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.
Human serum albumin (HSA) or anti-human serum albumin (anti-HSA) yields a catalytic hydrogen wave at about -1.85V (vs Ag/AgCl) in 0.25M NH(3).H(2)O-NH(4)Cl (pH 8.58) buffer. When 1.0 x 10(-2)M K(2)S(2)O(8) is present, the catalytic hydrogen wave is further catalyzed, producing a parallel catalytic wave of hydrogen as catalyst in nature, termed the parallel catalytic hydrogen wave. The sensitivity of the parallel catalytic hydrogen wave is higher by two orders of magnitude than that of the catalytic hydrogen wave. Using the parallel catalytic hydrogen wave of anti-HSA or HSA in the presence of K(2)S(2)O(8), two sensitive methods for the determination of anti-HSA were developed. One is a direct determination based on the parallel catalytic hydrogen wave of anti-HAS itself, and the other is a homogeneous immunoassay based on measuring the decrease of the peak current of the parallel catalytic hydrogen wave of HSA after homogeneous immunoreaction of HSA with anti-HSA. In the direct determination, the second-order derivative peak current of the parallel catalytic hydrogen wave of anti-HSA itself is rectilinear to its titer in the range from 1:1.0 x 10(7) to 1:8.4 x 10(6). In the homogeneous immunoassay, the decrease in the second-order derivative peak current of the parallel catalytic hydrogen wave of HSA is linearly related to the added anti-HSA in the titer range from 1:3.0 x 10(7) to 1:6.0 x 10(6). These assays are highly sensitive and rapid in operation and can be used to evaluate such antigens and their antibodies as those that could yield the parallel catalytic hydrogen wave.  相似文献   

2.
Pepstatin-insensitive carboxyl proteinases from Pseudomonas sp. (PCP) and Xanthomonas sp. (XCP) have no conserved catalytic residue sequences, -Asp*-Thr-Gly- (Asp is the catalytic residue) for aspartic proteinases. To identify the catalytic residues of PCP and XCP, we selected presumed catalytic residues based on their high sequence similarity, assuming that such significant sites as catalytic residues will be generally conserved. Several Ala mutants of Asp or Glu residues were constructed and analyzed. The D170A, E222A, and D328A mutants for PCP and XD79A, XD169A, and XD348A mutants for XCP were not converted to mature protein after activation, and no catalytic activity could be detected in these mutants. The specificity constants toward chromogenic substrate of the other PCP and XCP mutants, except for the D84A mutant of PCP, were similar to that of wild-type PCP or XCP. Coupled with the result of chemical modification (Ito, M., Narutaki, S., Uchida, K., and Oda, K. (1999) J. Biochem. (Tokyo) 125, 210-216), a pair of Asp residues (170 and 328) for PCP and a pair of Asp residues (169 and 348) for XCP were elucidated to be their catalytic residues, respectively. The Glu(222) residue in PCP or Asp(79) residue in XCP was excluded from the candidates as catalytic residues, since the corresponding mutant retained its original activity.  相似文献   

3.
Some preparations of both native aspartate transcarbamylase from Escherichia coli and catalytic subunit have fewer tight binding sites per oligomer for carbamyl-P than the number of catalytic peptide chains. In contrast, the number of sites for the tight-binding inhibitor N-(phosphonacetyl)-L-aspartate does equal the number of catalytic chains in each case. Binding of the labile carbamyl-P was determined using rapid gel filtration, with conversion to stable carbamyl-L-aspartate during collection. Native enzyme (six catalytic chains) obtained from cells grown under the conditions of J.C. Gerhart and H. Holoubek (J. Biol. Chem. (1967) 242, 2886-2892) has 5.4 tight sites for carbamyl-P at pH 8.0 (KD = 9.9 muM), whereas native enzyme from cells grown with higher concentrations of glucose, uracil, and histidine (to yield more enzyme per unit volume of culture) has only 1.9 tight sites at pH 8.0 (KD = 4.6 muM) and only 2.3 tight sites at pH 7.0 (KD = 2.6 muM). At pH 8.0, catalytic subunit (three catalytic chains) obtained from the former native enzyme has 2.2 tight sites for carbamyl-P (KD = 2.4 muM) and the number of sites is 2.3 in the presence of 35 mM succinate, whereas catalytic subunit obtained from the latter native enzyme has 1.8 tight sites (KD = 3.6 muM) in the absence of succinate and 2.3 tight sites in its presence. The number of tight binding sites is also less than the number of subunit peptide chains in 19F nuclear magnetic resonance experiments performed with catalytic subunit and two fluorinated analogs of carbamyl-P at comparable concentrations of analogs and active sites. A model is proposed in which incomplete removal of formylmethionine from the NH2 termini of the enzyme under conditions of extreme depression affects affinity for ligands.  相似文献   

4.
In mammals, eight aminoacyl-tRNA synthetases (AARSs) and three AARS-interacting multifunctional proteins (AIMPs) form a multi-tRNA synthetase complex (MSC). MSC components possess extension peptides for MSC assembly and specific functions. Human cytosolic methionyl-tRNA synthetase (MRS) has appended peptides at both termini of the catalytic main body. The N-terminal extension includes a glutathione transferase (GST) domain responsible for interacting with AIMP3, and a long linker peptide between the GST and catalytic domains. Herein, we determined crystal structures of the human MRS catalytic main body, and the complex of the GST domain and AIMP3. The structures reveal human-specific structural details of the MRS, and provide a dynamic model for MRS at the level of domain orientation. A movement of zinc knuckles inserted in the catalytic domain is required for MRS catalytic activity. Depending on the position of the GST domain relative to the catalytic main body, MRS can either block or present its tRNA binding site. Since MRS is part of a huge MSC, we propose a dynamic switching between two possible MRS conformations; a closed conformation in which the catalytic domain is compactly attached to the MSC, and an open conformation with a free catalytic domain dissociated from other MSC components.  相似文献   

5.
Konstantinov AA 《FEBS letters》2012,586(5):630-639
Several issues relevant to the current studies of cytochrome c oxidase catalytic mechanism are discussed. The following points are raised. (1) The terminology currently used to describe the catalytic cycle of cytochrome oxidase is outdated and rather confusing. Presumably, it would be revised so as to share nomenclature of the intermediates with other oxygen-reactive heme enzymes like P450 or peroxidases. (2) A "catalytic cycle" of cytochrome oxidase involving complete reduction of the enzyme by 4 electrons followed by oxidation by O(2) is a chimera composed artificially from two partial reactions, reductive and oxidative phases, that never operate together as a true multi-turnover catalytic cycle. The 4e(-) reduction-oxidation cycle would not serve a paradigm for oxygen reduction mechanism and protonmotive function of cytochrome oxidase. (3) The foremost role of the K-proton channel in the catalytic cycle may consist in securing faultless delivery of protons for heterolytic O-O bond cleavage in the oxygen-reducing site, minimizing the danger of homolytic scission reaction route. (4) Protonmotive mechanism of cytochrome oxidase may vary notably for the different single-electron steps in the catalytic cycle.  相似文献   

6.
Enterokinase (EC 3.4.21.9) is a serine proteinase in the duodenum that exhibits specificity for the sequence (Asp)(4)-Lys. It converts trypsinogen to trypsin. Its high specificity for the recognition site makes enterokinase (EK) a useful tool for in vitro cleavage of fusion proteins. cDNA encoding the catalytic chain of Chinese bovine enterokinase was cloned and its encoding amino acid sequence is identical to the previously reported sequence although there are two one-base mutations which do not change the encoded amino acid. The EK catalytic subunit cDNA was cloned into plasmid pET32a, and fused downstream to the fusion partner thioredoxin (Trx) and the following DDDDK enterokinase recognition sequence. The recombinant bovine enterokinase catalytic subunit was expressed in Escherichia coli BL21(DE3), and most products existed in soluble form. After an in vivo autocatalytic cleavage of the recombinant Trx-EK catalytic domain fusion protein, intact, biologically active EK catalytic subunit was released from the fusion protein. The recombinant intact EK catalytic subunit was purified to homogeneity with a specific activity of 720 AUs/mg protein through ammonium sulfate precipitation, DEAE chromatography, and gel filtration. The purified intact EK catalytic subunit has a K(m) of 0.17 mM, and K(cat) is 20.8s(-1). From 100 ml flask culture, 4.3 mg pure active EK catalytic subunits were obtained.  相似文献   

7.
Chlorophyllases (Chlases), cloned so far, contain a lipase motif with the active serine residue of the catalytic triad of triglyceride lipases. Inhibitors specific for the catalytic serine residue in serine hydrolases, which include lipases effectively inhibited the activity of the recombinant Chenopodium album Chlase (CaCLH). From this evidence we assumed that the catalytic mechanism of hydrolysis by Chlase might be similar to those of serine hydrolases that have a catalytic triad composed of serine, histidine and aspartic acid in their active site. Thus, we introduced mutations into the putative catalytic residue (Ser162) and conserved amino acid residues (histidine, aspartic acid and cysteine) to generate recombinant CaCLH mutants. The three amino acid residues (Ser162, Asp191 and His262) essential for Chlase activity were identified. These results indicate that Chlase is a serine hydrolase and, by analogy with a plausible catalytic mechanism of serine hydrolases, we proposed a mechanism for hydrolysis catalyzed by Chlase.  相似文献   

8.
The mechanism of inhibition of yeast mitochondrial F(1)-ATPase by its natural regulatory peptide, IF1, was investigated by correlating the rate of inhibition by IF1 with the nucleotide occupancy of the catalytic sites. Nucleotide occupancy of the catalytic sites was probed by fluorescence quenching of a tryptophan, which was engineered in the catalytic site (beta-Y345W). Fluorescence quenching of a beta-Trp(345) indicates that the binding of MgADP to F(1) can be described as 3 binding sites with dissociation constants of K(d)(1) = 10 +/- 2 nm, K(d2) = 0.22 +/- 0.03 microm, and K(d3) = 16.3 +/- 0.2 microm. In addition, the ATPase activity of the beta-Trp(345) enzyme followed simple Michaelis-Menten kinetics with a corresponding K(m) of 55 microm. Values for the K(d) for MgATP were estimated and indicate that the K(m) (55 microm) for ATP hydrolysis corresponds to filling the third catalytic site on F(1). IF1 binds very slowly to F(1)-ATPase depleted of nucleotides and under unisite conditions. The rate of inhibition by IF1 increased with increasing concentration of MgATP to about 50 mum, but decreased thereafter. The rate of inhibition was half-maximal at 5 microm MgATP, which is 10-fold lower than the K(m) for ATPase. The variations of the rate of IF1 binding are related to changes in the conformation of the IF1 binding site during the catalytic reaction cycle of ATP hydrolysis. A model is proposed that suggests that IF1 binds rapidly, but loosely to F(1) with two or three catalytic sites filled, and is then locked in the enzyme during catalytic hydrolysis of ATP.  相似文献   

9.
10.
Limited trypsin digestion of type I cAMP-dependent protein kinase holoenzyme results in a proteolytic-resistant Delta(1-72) regulatory subunit core, indicating that interaction between the regulatory and catalytic subunits extends beyond the autoinhibitory site in the R subunit at the NH(2) terminus. Sequence alignment of the two R subunit isoforms, RI and RII, reveals a significantly sequence diversity at this specific region. To determine whether this sequence diversity is functionally important for interaction with the catalytic subunit, specific mutations, R133A and D328A, are introduced into sites adjacent to the active site cleft in the catalytic subunit. While replacing Arg(133) with Ala decreases binding affinity for RII, interaction between the catalytic subunit and RI is not affected. In contrast, mutant C(D328A) showed a decrease in affinity for binding RI while maintaining similar affinities for RII as compared with the wild-type catalytic subunit. These results suggest that sequence immediately NH(2)-terminal to the consensus inhibition site in RI and RII interacts with different sites at the proximal region of the active site cleft in the catalytic subunit. These isoform-specific differences would dictate a significantly different domain organization in the type I and type II holoenzymes.  相似文献   

11.
In aspartate transcarbamylase (ATCase) each regulatory chain interacts with two catalytic chains each one belonging to a different trimeric catalytic subunit (R1-C1 and R1-C4 types of interactions as defined in Fig. 1). In order to investigate the interchain contacts that are involved in the co-operative interactions between the catalytic sites, a series of modified forms of the enzyme was prepared by site-directed mutagenesis. The amino acid replacements were devised on the basis of the previously described properties of an altered form of ATCase (pAR5-ATCase) which lacks the homotropic co-operative interactions between the catalytic sites. The results obtained (enzyme kinetics, bisubstrate analog influence and pH studies) show that the R1-C4 interaction is essential for the establishment of the enzyme conformation that has a low affinity for aspartate (T state), and consequently for the existence of co-operativity between the catalytic sites. This interaction involves the 236-250 region of the aspartate binding domain of the catalytic chain (240s loop) and the 143-149 region of the regulatory chain which comprises helix H3'.  相似文献   

12.
The presence of tissue factor is essential for factor VIIa (FVIIa) to reach its full catalytic potential. The previous work in this laboratory demonstrated that substitution of the EGF1 domain of factor VIIa with that of factor IX (FVII((IXegf1))a) results in a substantial decrease in TF-binding affinity and catalytic activity. Supporting simulations of the solution structures of Ca(2+)-bound factor VIIa and FVII((IXegf1))a with tissue factor are provided. Mutants are generated, based on the simulation model, to study the effect of EGF1 substitution on catalytic activity. The simulations show larger Gla-EGF1 and EGF1-EGF2 inter-domain motions for FVII((IXegf1))a than for factor VIIa. The catalytic domain of the chimeric factor VIIa has been disturbed and several surface loops in the catalytic domain of FVII((IXegf1))a (Loop 170s (170-182), Loop 1 (185-188) and Loop 2 (221A-225)) manifest larger position fluctuations than wild-type. The position of Loop 140s (142-152) of FVII((IXegf1))a, near the N terminus insertion site of the catalytic domain, shifts relative to factor VIIa, resulting in a slight alteration of the active site. The results suggest that these four loops mediate the effect of the EGF1 domain substitution on the S1 site and catalytic residues. To test the model, we prepared mutations of these surface loops, including four FVII mutants, D186A, K188A, L144A and R147A, a FVII mutant with multiple mutations (MM3: L144A+R147A+D186A) and a FVII mutant with Loop 170s partially deleted, Loop 170s(del). The catalytic activities towards a small peptidyl substrate decreased 2.4, 4.5 and 9-fold for Loop 170s(del)a (a, activated), L144Aa and D186Aa, respectively, while MM3a lost almost all catalytic activity. The combined results of the simulations and mutants provide insight into the mechanism by which tissue factor enhances factor VIIa catalytic activity.  相似文献   

13.
Aspartate transcarbamoylase undergoes a domain closure in the catalytic chains upon binding of the substrates that initiates the allosteric transition. Interdomain bridging interactions between Glu(50) and both Arg(167) and Arg(234) have been shown to be critical for stabilization of the R state. A hybrid version of the enzyme has been generated in vitro containing one wild-type catalytic subunit, one catalytic subunit in which Glu(50) in each catalytic chain has been replaced by Ala (E50A), and wild-type regulatory subunits. Thus, the hybrid enzyme has one catalytic subunit capable of domain closure and one catalytic subunit incapable of domain closure. The hybrid does not behave as a simple mixture of the constituent subunits; it exhibits lower catalytic activity and higher aspartate affinity than would be expected. As opposed to the wild-type enzyme, the hybrid is inhibited allosterically by CTP at saturating substrate concentrations. As opposed to the E50A holoenzyme, the hybrid is not allosterically activated by ATP at saturating substrate concentrations. Small angle x-ray scattering showed that three of the six interdomain bridging interactions in the hybrid is sufficient to cause the global structural change to the R state, establishing the critical nature of these interactions for the allosteric transition of aspartate transcarbamoylase.  相似文献   

14.
In Malaysia, there has been interest in the utilization of palm oil and oil palm biomass for the production of environmental friendly biofuels. A biorefinery based on palm oil and oil palm biomass for the production of biofuels has been proposed. The catalytic technology plays major role in the different processing stages in a biorefinery for the production of liquid as well as gaseous biofuels. There are number of challenges to find suitable catalytic technology to be used in a typical biorefinery. These challenges include (1) economic barriers, (2) catalysts that facilitate highly selective conversion of substrate to desired products and (3) the issues related to design, operation and control of catalytic reactor. Therefore, the catalytic technology is one of the critical factors that control the successful operation of biorefinery. There are number of catalytic processes in a biorefinery which convert the renewable feedstocks into the desired biofuels. These include biodiesel production from palm oil, catalytic cracking of palm oil for the production of biofuels, the production of hydrogen as well as syngas from biomass gasification, Fischer-Tropsch synthesis (FTS) for the conversion of syngas into liquid fuels and upgrading of liquid/gas fuels obtained from liquefaction/pyrolysis of biomass. The selection of catalysts for these processes is essential in determining the product distribution (olefins, paraffins and oxygenated products). The integration of catalytic technology with compatible separation processes is a key challenge for biorefinery operation from the economic point of view. This paper focuses on different types of catalysts and their role in the catalytic processes for the production of biofuels in a typical palm oil and oil palm biomass-based biorefinery.  相似文献   

15.
Phosphodiesterases (PDEs) comprise a superfamily of phosphohydrolases that degrade 3',5'-cyclic nucleotides. All known mammalian PDEs are dimeric, but the functional significance of dimerization is unknown. A deletion mutant of cGMP-binding cGMP-specific PDE (PDE5), encoding the 357 carboxyl-terminal amino acids including the catalytic domain, has been generated, expressed, and purified. The K(m) of the catalytic fragment for cGMP (5.5 +/- 0. 51 microM) compares well with those of the native bovine lung PDE5 (5.6 microM) and full-length wild type recombinant PDE5 (2 +/- 0.4 microM). The catalytic fragment and full-length PDE5 have similar IC(50) values for the inhibitors 3-isobutyl-1-methylxanthine (20 microM) and sildenafil (Viagra(TM))(4 nM). Based on measured values for Stokes radius (29 A) and sedimentation coefficient (2.9 S), the PDE5 catalytic fragment has a calculated molecular mass of 35 kDa, which agrees well with that predicted by amino acid content (43.3 kDa) and with that estimated using SDS-polyacrylamide gel electrophoresis (39 kDa). The combined data indicate that the recombinant PDE5 catalytic fragment is monomeric, and retains the essential catalytic features of the dimeric, full-length enzyme. Therefore, the catalytic activity of PDE5 holoenzyme requires neither interaction between the catalytic and regulatory domains nor interactions between subunits of the dimer.  相似文献   

16.
【目的】阐明嗜热细菌Clostridium thermocellum Xyn Z蛋白的阿魏酸酯酶催化域的酶学特性,为其在生物质能源及其它发酵工业中的应用奠定基础。【方法】分别构建了C.thermocellum Xyn Z的阿魏酸酯酶催化域(FAE)及该阿魏酸酯酶催化域和碳水化合物结合域(FAE-CBM6)编码基因的原核表达载体,并在大肠杆菌菌株BL21(DE3)中异源表达,在此基础上分析比较了温度、pH、底物、金属离子及CBM6结合域对阿魏酸酯酶活性的影响。【结果】重组FAE酶及FAE-CBM6酶发挥催化活性的适宜pH值为5.0-9.0,适宜温度为50-70°C,它们对不同金属离子的响应有差异。【结论】在同一反应条件下,FAE-CBM6酶的酶活均比FAE高,说明CBM6结合域的存在对于阿魏酸酯酶活性有促进作用。  相似文献   

17.
Huang X  Liu Y  Liang K  Tang Y  Liu J 《Biomacromolecules》2008,9(5):1467-1473
A new nanoenzyme model with glutathione peroxidase-like active site was constructed on polystyrene nanoparticle (PN1) via microemulsion polymerization. In this model system, two functional monomers were designed: one is a tellurium-containing compound that was introduced on the surface of the nanoparticle and acts as a catalytic center, and the other one is an arginine-containing compound designed as a binding site for the complexation of the carboxyl group of substrate 3-carboxy-4-nitrobenzenethiol (ArSH, 1). As a new glutathione peroxidase (GPx) mimic, it demonstrated excellent catalytic activity and substrate specificity. In ArSH assay system, it was at least 316,000-fold more efficient than PhSeSePh for the reduction of cumene hydroperoxide (CUOOH) by ArSH. In contrast to model PN2, which lacks of substrate binding site, PN1 exhibits an obvious enhancement in catalytic activity. To further promote the catalytic efficiency, a substrate ArSH surface-imprinted nanoenzyme model (I-PN) was developed. By correctly incorporating and positioning the catalytic center tellurium and functional binding factor guanidinium, a continuative activity enhancement of 596,000-fold for the reduction of CUOOH by catalyst I-PN compared with diphenyl diselenide (PhSeSePh) was observed. The results clearly show that polymeric nanoparticle can be developed as an excellent model for combining most of catalytic factors of enzyme into one scaffold.  相似文献   

18.
C-C hydrolase MhpC (2-hydroxy-6-keto-nona-1,9-dioic acid 5,6-hydrolase) from Escherichia coli catalyses the hydrolytic C-C cleavage of the meta-ring fission product on the phenylpropionic acid catabolic pathway. The crystal structure of E. coli MhpC has revealed a number of active-site amino acid residues that may participate in catalysis. Site-directed mutants of His263, Ser110, His114, and Ser40 have been analysed using steady-state and stopped-flow kinetics. Mutants H263A, S110A and S110G show 10(4)-fold reduced catalytic efficiency, but still retain catalytic activity for C-C cleavage. Two distinct steps are observed by stopped-flow UV/Vis spectrophotometry, corresponding to ketonisation and C-C cleavage: H263A exhibits very slow ketonisation and C-C cleavage, whereas S110A and S110G exhibit fast ketonisation, an intermediate phase, and slow C-C cleavage. H114A shows only twofold-reduced catalytic efficiency, ruling out a catalytic role, but shows a fivefold-reduced K(M) for the natural substrate, and an ability to process an aryl-containing substrate, implying a role for His114 in positioning of the substrate. S40A shows only twofold-reduced catalytic efficiency, but shows a very fast (500 s(-1)) interconversion of dienol (317 nm) to dienolate (394 nm) forms of the substrate, indicating that the enzyme accepts the dienol form of the substrate. These data imply that His263 is responsible for both ketonisation of the substrate and for deprotonation of water for C-C cleavage, a novel catalytic role in a serine hydrolase. Ser110 has an important but non-essential role in catalysis, which appears not to be to act as a nucleophile. A catalytic mechanism is proposed involving stabilisation of reactive intermediates and activation of a nucleophilic water molecule by Ser110.  相似文献   

19.
20.
Although belonging to the widely investigated peroxidase superfamily, lactoperoxidase (LPO) and myeloperoxidase (MPO) share structural and functional features that make them peculiar with respect to other enzymes of the same group. A survey of the available literature on their catalytic intermediates enabled us to ask some questions that remained unanswered. These questions concern controversial features of the LPO and MPO catalytic cycle, such as the existence of Compound I and Compound II isomers and the identification of their spectroscopic properties. After addressing each of these questions, we formulated a hypothesis that describes an integrated vision of the catalytic mechanism of both enzymes. The main points are: (a) a re-evaluation of the role of superoxide as a reductant in the catalytic cycle; (b) the existence of Cpd I isomers; (c) reciprocal interactions between catalytic intermediates and (d) a mechanistic explanation for catalase activity in both enzymes.  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