首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
PLC(Bc) is a 28.5 kDa monomeric enzyme that catalyzes the hydrolysis of the phosphodiester bond of phosphatidylcholine, phosphatidylethanolamine, and phosphatidylserine to provide a diacylglycerol and the corresponding phosphorylated headgroup. Because single replacements of Glu4, Tyr56, and Phe66 in the headgroup binding pocket led to changes in substrate specificity [Martin et al. (2000) Biochemistry 39, 3410-3415], a combinatorial library of approximately 6000 maltose binding protein-PLC(Bc) fusion protein mutants containing random permutations of these three residues was generated to identify PLC(Bc) mutants with altered specificity profiles and high catalytic activities. Members of this library were screened for hydrolytic activity toward the water soluble substrates C6PC, C6PE, and C6PS using a novel protocol that was conducted in a 96-well format and featured the in situ cleavage of the fusion protein to release the mutant PLC(Bc)s. Ten mutant enzymes that exhibited significant preferences toward C6PE or C6PS were selected and analyzed by steady-state kinetics to determine their specificity constants, k(cat)/K(M). The C6PS selective clones E4G, E4Q/Y56T/F66Y, and E4K/Y56V exhibited higher specificity constants toward C6PS than wt, whereas Y56T, F66Y, and Y56T/F66Y were C6PE selective and had comparable or higher specificity constants than wt for C6PE. The corresponding wt residues were singly reinserted back into the E4Q/Y56T/F66Y and E4K/Y56V mutants via site-directed mutagenesis, and the E4Q/F66Y mutant thus obtained exhibited a 10-fold higher specificity constant toward C6PS than wt, a value significantly higher than other PLC(Bc) mutants. On the basis of available data, an aromatic residue at position 66 appears important for significant catalytic activity toward all three substrates, especially C6PC and C6PE. The charge of residue 4 also appears to be a determinant of enzyme specificity as a negatively charged residue at this position endows the enzyme with C6PC and C6PE preference, whereas a polar neutral or positively charged residue results in C6PS selectivity. Replacing Tyr56 with Val, Ala, Thr, or Ser greatly reduces activity toward C6PC. Thus, the substrate specificity of PLC(Bc) can be modulated by varying three of the amino acid residues that constitute the headgroup binding pocket, and it is now apparent that this enzyme is not evolutionarily optimized to hydrolyze phospholipids with ethanolamine or serine headgroups.  相似文献   

2.
The phosphatidylcholine preferring phospholipase C from Bacillus cereus (PC-PLC(Bc)) catalyzes the hydrolysis of phospholipids in the following order of preference: phosphatidylcholine (PC)>phosphatidylethanolamine (PE)>phosphatidylserine (PS). In previous work, mutagenic, kinetic, and crystallographic experiments suggested that varying the amino acids at the 4th, 56th, and 66th positions had a significant influence upon the substrate specificity profile of PC-PLC(Bc). Here, we report the crystal structures of the native form of several PC-PLC(Bc) variants that exhibited altered substrate specificities for PC, PE, and PS at maximum resolutions of 1.90-2.05 Angstrom. Comparing the structures of these variants to the structure of the wild-type enzyme reveals only minor differences with respect to the number and location of active site water molecules and the side chain conformations of residues at the 4th and 56th positions. These results suggest that subtle changes in steric and electronic properties in the substrate binding site of PC-PLC(Bc) are responsible for the significant changes in substrate selectivity.  相似文献   

3.
Histidine 30 in human manganese superoxide dismutase (MnSOD) is located at a site partially exposed to solvent with its side chain participating in a hydrogen-bonded network that includes the active-site residues Tyr(166) and Tyr(34) and extends to the manganese-bound solvent molecule. We have replaced His(30) with a series of amino acids and Tyr(166) with Phe in human MnSOD. The crystal structure of the mutant of MnSOD containing Asn(30) superimposed closely with the wild type, but the side chain of Asn(30) did not participate in the hydrogen-bonded network in the active site. The catalytic activity of a number of mutants with replacements at position 30 and for the mutant containing Phe(166) showed a 10-40-fold decrease in k(cat). This is the same magnitude of decrease in k(cat) obtained with the replacement of Tyr(34) by Phe, suggesting that interrupting the hydrogen-bonded active-site network at any of the sites of these three participants (His(30), Tyr(34), and Tyr(166)) leads to an equivalent decrease in k(cat) and probably less efficient proton transfer to product peroxide. The specific geometry of His(30) on the hydrogen bond network is essential for stability since the disparate mutations H30S, H30A, and H30Q reduce T(m) by similar amounts (10-16 degrees C) compared with wild type.  相似文献   

4.
The phosphatidylcholine (PC)-preferring phospholipase C (PLC) from Bacillus cereus (PLC(Bc)) hydrolyzes various 1,2-diacyl derivatives of PC at different rates. Substrates with side chains having eight or more carbons are present in micellular form in aqueous media and are processed most rapidly. The catalytic efficiency (k(cat)/K(m)) for the hydrolyses of short-chain PCs at concentrations below their respective critical micelle concentrations also decreases as the side chains become shorter, and this loss of efficiency owes its origin to increases in K(m). In order to ascertain whether the observed increases in K(m) might arise from conformational changes in the glycerol backbone, nuclear magnetic resonance (NMR) experiments were performed in D(2)O to determine the (3)J(HH) and (3)J(CH) coupling constants along the glycerol subunit of 1, 2-dipropanoyl-sn-glycero-3-phosphocholine (K(m)=61 mM), 1, 2-dibutanoyl-sn-glycero-3-phosphocholine (K(m)=21.2 mM) and 1, 2-dihexanoyl-sn-glycero-3-phosphocholine (K(m)=2.4 mM). Using these coupling constants, the fractional populations for each rotamer about the backbone of each of substrate were calculated. Two rotamers, which were approximately equally populated, about the sn-1-sn-2 bond of each substrate were significantly preferred, and in these conformers, the oxygens on the sn-1 and sn-2 carbons of the backbone were synclinal to optimize intramolecular hydrophobic interactions between the acyl side chains. There was greater flexibility about the sn-2-sn-3 bond, and each of the three possible staggered conformations was significantly populated, although there was a slight preference for the rotamer in which the oxygen bearing the phosphate head group was synclinal to the oxygen at the sn-2 carbon and to the sn-1 carbon; in this orientation, the head group is folded back relative to the side chains. These studies demonstrate that there is no significant change in the conformation about the glycerol backbone as a function of side chain length in short-chain phospholipids. Thus, prior organization of the substrate seems an unlikely determinant of the catalytic efficiency of PLC(Bc), and other factors such as hydrophobic interactions or differential solvation/desolvation effects associated with the complexation of the substrate with PLC(Bc) may be involved.  相似文献   

5.
Falls LA  Furie B  Furie BC 《Biochemistry》2000,39(43):13216-13222
Phospholipid membranes play a significant role during the proteolytic activation of blood coagulation proteins. This investigation identifies a role for phosphatidylethanolamine (PE) during the activation of factor X by the tenase complex, an enzymatic complex composed of the serine protease, factor IXa, a protein cofactor, factor VIIIa, a phospholipid membrane, and Ca(2+). Phospholipid vesicles composed of PE, phosphatidylserine (PS), and phosphatidylcholine support factor Xa generation. The K(m) and k(cat) for factor X activation by the tenase complex are independent of PE in the presence of 20% PS. At lower PS concentrations, the presence of 20 or 35% PE lowers the K(m) and increases the k(cat) as compared to those in vesicles without PE. The effect of PE on the k(cat) of the tenase complex is mediated through factor VIIIa. PE also enhances factor Xa generation by facilitating tenase complex formation; PE lowers the K(d(app)) of factor IXa for both phospholipid/Ca(2+) and phospholipid/Ca(2+)/factor VIIIa complexes in the presence of suboptimal PS. In addition, the K(d)s of factor IXa and factor X were lower for phospholipid vesicles containing PE. N-Methyl-PE increased the k(cat) and decreased the K(d(app)), whereas N,N-dimethyl-PE had no effect on either parameter, indicating the importance of headgroup size. Lyso-PE had no effect on kinetic parameters, indicating the sn-2 acyl chain dependence of the PE effect. Together, these results demonstrate a role for PE in the assembly and activity of the tenase complex and further extend the understanding of the importance of PE-containing membranes in hemostasis.  相似文献   

6.
The polar headgroup structure of phosphatidylcholine (PC), phosphatidylethanolamine (PE) and phosphatidylserine (PS) in inverted micelles in chloroform or benzene was investigated by the selective 31P(H) nuclear Overhauser effect (NOE). In the frequency dependence of the 31P(1H) NOE, PC micelles in CDCl3 showed two maxima. The larger maximum was located at the resonance of the glycerol-CH2OP protons and the smaller at the resonance of the N-methyl protons. In PC/PE mixed micelles in C6D6, both PC and PE showed three maxima which were located at the resonance of the CH2OP protons, the N-methyl protons and the amino protons in the frequency dependence of the 31P-NOE. The N-methyl protons of PC and the amino protons of PE were closely spaced to the phosphate groups of neighboring lipid molecules. The polar headgroups of PC and PE in the mixed micelles were concluded to lie in the plane perpendicular to the molecular axes. The frequency dependence of the 31P(H) NOE for PS micelles in C6D6 showed the maxima at the resonances of the amino protons and the CH2OP protons. The polar headgroups of PS molecules were not extended parallel to the molecular axes in the inverted micelles.  相似文献   

7.
The contributions to substrate binding and catalysis of 13 amino acid residues of the Caenorhabditis elegans diadenosine tetraphosphate pyrophosphohydrolase (Ap(4)A hydrolase) predicted from the crystal structure of an enzyme-inhibitor complex have been investigated by site-directed mutagenesis. Sixteen glutathione S-transferase-Ap(4)A hydrolase fusion proteins were expressed and their k(cat) and K(m) values determined after removal of the glutathione S-transferase domain. As expected for a Nudix hydrolase, the wild type k(cat) of 23 s(-1) was reduced by 10(5)-, 10(3)-, and 30-fold, respectively, by replacement of the conserved P(4)-phosphate-binding catalytic residues Glu(56), Glu(52), and Glu(103) by Gln. K(m) values were not affected, indicating a lack of importance for substrate binding. In contrast, mutating His(31) to Val or Ala and Lys(83) to Met produced 10- and 16-fold increases in K(m) compared with the wild type value of 8.8 microm. These residues stabilize the P(1)-phosphate. H31V and H31A had a normal k(cat) but K83M showed a 37-fold reduction in k(cat). Lys(36) also stabilizes the P(1)-phosphate and a K36M mutant had a 10-fold reduced k(cat) but a relatively normal K(m). Thus both Lys(36) and Lys(83) may play a role in catalysis. The previously suggested roles of Tyr(27), His(38), Lys(79), and Lys(81) in stabilizing the P(2) and P(3)-phosphates were not confirmed by mutagenesis, indicating the absence of phosphate-specific binding contacts in this region. Also, mutating both Tyr(76) and Tyr(121), which clamp one substrate adenosine moiety between them in the crystal structure, to Ala only increased K(m) 4-fold. It is concluded that interactions with the P(1)- and P(4)-phosphates are minimum and sufficient requirements for substrate binding by this class of enzyme, indicating that it may have a much wider substrate range then previously believed.  相似文献   

8.
The primary specificity residue of a substrate or an inhibitor, called the P(1) residue, is responsible for the proper recognition by the cognate enzyme. This residue enters the S(1) pocket of the enzyme and establishes contacts (up to 50%) inside the proteinase substrate cavity, strongly affecting its specificity. To analyze the influence on bovine alpha-chymotrypsin substrate activity, aromatic non-proteinogenic amino acid residues in position P(1) with the sequence Ac-Phe-Ala-Thr-X-Anb(5,2)-NH(2) were introduced: L-pyridyl alanine (Pal), 4-nitrophenylalanine - Phe(p-NO(2)), 4-aminophenylalanine - Phe(p-NH(2)), 4-carboxyphenylalanine Phe(p-COOH), 4-guanidine phenylalanine - Phe(p-guanidine), 4-methyloxycarbonyl-phenylalanine - Phe(p-COOMe), 4-cyanophenylalanine - Phe(p-CN), Phe, Tyr. The effect of the additional substituent at the phenyl ring of the Phe residue was investigated. All peptides contained an amide of 5-amino-2-nitrobenzoic acid, which served as a chromophore. Kinetic parameters (k(cat), K(M) and k(cat)/K(M)) of the peptides synthesized with bovine alpha-chymotrypsin were determined. The highest value of the specificity constant k(cat)/K(M), reaching 6.0 x 10(5) [M(-1)xs(-1)], was obtained for Ac-Phe-Ala-Thr-Phe(p-NO(2))-Anb(5,2)-NH(2). The replacement of the acetyl group with benzyloxycarbonyl moiety yielded a substrate with the value of k(cat) more than three times higher. Peptide aldehydes were synthesized with selected residues (Phe, Pal, Tyr, Phe(p-NO(2)) in position P(1) and potent chymotrypsin inhibitors were obtained. The dissociation constant (K(i)) with the experimental enzyme determined for the most active peptide, Tos-Phe-Ala-Thr-Phe(p-NO(2))-CHO, amounted to 1.12 x 10(-8) M.  相似文献   

9.
The substrate specificity for phospholipase D from Streptomyces chromofuscus (PLD(Sc)) has been determined utilizing an assay based on the quantitation of inorganic phosphate. 1,2-Di-n-hexanoyl phosphatidylcholine (C6PC), phosphatidylethanolamine (C6PE), phosphatidylserine (C6PS), phosphatidylglycerol (C6PG), and an unnatural phospholipid bearing a neohexyl headgroup (C6PDB) were examined as substrates. The assay relies on the quenching of the PLD(Sc)-catalyzed hydrolysis of the phospholipid substrates with EDTA followed by the hydrolysis of the phosphatidic acid product with alkaline phosphatase. The inorganic phosphate thus released is quantitated through the formation of a complex with ammonium molybdate, which has an absorbance maximum at 700 nm. To minimize the time involved and the reagents consumed, the assay is conducted in 96-well plates. The results of this study indicate that the catalytic efficiency for PLD(Sc) on the substrates is C6PC > C6PS approximately C6PE > C6PG > C6PDB.  相似文献   

10.
We present the first structure of a glycoside hydrolase family 79 β-glucuronidase from Acidobacterium capsulatum, both as a product complex with β-D-glucuronic acid (GlcA) and as its trapped covalent 2-fluoroglucuronyl intermediate. This enzyme consists of a catalytic (β/α)(8)-barrel domain and a β-domain with irregular Greek key motifs that is of unknown function. The enzyme showed β-glucuronidase activity and trace levels of β-glucosidase and β-xylosidase activities. In conjunction with mutagenesis studies, these structures identify the catalytic residues as Glu(173) (acid base) and Glu(287) (nucleophile), consistent with the retaining mechanism demonstrated by (1)H NMR analysis. Glu(45), Tyr(243), Tyr(292)-Gly(294), and Tyr(334) form the catalytic pocket and provide substrate discrimination. Consistent with this, the Y292A mutation, which affects the interaction between the main chains of Gln(293) and Gly(294) and the GlcA carboxyl group, resulted in significant loss of β-glucuronidase activity while retaining the side activities at wild-type levels. Likewise, although the β-glucuronidase activity of the Y334F mutant is ~200-fold lower (k(cat)/K(m)) than that of the wild-type enzyme, the β-glucosidase activity is actually 3 times higher and the β-xylosidase activity is only 2.5-fold lower than the equivalent parameters for wild type, consistent with a role for Tyr(334) in recognition of the C6 position of GlcA. The involvement of Glu(45) in discriminating against binding of the O-methyl group at the C4 position of GlcA is revealed in the fact that the E45D mutant hydrolyzes PNP-β-GlcA approximately 300-fold slower (k(cat)/K(m)) than does the wild-type enzyme, whereas 4-O-methyl-GlcA-containing oligosaccharides are hydrolyzed only 7-fold slower.  相似文献   

11.
The self-assembled supramolecular structures of diacylphosphatidylcholine (diC(n)PC), diacylphosphatidylethanolamine (diC(n)PE), diacylphosphatidyglycerol (diC(n)PG), and diacylphosphatidylserine (diC(n)PS) were investigated by (31)P nuclear magnetic resonance (NMR) spectroscopy as a function of the hydrophobic acyl chain length. Short-chain homologs of these lipids formed micelles, and longer-chain homologs formed bilayers. The shortest acyl chain lengths that supported bilayer structures depended on the headgroup of the lipids. They increased in the order PE (C(6)) < PC (C(9)) < or = PS (C(9) or C(10)) < PG (C(11) or C(12)). This order correlated with the effective headgroup area, which is a function of the physical size, charge, hydration, and hydrogen-bonding capacity of the four headgroups. Electrostatic screening of the headgroup charge with NaCl reduced the effective headgroup area of PS and PG and thereby decreased the micelle-to-bilayer transition of these lipid classes to shorter chain lengths. The experimentally determined supramolecular structures were compared to the assembly states predicted by packing constraints that were calculated from the hydrocarbon-chain volume and effective headgroup area of each lipid. The model accurately predicted the chain-length threshold for bilayer formation if the relative displacement of the acyl chains of the phospholipid were taken into account. The model also predicted cylindrical rather than spherical micelles for all four diacylphospholipid classes and the (31)P-NMR spectra provided evidence for a tubular network that appeared as an intermediate phase at the micelle-to-bilayer transition. The free energy of micellization per methylene group was independent of the structure of the supramolecular assembly, but was -0.95 kJ/mol (-0.23 kcal/mol) for the PGs compared to -2.5 kJ/mol (-0.60 kcal/mol) for the PCs. The integral membrane protein OmpA did not change the bilayer structure of thin (diC(10)PC) bilayers.  相似文献   

12.
The amino acid binding domains of the tryptophanyl (TrpRS)- and tyrosyl-tRNA synthetases (TyrRS) of Bacillus stearothermophilus are highly homologous. These similarities suggest that conserved residues in TrpRS may be responsible for both determining tryptophan recognition and discrimination against tyrosine. This was investigated by the systematic mutation of TrpRS residues based upon the identity of homologous positions in TyrRS. Of the four residues which interact directly with the aromatic side chain of tryptophan (Phe5, Met129, Asp132, and Val141) replacements of Asp132 led to significant changes in the catalytic efficiency of Trp aminoacylation (200-1250-fold reduction in k(cat)/K(M)) and substitution of Val141 by the larger Glu side chain reduced k(cat)/K(M) by 300-fold. Mutation of Pro127, which determines the position of active-site residues, did not significantly effect Trp binding. Of the mutants tested, D132N TrpRS also showed a significant reduction in discrimination against Tyr, with Tyr acting as a competitive inhibitor but not a substrate. The analogous residue in B. stearothermophilusTyrRS (Asp176) has also been implicated as a determinant of amino acid specificity in earlier studies [de Prat Gay, G., Duckworth, H. W., and Fersht, A. R. (1993) FEBS Lett. 318, 167-171]. This striking similarity in the function of a highly conserved residue found in both TrpRS and TyrRS provides mechanistic support for a common origin of the two enzymes.  相似文献   

13.
Sierks MR  Svensson B 《Biochemistry》2000,39(29):8585-8592
Molecular recognition using a series of deoxygenated maltose analogues was used to determine the substrate transition-state binding energy profiles of 10 single-residue mutants at the active site of glucoamylase from Aspergillus niger. The individual contribution of each substrate hydroxyl group to transition-state stabilization with the wild type and each mutant GA was determined from the relation Delta(DeltaG()) = -RT ln[(k(cat)/K(M))(x)/(k(cat)/K(M))(y)], where x represents either a mutant enzyme or substrate analogue and y the wild-type enzyme or parent substrate. The resulting binding energy profiles indicate that disrupting an active site hydrogen bond between enzyme and substrate, as identified in crystal structures, not only sharply reduces or eliminates the energy contributed from that particular hydrogen bond but also perturbs binding contributions from other substrate hydroxyl groups. Replacing the active site acidic groups, Asp55, Glu180, or Asp309, with the corresponding amides, and the neutral Trp178 with the basic Arg, all substantially reduced the binding energy contribution of the 4'- and 6'-OH groups of maltose at subsite -1, even though both Glu180 and Asp309 are localized at subsite 1. In contrast, the substitution, Asp176 --> Asn, located near subsites -1 and 1, did not substantially perturb any of the individual hydroxyl group binding energies. Similarly, the substitutions Tyr116 --> Ala, Ser119 --> Tyr, or Trp120 --> Phe also did not substantially alter the energy profiles even though Trp120 has a critical role in directing conformational changes necessary for activity. Since the mutations at Trp120 and Asp176 reduced k(cat) values by 50- and 12-fold, respectively, a large effect on k(cat) is not necessarily accompanied by changes in hydroxyl group binding energy contributions. Two substitutions, Asn182 --> Ala and Tyr306 --> Phe, had significant though small effects on interactions with 3- and 4'-OH, respectively. Binding interactions between the enzyme and the glucosyl group in subsite -1, particularly with the 4'- and 6'-OH groups, play an important role in substrate binding, while subsite 1 interactions may play a more important role in product release.  相似文献   

14.
Although antiviral nucleoside analog therapy successfully delays progression of HIV infection to AIDS, these drugs cause unwelcome side-effects by inducing mitochondrial toxicity. We and others have demonstrated that the mitochondrial polymerase, DNA polymerase gamma (pol gamma), participates in mitochondrial toxicity by incorporating these chain-terminating antiviral nucleotide analogs into DNA. Here, we explore the role of three highly conserved amino acid residues in the active site of human pol gamma that modulate selection of nucleotide analogs as substrates for incorporation. Sequence alignments, crystal structures and mutagenesis studies of family A DNA polymerases led us to change Tyr951 and Tyr955 in polymerase motif B to Phe and Ala, and Glu895 in polymerase motif A was changed to Ala. The mutant polymerases were tested for their ability to incorporate natural nucleotides and the five antiviral nucleoside analogs currently approved for antiviral therapy: AZT, ddC, D4T, 3TC and carbovir. Steady-state kinetic analysis of the pol gamma derivatives with the normal and antiviral nucleotides demonstrated that Tyr951 is largely responsible for the ability of pol gamma to incorporate dideoxynucleotides and D4T-MP. Mutation of Tyr951 to Phe renders the enzyme resistant to dideoxynucleotides and D4T-TP without compromising the activity of the polymerase. Alteration of Glu895 and Tyr955 to Ala had the largest effect on overall polymerase activity with normal nucleotides, producing dramatic increases in K(m(dNTP)) and large decreases in k(cat). Mutation of Tyr955 in pol gamma causes the degenerative disease progressive external ophthalmoplegia in humans, and we show that this residue partially accounts for the ability of pol gamma to incorporate D4T-MP and carbovir. Alteration of Glu895 to Ala slightly increased discrimination against dideoxynucleotides and D4T-TP. The mechanisms by which pol gamma selects certain nucleotide analogs are discussed.  相似文献   

15.
Molecular cloning of the gene and the crystal structure of the prolyl aminopeptidase [EC 3.4.11.5] from Serratia marcescens have been studied by us [J. Biochem. 122, 601-605 (1997); ibid. 126, 559-565 (1999)]. Through these studies, Phe139, Tyr149, Glu204, and Arg136 were estimated to be concerned with substrate recognition. To elucidate the details of the mechanism for the substrate specificity, the site-directed mutagenesis method was applied. The F139A mutant showed an 80-fold decrease in catalytic efficiency (k(cat)/K(m)), but the Y149A mutant did not show a significant change in catalytic efficiency. The catalytic efficiency of the E204Q mutant was about 4% of that of the wild type. The peptidase activity of the mutant (R136A) was markedly decreased, however, arylamidase activity with Pyr-bNA was retained as in the wild-enzyme. From these results, it was clarified that the pyrrolidine ring and the amino group of proline at the S1 site were recognized by Phe139 and Glu204, respectively. P1' of a substrate was recognized by Arg136. On the other hand, the enzyme had two cysteine residues. Mutants C74A and C271A were inhibited by PCMB, but the double mutated enzyme (C74/271A) was resistant to it.  相似文献   

16.
Hirata A  Adachi M  Utsumi S  Mikami B 《Biochemistry》2004,43(39):12523-12531
The optimum pH of Bacillus cereus beta-amylase (BCB, pH 6.7) differs from that of soybean beta-amylase (SBA, pH 5.4) due to the substitution of a few amino acid residues near the catalytic base residue (Glu 380 in SBA and Glu 367 in BCB). To explore the mechanism for controlling the optimum pH of beta-amylase, five mutants of BCB (Y164E, Y164F, Y164H, Y164Q, and Y164Q/T47M/Y164E/T328N) were constructed and characterized with respect to enzymatic properties and X-ray structural crystal analysis. The optimum pH of the four single mutants shifted to 4.2-4.8, approximately 2 pH units and approximately 1 pH unit lower than those of BCB and SBA, respectively, and their k(cat) values decreased to 41-3% of that of the wild-type enzyme. The X-ray crystal analysis of the enzyme-maltose complexes showed that Glu 367 of the wild type is surrounded by two water molecules (W1 and W2) that are not found in SBA. W1 is hydrogen-bonded to both side chains of Glu 367 and Tyr 164. The mutation of Tyr 164 to Glu and Phe resulted in the disruption of the hydrogen bond between Tyr 164 Oeta and W1 and the introduction of two additional water molecules near position 164. In contrast, the triple mutant of BCB with a slightly decreased pH optimum at pH 6.0 has no water molecules (W1 and W2) around Glu 367. These results suggested that a water-mediated hydrogen bond network (Glu 367...W1...Tyr 164...Thr 328) is the primary requisite for the increased pH optimum of wild-type BCB. This strategy is completely different from that of SBA, in which a hydrogen bond network (Glu 380...Thr 340...Glu 178) reduces the optimum pH in a hydrophobic environment.  相似文献   

17.
Attempts to alter the guanine specificity of ribonuclease T1 (RNase T1) by rational or random mutagenesis have failed so far. The RNase T1 variant RV (Lys41Glu, Tyr42Phe, Asn43Arg, Tyr45Trp, and Glu46Asn) designed by combination of a random and a rational mutagenesis approach, however, exhibits a stronger preference toward adenosine residues than wild-type RNase T1. Steady state kinetics of the cleavage reaction of the two dinucleoside phosphate substrates adenylyl-3',5'-cytidine and guanylyl-3',5'-cytidine revealed that the ApC/GpC ratio of the specificity coefficient (k(cat)/K(m)) was increased approximately 7250-fold compared to that of the wild-type. The crystal structure of the nucleotide-free RV variant has been refined in space group P6(1) to a crystallographic R-factor of 19.9% at 1.7 A resolution. The primary recognition site of the RV variant adopts a similar conformation as already known from crystal structures of RNase T1 not complexed to any nucleotide. Noteworthy is a high flexibility of Trp45 and Asn46 within the three individual molecules in the asymmetric unit. In addition to the kinetic studies, these data indicate the participation of Asn46 in the specific recognition of the base and therefore a specific binding of adenosine.  相似文献   

18.
Binz T  Bade S  Rummel A  Kollewe A  Alves J 《Biochemistry》2002,41(6):1717-1723
The botulinum neurotoxin type A (BoNT/A) light chain (LC) acts as zinc endopeptidase. The X-ray structure of the toxin demonstrated that Zn(2+) is coordinated by His(222) and His(226) of the Zn(2+) binding motif HisGluXXHis and Glu(261), whereas Glu(223) coordinates the water molecule required for hydrolysis as the fourth ligand. Recent analysis of a cocrystal of the BoNT/B LC and its substrate synaptobrevin 2 suggested that Arg(362) and Tyr(365) of the homologous BoNT/A may be directly involved in catalysis. Their role and that of Glu(350) which is also found in the vicinity to the active site were analyzed by site-directed mutagenesis. Various replacements of Arg(362) and substitution of Tyr(365) with Phe resulted in 79- and 34-fold lower k(cat)/K(m) values, respectively. These changes were provoked by decreased catalytic rates (k(cat)) and not by alterations of ground state substrate binding as evidenced by largely unchanged K(d) and K(m) values. None of these mutations affected the overall secondary structure or zinc content of the LC. These findings suggest that the guanidino group of Arg(362) and the hydroxyl group of Tyr(365) together accomplish transition state stabilization as was proposed for thermolysin, being the prototypical member of the gluzincin superfamily of metalloproteases. Mutation of Glu(350) dramatically diminished the hydrolytic activity which must partly be attributed to an altered active site fine structure as demonstrated by an increased sensitivity toward heat-induced denaturing and a lower Zn(2+) binding affinity. Glu(350) apparently occupies a central position in the active site and presumably positions His(222) and Arg(362).  相似文献   

19.
F Ni  Y Konishi  H A Scheraga 《Biochemistry》1990,29(18):4479-4489
The interaction of the C-terminal fragments (residues 52-65 and 55-65) of the thrombin-specific inhibitor hirudin with bovine thrombin was studied by use of one- and two-dimensional NMR techniques in aqueous solution. Thrombin induces specific line broadening of the proton resonances of residues Asp(55) to Gln(65) of the synthetic hirudin fragments H-Asn-Asp-Gly-Asp(55)-Phe-Glu-Glu-Ile-Pro-Glu-Glu-Tyr(63)-Leu-Gln-COOH and acetyl-Asp(55)-Phe-Glu-Glu-Ile-Pro-Glu-Glu-Tyr(63)-Leu-Gln-COOH. This demonstrates that residues 55-65 are the predominant binding site of hirudin fragments with thrombin. Hirudin fragments take on a well-defined structure when bound to thrombin as indicated by several long-range transferred NOEs between the backbone and side-chain protons of the peptides, but they are not structured when free in solution. Particularly, transferred NOEs exist between the alpha CH proton of Glu(61) and the NH proton of Leu(64) [d alpha N(i,i+3)], between the alpha CH proton of Glu(61) and the beta CH2 protons of Leu(64) [d alpha beta(i,i+3)], and between the alpha CH proton of Glu(62) and the gamma CH2 protons of Gln(65) [d alpha gamma(i,i+3)]. These NOEs are characteristic of an alpha-helical structure involving residues Glu(61) to Gln(65). There are also NOEs between the side-chain protons of residues Phe(56), Ile(59), Pro(60), Tyr(63), and Leu(64). Distance geometry calculations suggest that in the structure of the thrombin-bound hirudin peptides all the charged residues lie on the opposite side of a hydrophobic cluster formed by the nonpolar side chains of residues Phe(56), Ile(59), Pro(60), Tyr(63), and Leu(64).  相似文献   

20.
Grant MA  Baikeev RF  Gilbert GE  Rigby AC 《Biochemistry》2004,43(49):15367-15378
The binding of factor IX to cell membranes requires a structured N-terminal omega-loop conformation that exposes hydrophobic residues for a highly regulated interaction with a phospholipid. We hypothesized that a peptide comprised of amino acids Gly4-Gln11 of factor IX (fIX(G4)(-)(Q11)) and constrained by an engineered disulfide bond would assume the native factor IX omega-loop conformation in the absence of Ca(2+). The small size and freedom from aggregation-inducing calcium interactions would make fIX(G4)(-)(Q11) suitable for structural studies for eliciting details about phospholipid interactions. fIX(G4)(-)(Q11) competes with factor IXa for binding sites on phosphatidylserine-containing membranes with a K(i) of 11 microM and inhibits the activation of factor X by the factor VIIIa-IXa complex with a K(i) of 285 microM. The NMR structure of fIX(G4)(-)(Q11) reveals an omega-loop backbone fold and side chain orientation similar to those found in the calcium-bound factor IX Gla domain, FIX(1-47)-Ca(2+). Dicaproylphosphatidylserine (C(6)PS) induces HN, Halpha backbone, and Hbeta chemical shift perturbations at residues Lys5, Leu6, Phe9, and Val10 of fIX(G4)(-)(Q11), while selectively protecting the NHzeta side chain resonance of Lys5 from solvent exchange. NOEs between the aromatic ring protons of Phe9 and specific acyl chain protons of C(6)PS indicate that these phosphatidylserine protons reside 3-6 A from Phe9. Stabilization of the phosphoserine headgroup and glycerol backbone of C(6)PS identifies that phosphatidylserine is in a protected environment that is spatially juxtaposed with fIX(G4)(-)(Q11). Together, these data demonstrate that Lys5, Leu6, Phe9, and Val10 preferentially interact with C(6)PS and allow us to correlate known hemophilia B mutations of factor IX at Lys5 or Phe9 with impaired phosphatidylserine interaction.  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