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 reaction of bovine pancreatic trypsin with human plasma α2-macroglobulin (α2M) was studied at 25°C, using equimolar mixtures of E and I in 50 mM potassium phosphate buffer, pH 7. The conformational change in α2M was monitored through the increase in protein fluorescence at 320 nm (exc λ, 280 nm). At [α2M]0 = [E]0 = 11.5-200 nM, the fluorescence change data fit the integrated second-order rate equation, (F - F0)/(F - F1) = 1 + ki,obsd2M]0t, indicating that cleavage of the bait region in α2M was the rate-determining step.

The apparent rate constant (ki,obsd) was found to be inversely related to reactant concentration. The kinetic behavior of the system was compatible with a model involving reversible, non-bait region binding of E to α2M, competitively limiting the concentration of E available for bait region cleavage. The intrinsic value of ki was (1.7±0.24) × 107 M-l s -1. Kp, the inhibitory constant associated with peripheral binding, was estimated to be in the submicromolar range.

The results of the present study point to a potential problem in interpreting kinetic data relating to protease-induced structural changes in macromolecular substrates. If there is nonproductive binding, as in the case of trypsin and α2M, and the reactions are monitored under pseudo first-order conditions ([S]0 ? [E]0), an intrinsically second-order process (such as the rate-limiting bait region cleavage in α2M) may become kinetically indistinguishable from an intrinsically first-order process (e.g. rate-limiting conformational change). Hence an excess of one component over the other should be avoided in kinetic studies addressing such systems.  相似文献   

2.
Summary The apparent Michaelis constant (K m) of NADH for muscle-type (M4 isozyme) lactate dehydrogenases (LDHs) is highest, at any given temperature of measurement, for LDHs of cold-adapted vertebrates (Table 1). However, these interspecific differences in theK m of NADH are not due to variations in LDH-NADH binding affinity. Rather, theK m differences result entirely from interspecific variation in the substrate turnover constant (k cat) (Fig. 1; Table 2). This follows from the fact that theK m of NADH is equal tok cat divided by the on constant for NADH binding to LDH,k 1, so that interspecific differences ink cat, combined with identical values fork 1 among different LDH reactions, make the magnitude of theK m of NADH a function of substrate turnover number. The temperature dependence of theK m of NADH for a single LDH homologue is the net result of temperature dependence of bothk cat andk 1 (Figs. 3 and 4). Temperature independentK m values can result from simultaneous, and algebraically offsetting, increases ink cat andk 1 with rising temperature. Salt-induced changes in theK m of NADH also may be due to simultaneous perturbation of bothk cat andk 1 (Table 3). These findings are discussed from the standpoint of the evolution of LDH kinetic properties, particularly the interspecific conservation of catalytic and regulatory functions, in differently-adapted species.  相似文献   

3.
The paper presents a kinetic analysis of a reversible enzymatic reaction S⇄P involving two intermediate compounds under the condition [E]0 ≫ [S]0 + [P]0. For the case of mono-exponential behavior, we derive an equation for k obs as a function of [E]0, which emphasizes the pitfalls of oversimplifying kinetic schemes (such as the Michaelis-Menten model) for ribozyme studies. This novel apparent rate constant, which has been arrived at through mechanistic considerations, is analyzed, and the characteristic parameters obtained. The equation, which seems to fit experimental data better than conventional approximations, is used to analyze a single turnover study on an ADC1 ribozyme drawn from hepatitis delta virus RNA. The microscopic kinetic constants for such enzyme are evaluated and its mono-exponential behavior verified.  相似文献   

4.
1. Evolutionary changes in the structure of an enzyme that provide an increase in its Km value are considered. Provided that Km increases as a result of increases in the forward rate constants of the catalysis relative to the reverse rate constants, the enzyme catalyses the conversion of a fixed concentration of its substrate more rapidly when its structure provides that Km>[S] than when Km<[S]. 2. Catalytic efficiency of enzymes is discussed in terms of the simplest plausible model, the Haldane [(1930) Enzymes, Longmans, London] reversible three-step model: [Formula: see text] The rate equation for the forward reaction of this model (formation of P) may be written in the simple form: [Formula: see text] Keq. is the equilibrium constant (=[P]eq./[S]eq.), and kcat.=V/[E]T, where [E]T is the total enzyme concentration. 3. To assess the effectiveness of an enzyme, it is necessary only to determine the extent to which the constraints of a particular kinetic mechanism permit v2 (v when Km»[S]) to approach vd (the diffusion-limited rate). 4. The value of the optimal rate of catalysis (vopt., the maximal value of v2) is dictated by the equilibrium constant for the reaction, Keq.; v2=vd/a, where [Formula: see text] when k+1 is assumed equal to k−3, and vopt.=vd/amin.. When Keq.≥1, it is necessary that k+2»k−1 for a to take its minimum value, amin.; when Keq.«1, it is necessary only that k+2»Keq.·k−1, i.e. a can equal amin. even if k+2<k−1. When Keq.»1, vopt.=vd; when Keq.=1, vopt.=vd/2, and when Keq.«1, vopt.=Keq.·vd. 5. The analysis, together with predicted effects of evolutionary pressure, suggests that in practice the rates of the fastest enzyme-catalysed freely reversible reactions might be expected to be lower than the value of k+1[E]T[S] by about an order of magnitude, particularly if Keq.<1. 6. The existing literature suggests that, in general, appropriate values of Km have evolved for the provision of high rates of catalysis but that many values of kcat. are not large enough to provide optimal rates of catalysis unless the value of k+1 in vivo is lower than its value in free solution.  相似文献   

5.
通过易错PCR提高鼠伤寒沙门氏菌丙氨酸消旋酶催化活性   总被引:2,自引:1,他引:1  
[目的] 通过易错PCR技术提高鼠伤寒沙门氏菌中丙氨酸消旋酶的催化活性。[方法] 利用易错PCR技术构建丙氨酸消旋酶基因alrSt的突变体文库,采用缺陷菌株UT5028筛选突变体基因,以D-氨基酸氧化酶偶联法检测各突变蛋白的活性,通过凝胶过滤层析法分析酶蛋白寡聚化状态,并采用HPLC检测酶蛋白的动力学参数。[结果] 经过易错PCR及定点突变技术最终获得了3个催化活性有所提高的突变体A3V、Y343H和A3VY343H,酶学特性分析发现,与野生型蛋白StAlr相比,突变体Y343H仅对底物L/D-丝氨酸的催化效率略有提高,kcat/Km值分别是StAlr的2.01和3.68倍;而突变体A3V则对底物L/D-丙氨酸或L/D-丝氨酸的Kmkcatkcat/Km值均有较大幅度的改变,其kcat/Km值分别是StAlr的105.51、97.36、4.63和10.73倍。凝胶过滤层析结果显示,突变体A3V在蛋白含量极低时就呈现出单体和二聚体共存状态,且随着蛋白含量的增加,其向二聚体状态迁移的速率最为明显。[结论] 丙氨酸消旋酶StAlr的第3位点是影响其催化活性和低聚合状态的关键位点。  相似文献   

6.
The synthesis of the protected dipeptide BocGlyPheOMe, has been modellised when working in an aqueousorganic biphasic system, with papain as a catalyst. The mathematical model takes into account that one of the substrates, PheOMe, has parallel hydrolysis reactions and that the reaction only takes place in the aqueous phase while the whole reaction system is biphasic. The reaction system has been modellised when working in batch as well as when working in fed-batch mode, achieving a good prediction of the product evolution for both working strategies. When working in fed-batch mode, the extension of the undesired parallel reactions has been diminished, the model has been used for a computer aided optimisation of the addition sequence of PheOMe. The results obtained led to a process operation strategy with a compromise between yield and productivity.List of Symbols [i] concentration of any component i - [i] aq concentration of i in the aqueous phase - [i] bi concentration of i in the biphasic system - [E] 0 initial concentration of enzyme - k e, kq first order kinetic constants - K A, KB equilibrium constants - r m maximum rate of reaction This worked was financed by the Interministerial Commission for Science and Technology (CICYT)from the Spanish Government under projects number BIO/88-370 and SAF92-0261-CO2-02.  相似文献   

7.
Abstract Acetylcholinesterase (AChE) in the susceptible (S) and the resistant (R) strains of housefly (Musca domestica) was investigated using kinetic analysis. The Vmax values of AChE for hydrolyzing acetylthiocholine (ATCh) and butyrylthiocholine (BTCh) were 4578.50 and 1716.08nmol/min/mg* protein in the R strain, and were 1884.75 and 864.72 nmol/min/mg. protein in the Sstrain, respectively. The Vmax ratios of R to S enzyme were 2.43 for ATCh and 1.98 for BTCh. The Km values of AChE for ATCh and BTCh were 0.069 and 0.034 mmol/L in the S strain, and 0.156, 0.059 mmol/L in the R strain, respectively. The Km ratios of R to S enzyme were 2.26 for ATCh and 1.74 for BTCh. The ki ratios of S to R enzyme for three insecticides propoxur, methomyl and paraoxon were 46.04, 4.17 and 2. 86, respectively. In addition, kcat and kcat/Km for measuring turnover and catalytic efficiency of AChE were determined using eserine as titrant. The kcat values of AChE from the R strain for both ATCh and BTCh were higher than those values from the S strain. But the values of kcat/Km were in contrary to the kcat values with R enzyme compared to S enzyme. The AChE catalytic properties and sensitivity to the inhibition by three insecticides in the R and S strains of housefly were discussed based on contribution of Vmax, Km, ki, kcat and kcat/Km. All these data implied that AChE from the R strain might be qualitatively altered. We also observed an intriguing phenomenon that inhibitors could enhance the activity of AChE from the resistant strain. This “flight reaction” of the powerful enzyme might be correlated with the developing resistance of housefly to organophosphate or carbamate insecticides.  相似文献   

8.
The thermal denaturation of the simple, redox-active iron protein rubredoxin is characterized by a slow, irreversible decay of the characteristic red color of the iron center at elevated temperatures in the presence of oxygen at pH 7.8. The denaturation rate is essentially constant and the time period for complete bleaching is nearly independent of protein concentration. These two characteristics of the kinetics can be fit by a simple self-catalyzed kinetics model consisting of the combination of a first-order decay and catalysis by some product of that decay, i.e., dP/dt=k 1[A]+(k 2[P][A])/(K m+[A]), where A is native rubredoxin, P, is unspecified product, k 1 is a first-order rate constant, and k 2 and K m are the catalytic constants. In order for the second term to be of this simple form over the full course of a decay, the model must include the condition that the reaction is effectively irreversible. This model has properties which suggest other biological roles in regulation (changes in k 1 or k 2 can dramatically modulate the kinetics), in timing (titer-independent fixed reaction time), and in self-activation reactions. At one extreme (k1 k2) the kinetics becomes exponential, but at the other extreme (k2 k1) they show a dramatic and rapid terminal increase after a lag period. Some obvious possible roles in the kinetics of programmed cell death, prion disease, and protease autoactivation are discussed.  相似文献   

9.
Tobacco expresses four isomers of assimilatory nitrite reductase (aNiR), leaf‐type (Nii1 and Nii3), and root‐type (Nii2 and Nii4). The high‐resolution crystal structures of Nii3 and Nii4, determined at 1.25 and 2.3 Å resolutions, respectively, revealed that both proteins had very similar structures. The Nii3 structure provided detailed geometries for the [4Fe–4S] cluster and the siroheme prosthetic groups. We have generated two types of Nii3 variants: one set focuses on residue Met175 (Nii3‐M175G, Nii3‐M175E, and Nii3‐M175K), a residue that is located on the substrate entrance pathway; the second set targets residue Gln448 (Nii3‐Q448K), a residue near the prosthetic groups. Comparison of the structures and kinetics of the Nii3 wild‐type (Nii3‐WT) and the Met175 variants showed that the hydrophobic side‐chain of Met175 facilitated enzyme efficiency (kcat/Km). The Nii4‐WT has Lys449 at the equivalent position of Gln448 in Nii3‐WT. The enzyme activity assay revealed that the turnover number (kcat) and Michaelis constant (Km) of Nii4‐WT were lower than those of Nii3‐WT. However, the kcat/Km of Nii4‐WT was about 1.4 times higher than that of Nii3‐WT. A comparison of the kinetics of the Nii3‐Q448K and Nii4‐K449Q variants revealed that the change in kcat/Km was brought about by the difference in Residue 448 (defined as Gln448 in Nii3 and Lys449 in Nii4). By combining detailed crystal structures with enzyme kinetics, we have proposed that Nii3 is the low‐affinity and Nii4 is the high‐affinity aNiR.  相似文献   

10.
Fluorescently labeled derivatives of a xyloglucan (XG) nonasaccharide Glc4Xyl3Gal2 (XLLG) were used as glycosyl acceptors in assays of xyloglucan endotransglycosylase (XET) from germinated nasturtium (Tropaeolum majus) seeds. We have investigated how the type of the oligosaccharide label influences the kinetic parameters of the reaction. The fluorescent probes used to label XLLG were anthranilic acid (AA), 8-aminonaphtalene-1,3,6-trisulfonic acid (ANTS), fluorescein isothiocyanate (FITC), and sulforhodamine (SR), respectively. The obtained data were compared with those of the reactions where aldose and/or alditol forms of tritium-labeled xyloglucan-derived nonasaccharide served as the respective acceptors. Modification at C-1 of the reducing-end glucose in XLLG by substitution with the fluorophore markedly affected the kinetic parameters of the reaction. The Michaelis constants Km for individual acceptors increased in the order [1-3H]XLLG < XLLG-SR < [1-3H]XLLGol < XLLG-FITC < XLLG-ANTS < XLLG-AA, while the turnover numbers characterized by kcat decreased in the order XLLG-FITC > XLLG-SR > XLLG-ANTS > [1-3H]XLLGol > [1-3H]XLLG > XLLG-AA. Catalytic efficiency (expressed as kcat/Km) with XLLG labeled with SR or FITC was 15 and 28 times, respectively, higher than with the tritium-labeled natural substrate [1-3H]XLLG. Comparison of the kinetic parameters found with acceptors labeled with different types of labels enables to select the most effective substrates for the high-throughput assays of XET.  相似文献   

11.
A difficulty associated with high throughput screening for enzyme inhibitors is to establish reaction conditions that maximize the sensitivity and resolution of the assay. Deduction of information from end-point assays at single concentrations requires a detailed understanding of the time progress of the enzymatic reaction, an essential but often difficult process to model. A tool to simulate the time progress of enzyme catalyzed reactions and allows adjustment of reactant concentrations and parameters (initial concentrations, K m, k cat, K i values, enzyme half-life, product•enzyme dissociation constant, and the rate constant for the reversed reaction) has been developed. This tool provides comparison of the progress of uninhibited versus inhibited reactions for common inhibitory mechanisms, and guides the tuning of reaction conditions. Possible applications include: analysis of substrate turnover, identification of the point of maximum difference in product concentration (Δmax[P]) between inhibited and uninhibited reactions, determination of an optimal observation window unbiased for inhibitor mechanisms or potency, and interpretation of observed inhibition in terms of true inhibition. An important observation that can be utilized to improve assay signal strength and resolution is that Δmax[P] occurs at a high degree of substrate consumption (commonly >75%) and that observation close to this point does not adversely affect observed inhibition or IC50 values.  相似文献   

12.
Summary Kinetic data for alternative substrates of recombinant trypanothione reductase fromTrypanosoma cruzi were measured for a series ofN-substituted-L-cysteinylglycyl-3-dimethylaminopropylamides, in which the cysteineN-substituent was either a variant of the benzyloxycarbonyl group or was L-phenylalanine or L-tryptophan. Replacing the benzylic ether oxygen atom by CH2. or NH had relatively minor effects on kcat, but raised the value of Km, 4.5- and 10-fold, respectively. Similarly, relative to the carbobenzoxy group, anN-L-phenylalanyl orN-L-tryptophanyl replacement on the cysteine hardly altered kcat, but increased Km, values by 16.6 and 7.4 fold, respectively. These observations were consistent with the Km, values referring primarily to binding for this series of nonspecific substrates.Abbreviations DCC N,N-dicyclohexylcarbodiimide - dmapa dimethylaminopropylamine - DMF dimethylformamide - GR glutathione reductase - GSSG glutathione disulphide - GSH reduced glutathione - T[S]2 trypanothione disulphide - Hbt hydroxybenzotriazole - TFA trifluoroacetic acid - TLC thin layer chromatography - T[SH]2 reduced trypanothione as dithiol - TR trypanothione reductase - Z.cys.gly.dmapa N-benzyloxycarbonyl-Lcysteinylglycyl-3-dimethylpropylamide  相似文献   

13.
Carbapenem-hydrolyzing β-lactamase from Serratia marcescens FHSM4055 was purified 926-fold by means of carboxylmethyl Sephadex C-50, Sephacryl S-200, and Mono S column chromatography. The molecular weight was 30,000 by SDS-PAGE and the isoelectric point was 8.7. The enzyme activity was inhibited by EDTA, and restored by adding zinc (II) or manganese (II). It was inhibited by p-chloromercuribenzoate and iodine as well as the heavy metals, Hg (II), Fe (II), Fe (III), and Cu (II). These results indicate that the enzyme is a metallo-β-lactamase and that the SH-group of only one cysteine residue probably binds to the metal ion, thus contributing to the stability of the enzyme active center. The specific constant (kcat/Km) showed that the enzyme hydrolyzed various β-lactam antibiotics such as carbapenems, cephalosporins, moxalactam, cephamycins, and penicillins other than monobactams. Ampicillin and piperacillin with respective amino- and imino-groups, ceftazidime with a carboxypropyloxyimino-group, and cefclidin with a carbamoylquinuclidine-group were poor substrates among the β-lactam antibiotics other than the monobactams tested. The plots of the turnover number (kcat) against pH for the hydrolysis of cephaloridine gave an asymmetrical curve with the ‘tail’ on the acid side (pK1, 5.9; pK2, 9.0; pK3, 10.8), whereas those of kcat/Km gave a bell-shaped curve (pK1, 5.8; pK2, 9.8). Both results suggest that two ionic forms of an intermediate yield the same product at different rates and that the enzyme is stable under alkaline conditions. Since the N-terminal amino acid sequence of 27 residues determined was consistent with that of the metalloenzyme (Antimicrob. Agents Chemother., 1994, 38: 71-78), the above enzymatic characteristics seem to coincide.  相似文献   

14.
The three-dimensional structure of the catalytically efficient β-xylosidase from Selenomonas ruminantium in complex with competitive inhibitor 1,3-bis[tris(hydroxymethyl)methylamino]propane (BTP) was determined by using X-ray crystallography (1.3 Å resolution). Most H bonds between inhibitor and protein occur within subsite −1, including one between the carboxyl group of E186 and an N group of BTP. The other N of BTP occupies subsite +1 near K99. E186 (pKa 7.2) serves as catalytic acid. The pH (6-10) profile for is bell-shaped with pKa’s 6.8 and 7.8 on the acidic limb assigned to E186 and inhibitor groups and 9.9 on the basic limb assigned to inhibitor. Mutation K99A eliminates pKa 7.8, strongly suggesting that the BTP monocation binds to the dianionic enzyme D14E186. A sedimentation equilibrium experiment estimates a Kd ([dimer]2/[tetramer]) of 7 × 10−9 M. Similar kcat and kcat/Km values were determined when the tetramer/dimer ratio changes from 0.0028 to 26 suggesting that dimers and tetramers are equally active forms.  相似文献   

15.
The solvent kinetic isotope effects (SKIE) on the yeast α-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 °C. With p-nitrophenyl-d-glucopyranoside (pNPG), the dependence of kcat/Km on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, DOD(kcat/Km), of 1.9 (±0.3). The two pKas characterizing the pH profile were increased in D2O. The shift in pKa2 of 0.6 units is typical of acids of comparable acidity (pKa=6.5), but the increase in pKa1 (=5.7) of 0.1 unit in going from H2O to D2O is unusually small. The initial velocities show substrate inhibition (Kis/Km~200) with a small solvent isotope effect on the inhibition constant [DODKis=1.1 (±0.2)]. The solvent equilibrium isotope effects on the Kis for the competitive inhibitors d-glucose and α-methyl d-glucoside are somewhat higher [DODKi=1.5 (±0.1)]. Methyl glucoside is much less reactive than pNPG, with kcat 230 times lower and kcat/Km 5×104 times lower. The solvent isotope effect on kcat for this substrate [=1.11 (±0. 02)] is lower than that for pNPG [=1.67 (±0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

16.
A -glycosidase of a thermophile, Thermus thermophilus, belonging to the glycoside hydrolase family 1, was cloned and overexpressed in Escherichia coli. The purified enzyme (Ttgly) has a broad substrate specificity towards -D-glucoside, -D-galactoside and -D-fucoside derivatives. The thermostability of Ttgly was exploited to study its kinetic properties within the range 25–80[emsp4 ]°C. Whatever the temperature, except around 60[emsp4 ]°C, the enzyme displayed non-Michaelian kinetic behavior. Ttgly was inhibited by high concentrations of substrate below 60[emsp4 ]°C and was activated by high concentrations of substrate above 60[emsp4 ]°C. The apparent kinetic parameters (k cat and K m ) were calculated at different temperatures. Both k cat and K m increased with an increase in temperature, but up to 75[emsp4 ]°C the values of k cat increased much more rapidly than the values of K m . The observed kinetics might be due to a combination of factors including inhibition by excess substrate and stimulation due to transglycosylation reactions. Our results show that the substrate could act not only as a glycosyl donor but also as a glycosyl acceptor. In addition, when the glucose was added to reaction mixtures, inhibition or activation was observed depending on both substrate concentration and temperature. A reaction model is proposed to explain the kinetic behavior of Ttgly. The scheme integrates the inhibition observed at high concentrations of substrate and the activation due to transglycosylation reactions implicating the existence of a transfer subsite.  相似文献   

17.
Using site-directed mutagenesis and steady-state kinetic measurements, the functional role of the conserved glycine 127 in a human vaccinia H1-related phosphatase (VHR) was investigated. The mutations of Gly127 to Ala and Pro resulted in a significant decrease in k cat/K m, and increase in K i for arsenate, indicating that flexibility at the Gly127 site has a large effect on substrate binding and catalytic activity. No substantial decrease in k cat/K m and increase in K i values were observed for G127 deletion mutant. This showed the conformational flexibility of the PTP loop also affected the enzymatic activity of VHR. Our data suggest that the flexibility of the PTP loop in VHR is probably controlled by Gly127, and that even subtle changes in the loop flexibility may interfere with substrate binding and enzymatic reaction.  相似文献   

18.
The hydrolysis reaction of N α-benzoyl-L-arginine ethyl ester catalyzed by trypsin from pig pancreas was comparatively studied in an aqueous buffer solution and in the system of reversed micelles of Aerosol OT in octane (pH 8.5) to determine the mechanisms of influence of the enzyme microenvironment on the rate constants of the elementary stages of the enzymatic reaction. The temperature dependences of the catalytic constant k cat and the rate constant of the second order k cat/K m (s, catalysis efficiency) allowed the determination of the rate constants and the activation energy of elementary stages of the enzymatic reaction. It was revealed that a decrease in the efficiency of catalytic action of trypsin in reverse micelles in comparison with an aqueous solution is first of all determined by a decrease in the rate constant of formation of the enzyme-substrate complex k 1. Possible mechanisms of the effect of the microenvironment on the elementary stages of catalytic action of the enzyme are discussed.  相似文献   

19.
Acetylcholinesterase (AChE) was investigated inWalterinnesia aegyptia venom and characterized with respect to its kinetic properties. It was found that 4.0 ug of crude venom protein and an incubation time of 4.0 min were suitable conditions for linearity of AChE activity at 25°C. The optimum strength of the sodium phosphate buffer was 0.05 M, and the optimum pH was 7.75. The optimum temperature was 30°C. The activation energy and the heat of activation were observed to be 6510 and 5922 cal/mole. The AChE was specific for acetylthiocholine but it did not hydrolyse butyrylthiocholine. The optimum substrate concentration was 3.0 mM but at higher substrate concentrations, the AChE activity declined. The ASCh concentration ranges for different orders of the reactions were determined and kinetic parameters (Km, Vmax, kcat, and ksp) were established at each order of the reaction.Abbreviations AChE acetylcholinesterase - ASCh acetylthiocholine - Km Michaelis-Menten constant - Vmax the limiting maximal velocity - AChEa acylated enzyme - kcat turnover number - ksp specificity constant  相似文献   

20.
Ribavirin is a broad-spectrum antiviral drug and can be produced by enzymatic synthesis by purine nucleoside phosphorylase (PNP). In this study, we describe the application of such a cold-adapted XmPNP in ribavirin bioconversion which showed approximately 15°C lower optimum temperature and 1.80-fold higher catalytic efficiency (kcat/Km) at 37°C within substrate inosine than homolog in E. coli. By contrast, E. coli (XmPNP) took only 12 h to reach maximum substrate conversion rate (70%) under its optimum temperature (50°C) by using recombinant strain cell as enzyme source, but E. coli (EcPNP) did at 24 h. These results suggest cold-adapted PNP is one attractive candidate for ribavirin bioconversion and other nucleoside medications to improve the catalytic efficiency.  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