active-site volume available for binding larger substrates, while also confirming that the integrity of the interactions important for catalysis is maintained. In addition to this, the mode of binding of the bulky phenylglyoxylate at the active site is in accordance with the operation of a hydride transfer mechanism for substrate oxidation/flavin reduction in flavocytochrome b(2), whereas a mechanism involving the formation of a carbanion intermediate would appear to be sterically prohibited.  相似文献   

15.
Complex II superfamily members catalyze the kinetically difficult interconversion of succinate and fumarate. Due to the relative simplicity of complex II substrates and their similarity to other biologically abundant small molecules, substrate specificity presents a challenge in this system. In order to identify determinants for on-pathway catalysis, off-pathway catalysis, and enzyme inhibition, crystal structures of Escherichia coli menaquinol:fumarate reductase (QFR), a complex II superfamily member, were determined bound to the substrate, fumarate, and the inhibitors oxaloacetate, glutarate, and 3-nitropropionate. Optical difference spectroscopy and computational modeling support a model where QFR twists the dicarboxylate, activating it for catalysis. Orientation of the C2-C3 double bond of activated fumarate parallel to the C(4a)-N5 bond of FAD allows orbital overlap between the substrate and the cofactor, priming the substrate for nucleophilic attack. Off-pathway catalysis, such as the conversion of malate to oxaloacetate or the activation of the toxin 3-nitropropionate may occur when inhibitors bind with a similarly activated bond in the same position. Conversely, inhibitors that do not orient an activatable bond in this manner, such as glutarate and citrate, are excluded from catalysis and act as inhibitors of substrate binding. These results support a model where electronic interactions via geometric constraint and orbital steering underlie catalysis by QFR.  相似文献   

16.
Chloride-dependent alpha-amylases constitute a well conserved family of enzymes thereby allowing investigation of the characteristics of each member to understand, for example, relevant properties required for environmental adaptation. In this context, we have constructed a double mutant (Q58C/A99C) of the cold-active and heat-labile alpha-amylase from the Antarctic bacterium Pseudoalteromonas haloplanktis, defined on the basis of its strong similarity with the mesophilic enzyme from pig pancreas. This mutant was characterized to understand the role of an extra disulfide bond specific to warm-blooded animals and located near the entrance of the catalytic cleft. We show that the catalytic parameters of the mutant are drastically modified and similar to those of the mesophilic enzyme. Calorimetric studies demonstrated that the mutant is globally stabilized (DeltaDeltaG = 1.87 kcal/mol at 20 degrees C) when compared with the wild-type enzyme, although the melting point (T(m)) was not increased. Moreover, fluorescence quenching experiments indicate a more compact structure for the mutated alpha-amylase. However, the strain imposed on the active site architecture induces a 2-fold higher thermal inactivation rate at 45 degrees C as well as the appearance of a less stable calorimetric domain. It is concluded that stabilization by the extra disulfide bond arises from an enthalpy-entropy compensation effect favoring the enthalpic contribution.  相似文献   

17.
J Eder  M Wilmanns 《Biochemistry》1992,31(18):4437-4444
A disulfide bond has been introduced in the beta/alpha-barrel enzyme N-(5'-phosphoribosyl)anthranilate isomerase from Saccharomyces cerevisiae. The design of this disulfide bond was based on a model structure of this enzyme, built from the high-resolution crystal structure of the N-(5'-phosphoribosyl)anthranilate isomerase domain from Escherichia coli. The disulfide cross-link is spontaneously formed in vitro between residues 27 and 212, located in the structurally adjacent alpha-helices 1 and 8 of the outer helical ring of the beta/alpha-barrel. It creates a loop of 184 residues that account for 83% of the sequence of this enzyme, thus forming a quasi circular protein. The cross-linked mutant enzyme displays wild-type steady-state kinetic parameters. Measurements of the equilibrium constant for the reduction of this disulfide bond by 1,4-dithiothreitol show that its bond strength is comparable to that of other engineered protein disulfide bonds. The oxidized, cross-linked N-(5'-phosphoribosyl)anthranilate isomerase mutant is about 1.0 kcal/mol more stable than the wild-type enzyme, as estimated from its equilibrium unfolding transitions by guanidine hydrochloride.  相似文献   

18.
Cholera toxin (CT) produced by Vibrio cholerae and heat-labile enterotoxin (LT-I), produced by enterotoxigenic Escherichia coli, are AB5 heterohexamers with an ADP-ribosylating A subunit and a GM1 receptor binding B pentamer. These toxins are among the most potent mucosal adjuvants known and, hence, are of interest both for the development of anti-diarrheal vaccines against cholera or enterotoxigenic Escherichia coli diarrhea and also for vaccines in general. However, the A subunits of CT and LT-I are known to be relatively temperature sensitive. To improve the thermostability of LT-I an additional disulfide bond was introduced in the A1 subunit by means of the double mutation N40C and G166C. The crystal structure of this double mutant of LT-I has been determined to 2.0 A resolution. The protein structure of the N40C/G166C double mutant is very similar to the native structure except for a few local shifts near the new disulfide bond. The introduction of this additional disulfide bond increases the thermal stability of the A subunit of LT-I by 6 degrees C. The enhancement in thermostability could make this disulfide bond variant of LT-I of considerable interest for the design of enterotoxin-based vaccines.  相似文献   

19.
Cysteine residues 110 and 187 are essential for the formation of the correct bovine rhodopsin structure (Karnik, S. S., Sakmar, T. P., Chen, H.-B., and Khorana, H. G. (1988) Proc. Natl. Acad. Sci. U. S. A. 85, 8459-8463). We now show that the sulfhydryl groups of these 2 cysteine residues interact to form a disulfide bond. Rhodopsin mutants containing cysteine----serine substitutions were prepared as follows. In one mutant, CysVII, all the 10 cysteine residues of rhodopsin were replaced by serines. A second mutant, CysVIII, contained only C110 and C185; a third mutant, CysIX, contained only C185 and C187 while the fourth mutant, CysX, contained only C110 and C187. Only mutant CysX formed functional rhodopsin. Mutants CysVIII and CysIX reacted with [3H]iodoacetic acid showing the presence of free sulfhydryl groups while mutant CysX was inert to this reagent. CysX reacted with cyanide ion to form a thiocyanate derivative showing the presence of a disulfide bond. The C110-C187 disulfide bond is buried in rhodopsin because reactions with disulfide reducing agents and cyanide ion require prior treatment with denaturants.  相似文献   

20.
The contribution of the disulfide bridge in CotA-laccase from Bacillus subtilis is assessed with respect to the enzyme’s functional and structural properties. The removal of the disulfide bond by site-directed mutagenesis, creating the C322A mutant, does not affect the spectroscopic or catalytic properties and, surprisingly, neither the long-term nor the thermodynamic stability parameters of the enzyme. Furthermore, the crystal structure of the C322A mutant indicates that the overall structure is essentially the same as that of the wild type, with only slight alterations evident in the immediate proximity of the mutation. In the mutant enzyme, the loop containing the C322 residue becomes less ordered, suggesting perturbations to the substrate binding pocket. Despite the wild type and the C322A mutant showing similar thermodynamic stability in equilibrium, the holo or apo forms of the mutant unfold at faster rates than the wild-type enzyme. The picosecond to nanosecond time range dynamics of the mutant enzyme was not affected as shown by acrylamide collisional fluorescence quenching analysis. Interestingly, copper uptake or copper release as measured by the stopped-flow technique also occurs more rapidly in the C322A mutant than in the wild-type enzyme. Overall the structural and kinetic data presented here suggest that the disulfide bridge in CotA-laccase contributes to the conformational dynamics of the protein on the microsecond to millisecond timescale, with implications for the rates of copper incorporation into and release from the catalytic centres.  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