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.
Phytopathogenic fungi penetrate plants by breaking down the cuticular barrier with cutinase. Cutinases are extracellular hydrolytic enzymes that degrade cutin, a polyester composed of hydroxy and epoxy fatty acids. Until now, cutinase has been recognized by its ability to release labeled cutin monomers or by a non-specific esterase assay based on the hydrolysis of p-nitrophenyl esters of short fatty acids. In this work, an insoluble p-nitrophenyl derivative was synthesized and purified, and its structure was determined to be 4-nitrophenyl (16-methyl sulfone ester) hexadecanoate (pNMSEH) by nuclear magnetic resonance (H+ NMR) analysis. pNMSEH was tested as a new cutinase substrate with Pseudomonas mandocino cutinase and porcine liver esterase. While a linear release over time of p-nitrophenol (pNP) was recorded in the presence of cutinase, no response was obtained with the esterase. The calculated kinetic parameters of pNMSEH hydrolysis by cutinase revealed a high specificity (Km=1.8mM), albeit a low catalytic rate (Vmax=10.5 micromol min(-l)l(-1)). This new synthetic substrate may be helpful for detecting and assaying cutinase activity in mixed solutions, such as crude fungal extracellular extracts.  相似文献   

2.
Cytoplasmic soluble proteins from ungerminated conidia of Botrytis cinerea exhibited cutinase activity, while cell wall binding proteins lacked this activity. Cutinase activity in proteins extracted from cell walls and cytoplasm of ungerminated conidia of Botrytis cinerea was determined using p-nitrophenyl butyrate (PNB) and TLC analysis of products derived from hydrolysis of [3H]cutin. Treatment of conidia with indoxyl acetate, a substrate indicative of non-specific esterase and cutinase activity, also gave a positive reaction in the cytoplasm of ungerminated conidia. The possible role of a putative constitutive cutinase in the cytoplasm of conidia in the early stages of infection of plants by B. cinerea is discussed.  相似文献   

3.
The reactivity, stability and unfolding of wild-type (WT) Fusarium solani pisi cutinase and L153Q, S54D and T179C variants were studied in the absence and presence of the dioctyl sulfosuccinate sodium salt (AOT) surfactant. In the absence of surfactant the S54D variant catalytic activity is similar to that of the WT cutinase, whereas L153Q and T179C variants show a lower activity. AOT addition induces an activity reduction for WT cutinase and its variants, although for low AOT concentrations a small increase of activity was observed for S54D and T179C. The enzyme deactivation in the presence of 0.5 mM AOT is relatively slow for the S54D and T179C variants when compared to wild-type cutinase and L153Q variant. These results were correlated with secondary and tertiary structure changes assessed by the CD spectrum and fluorescence of the single tryptophan and the six tyrosine residues. The WT cutinase and S54D variant have similar secondary and tertiary structures that differ from those of T179C and L153Q variants. L153Q, S54D and T179C mutations prevent the formation of hydrophobic crevices responsible for the unfolding by anionic surfactants, with the consequent decrease of the AOT-cutinase interactions.  相似文献   

4.
We studied the reaction between vinyl butyrate and 2-phenyl-1-propanol in acetonitrile catalyzed by Fusarium solani pisi cutinase immobilized on zeolites NaA and NaY and on Accurel PA-6. The choice of 2-phenyl-1-propanol was based on modeling studies that suggested moderate cutinase enantioselectivity towards this substrate. With all the supports, initial rates of transesterification were higher at a water activity (a(w)) of 0.2 than at a(w) = 0.7, and the reverse was true for initial rates of hydrolysis. By providing acid-base control in the medium through the use of solid-state buffers that control the parameter pH-pNa, which we monitored using an organo-soluble chromoionophoric indicator, we were able, in some cases, to completely eliminate dissolved butyric acid. However, none of the buffers used were able to improve the rates of transesterification relative to the blanks (no added buffer) when the enzyme was immobilized at an optimum pH of 8.5. When the enzyme was immobilized at pH 5 and exhibited only marginal activity, however, even a relatively acidic buffer with a pK(a) of 4.3 was able to restore catalytic activity to about 20% of that displayed for a pH of immobilization of 8.5, at otherwise identical conditions. As a(w) was increased from 0.2 to 0.7, rates of transesterification first increased slightly and then decreased. Rates of hydrolysis showed a steady increase in that a(w) range, and so did total initial reaction rates. The presence or absence of the buffers did not impact on the competition between transesterification and hydrolysis, regardless of whether the butyric acid formed remained as such in the reaction medium or was eliminated from the microenvironment of the enzyme through conversion into an insoluble salt. Cutinase enantioselectivity towards 2-phenyl-1-propanol was indeed low and was not affected by differences in immobilization support, enzyme protonation state, or a(w).  相似文献   

5.
The backbone dynamics of Fusarium solani pisi cutinase has been studied by a variety of nuclear magnetic resonance experiments to probe internal motions on different time scales. The core of cutinase appears to be highly rigid. The binding site, including the oxyanion hole, is mobile on the microsecond to millisecond time scale, in contrast to the well-defined active site and preformed oxyanion hole elucidated by X-ray crystallography [Martinez, C., de Geus, P., Lauwereys, M., Matthyssens, G., and Cambillau, C. (1992) Nature 356, 615-618]. In this crystal structure, cutinase has a rather open conformation, in which the hydrophobic binding site is exposed. The observed mobility in solution most likely represents the interconversion between open and more closed conformations, like in a true lipase. The opening and closing motions are on a time scale which corresponds with the kinetics of the hydrolysis reaction, i.e., the millisecond range, which suggests that these conformational rearrangements form the rate-limiting step in catalysis. We conclude that the crystal structure probably represents one of the multiple conformations present in solution, which fortuitously is the active conformation. The implications of our findings are discussed with particular reference to the explanation of the lack of interfacial activation as found for cutinase.  相似文献   

6.
Xanthine dehydrogenase (XDH) from the bacterium Rhodobacter capsulatus catalyzes the hydroxylation of xanthine to uric acid with NAD+ as the electron acceptor. R. capsulatus XDH forms an (alphabeta)2 heterotetramer and is highly homologous to homodimeric eukaryotic xanthine oxidoreductases. Here we first describe reductive titration and steady state kinetics on recombinant wild-type R. capsulatus XDH purified from Escherichia coli, and we then proceed to evaluate the catalytic importance of the active site residues Glu-232 and Glu-730. The steady state and rapid reaction kinetics of an E232A variant exhibited a significant decrease in both kcat and kred as well as increased Km and Kd values as compared with the wild-type protein. No activity was determined for the E730A, E730Q, E730R, and E730D variants in either the steady state or rapid reaction experiments, indicating at least a 10(7) decrease in catalytic effectiveness for this variant. This result is fully consistent with the proposed role of this residue as an active site base that initiates catalysis.  相似文献   

7.
A bacterial cutinase from Thermobifida fusca, named Tfu_0883, was genetically modified by site-directed mutagenesis to enhance its activity on poly(ethylene terephthalate) (PET). The new mutations tailored the catalytic site for PET, increasing the affinity of cutinase to this hydrophobic substrate and the ability to hydrolyze it. The mutation I218A was designed to create space and the double mutation Q132A/T101A was designed both to create space and to increase hydrophobicity. The activity of the double mutant on the soluble substrate p-nitrophenyl butyrate increased two-fold compared to wild-type cutinase, while on PET both single and double mutants exhibited considerably higher hydrolysis efficiency. The replacement of specific amino acids at the active site was an effective approach for the improvement of the Tfu_0883 cutinase capacity to hydrolyze polyester surfaces. Thus, this study provides valuable insight on how the function and stability of enzymes can be improved by molecular engineering for their application in synthetic fiber biotransformation.  相似文献   

8.
The activity of various lipases was compared, in both free and immobilized forms, using the kinetics of the hydrolysis reaction of p-nitrophenyl butyrate, which was followed with in situ UV/Vis diode array spectrophotometry. Several enzymes were used to catalyze the reaction, namely Candida antarctica lipase B and Fusarium solani pisi cutinase wildtype and three single-mutation variants. The enzymes were tested in three different forms: free, immobilized as cross-linked aggregates and supported on zeolite NaY. A simple kinetic model was used to allow a quantitative comparison of the behavior of the different catalysts. It was concluded that although immobilization reduces the activity of the enzyme, the zeolite offers a much higher specific activity when compared to the cross-linked aggregates, thus supplying a heterogeneous catalyst with promising catalytic properties.  相似文献   

9.
Phospholipase C catalyzed hydrolysis of dimyristoyl phosphatidylcholine (DMPC) in phospholipid-bile salt mixed micelles was studied with particular attention on the relationship between interfacial enzyme activity and the physicochemical properties of substrate aggregates. Steady state kinetics is observed and it is argued that conditions for steady state exist because the enzyme encounters a steady supply of substrate by hopping between micelles at a rate faster than the chemical reaction rate. An existing kinetic model is reformulated to a more usable form. This presents a new approach to treating the kinetic data and allows extraction of the kinetic parameters of the model from the activity dependence on micellar lipid substrate surface concentration. The kinetic parameters were found to depend on the physicochemical properties of substrate aggregates, but remain constant over a range of lipid and bile salt concentrations. The substrate aggregates were characterized by time-resolved fluorescence quenching (TRFQ). The activity values and the micelle sizes group into two sets: (i) larger micelles for bile salt/lipid 5 with lower activity and longer steady state ( approximately 10 min). At least two sets of parameters, for bile salt/lipid 5, characterize the kinetics. Higher enzyme-micelle dissociation constant and lower catalytic rate are found for the group of smaller micelles. An explanation supporting our finding is that as micelles become smaller the overlap area for enzyme-micelle binding decreases, leading to weaker binding. Consequently the enzyme dissociation constant increases. Extension of the present approach to other phospholipases and substrates to establish its generality and correlation between micelle size and the catalytic rate are areas for future investigations.  相似文献   

10.
ATPase activities were measured in 10 mM MgCl2, 5 mM ATP, 1 mM ADP, and 1 microM FCCP with submitochondrial particles from bovine heart that had been stimulated by delta mu H+-forming substrates and with particles whose natural inhibitor protein was partially removed by heating. The activities were not linear with time. With both particles, the rate of ATP hydrolysis in the 7-fold greater than that in the steady state. Pre-steady-state and steady-state kinetic studies showed that the decrease of ATPase activity was due to the binding of ADP in a high-affinity site of the enzyme (K0.5 of 10 microM). Inhibition of ATP hydrolysis was accompanied by the binding of approximately 1 mol of ADP/mol of particulate F1; 10 microM ADP gave half-maximal binding. ADP could be replaced by IDP, but with an affinity 50-fold lower (K0.5 of 0.5 mM). Maximal inhibition by ADP and IDP was achieved in less than 5 s. Inhibition was enhanced by uncouplers. Even in the presence of pyruvate kinase and phosphoenolpyruvate, the rates of hydrolysis were about 2.5-fold higher in the first seconds of reaction than in the steady state. This decrease of ATPase activity also correlated with the binding of nearly 1 mol of ADP/mol of F1. This inhibitory ADP remained bound to the enzyme after several thousand turnovers. Apparently, it is possible to observe maximal rates of hydrolysis only in the first few catalytic cycles of the enzyme.  相似文献   

11.
We have studied the kinetics of hydrolysis of triacylglycerols, vinyl esters and p-nitrophenyl butyrate by four carboxylesterases of the HSL family, namely recombinant human hormone-sensitive lipase (HSL), EST2 from Alicyclobacillus acidocaldarius, AFEST from Archeoglobus fulgidus, and protein RV1399C from Mycobacterium tuberculosis. The kinetic properties of enzymes of the HSL family have been compared to those of a series of lipolytic and non-lipolytic carboxylesterases including human pancreatic lipase, guinea pig pancreatic lipase related protein 2, lipases from Mucor miehei and Thermomyces lanuginosus, cutinase from Fusarium solani, LipA from Bacillus subtilis, porcine liver esterase and Esterase A from Aspergilus niger. Results indicate that human HSL, together with other lipolytic carboxylesterases, are active on short chain esters and hydrolyze water insoluble trioctanoin, vinyl laurate and olive oil, whereas the action of EST2, AFEST, protein RV1399C and non-lipolytic carboxylesterases is restricted to solutions of short chain substrates. Lipolytic and non-lipolytic carboxylesterases can be differentiated by their respective value of K(0.5) (apparent K(m)) for the hydrolysis of short chain esters. Among lipolytic enzymes, those possessing a lid domain display higher activity on tributyrin, trioctanoin and olive oil suggesting, then, that the lid structure contributes to enzyme binding to triacylglycerols. Progress reaction curves of the hydrolysis of p-nitrophenyl butyrate by lipolytic carboxylesterases with lid domain show a latency phase which is not observed with human HSL, non-lipolytic carboxylesterases, and lipolytic enzymes devoid of a lid structure as cutinase.  相似文献   

12.
Within the family of large GTP-binding proteins, human guanylate binding protein 1 (hGBP1) belongs to a subgroup of interferon-inducible proteins. GTP hydrolysis activity of these proteins is much higher compared with members of other GTPase families and underlies mechanisms that are not understood. The large GTP-binding proteins form self-assemblies that lead to stimulation of the catalytic activity. The unique result of GTP hydrolysis catalyzed by hGBP1 is GDP and GMP. We investigated this reaction mechanism by transient kinetic methods using radioactively labeled GTP as well as fluorescent probes. Substrate binding and formation of the hGBP1 homodimer are fast as no lag phase is observed in the time courses of GTP hydrolysis. Instead, multiple turnover experiments show a rapid burst of P(i) formation prior to the steady state phase, indicating a rate-limiting step after GTP cleavage. Both molecules are catalytically active and cleave off a phosphate ion in the first step. Then bifurcation into catalytic inactivation, probably by irreversible dissociation of the dimer, and into GDP hydrolysis is observed. The second cleavage step is even faster than the first step, implying a rapid rearrangement of the nucleotide within the catalytic center of hGBP1. We could also show that the release of the products, including the phosphate ions, is fast and not limiting the steady state activity. We suggest that slow dissociation of the GMP-bound homodimer gives rise to the burst behavior and controls the steady state. The assembled forms of the GDP- and GMP-bound states of hGBP1 are accessible only through GTP binding and hydrolysis and achieve a lifetime of a few seconds.  相似文献   

13.
Down regulation of prolactin (PRL) receptors resulting from i.v. infusion of oPRL or human growth hormone (hGH) into female virgin rats was demonstrated. A decrease of over 85% in the number of free receptors was achieved within 15 - 30 min using infusion of oPRL or hGH at 25 micrograms/h and remained at this level until the end of infusion. Ovine growth hormone or recombinant bovine growth hormone at ten-fold higher concentration had no effect at all. The decrease in the specific binding resulted from a lower number of binding sites and not from change in the dissociation constants. The decrease in the total receptors in the liver was more gradual and leveled off at 40 - 50% of the initial value. Our results suggest that a change in blood PRL or hGH level may lead to a new steady state in the number, occupancy and distribution of prolactin receptors.  相似文献   

14.
The effects of aqueous surfactant solutions on the kinetics and stability of cutinase from Fusarium solani pisi were studied. The surfactant sodium bis[2-ethylhexyl]ester sulfosuccinic acid (AOT) acts as a pseudo-competitive inhibitor within a limited concentration range relative to the hydrolysis of short-chain p-nitrophenyl esters. For higher concentrations a hyperbolic mixed inhibition takes place. A pseudo-activation of hydrolysis in presence of AOT and hexadecyltrimethyl-ammonium bromide (CTAB) was observed. CTAB has similar effects on kinetics of cutinase. Cutinase revealed to be stable in CTAB solutions, with activity retention as high as 80%. AOT has a deleterious effect on the enzyme in the time course, resulting in acute loss of activity possibly related with unfolding of the protein structure. A relation between deactivation rate constants and AOT/cutinase concentration ratios is suggested. The presence of the linear alcohol, 1-hexanol, was included in these solutions, in the attempt to interpret the deactivation of cutinase when encapsulated in reversed micelle systems in the absence of this co-surfactant.  相似文献   

15.
The kinetic mechanism of dextransucrase was studied using the Streptococcus mutans enzyme purified by affinity chromatography to a specific activity of 36.9 mumol/min/mg of enzyme. In addition to dextran synthesis, the enzyme catalyzed sucrose hydrolysis and isotope exchange between fructose and sucrose. The rates of sucrose hydrolysis and dextran synthesis were partitioned as a function of dextran concentration such that exclusive sucrose hydrolysis was observed in the absence of dextran and exclusive dextran synthesis at high dextran concentrations. An analogous situation was observed with fructose-dependent partitioning of sucrose hydrolysis and fructose exchange. Steady state dextran synthesis and fructose isotope exchange kinetics were simplified by assay at dextran or fructose concentrations high enough to eliminate significant contributions from sucrose hydrolysis. This limited dextran synthesis assays to dextran concentrations above apparent saturation. The limitation was diminished by establishing conditions in which the enzyme does not distinguish between dextran as a substrate and product which allowed initial discrimination among mechanisms on the basis of the presence or absence of dextran substrate inhibition. No inhibition was observed, which excluded ping-pong and all but three common sequential mechanisms. Patterns of initial velocity fructose production inhibition and fructose isotope exchange at equilibrium were consistent with dextran synthesis proceeding by a rapid equilibrium random mechanism. A nonsequential segment was apparent in the exchange reaction between fructose and sucrose assayed in the absence of dextran. However, the absence of detectable glucosyl exchange between dextrans and the lack of steady state dextran substrate inhibition indicate that glucosyl transfer to dextran must occur almost exclusively through the sequential route. A review of the kinetic constants from steady state dextran synthesis, fructose product inhibition, and fructose isotope exchange showed a consistency in constants derived from each reaction and revealed that dextran binding increases the affinity of sucrose and fructose for dextransucrase.  相似文献   

16.
The effect of sarcomere length and stretching on the tension and the rate of ATP splitting was studied using small fiber bundles from glycerinated rabbit psoas muscle. The rate of ATP slitting was determined by measuring ADP production, while the tension development in response to a contracting solution (at pCa 5.3) was recorded in the same preparation. The isometric tension developed by the preparation decreased when the sarcomere length was increased. The decrease of tension development was accompanied by a decrease in the rate of ATP splitting. If a preparation exerting steady isometric tension was stretched by 5--10% at a velocity of 0.1 mm/s, the rate of ATP splitting was increased after stretching, while the steady isometric tension attained after stretching was also higher than the initial value. The extent of the excess ATP splitting caused by stretching decreased with increasing sarcomere length. These results suggest that the rate of the interaction cycle between actin and myosin molecules may increase as a result of stretching.  相似文献   

17.
Smooth muscle contraction is controlled in part by the state of phosphorylation of myosin. A recently discovered actin and calmodulin-binding protein, named caldesmon, may also be involved in regulation of smooth muscle contraction. Caldesmon cross-links actin filaments and also inhibits actin-activated ATP hydrolysis by myosin, particularly in the presence of tropomyosin. We have studied the effect of caldesmon on the rate of hydrolysis of ATP by skeletal muscle myosin subfragment-1, a system in which phosphorylation of the myosin is not important in regulation. Caldesmon is a very effective inhibitor of ATP hydrolysis giving up to 95% inhibition. At low ionic strength (approximately 20 mM) this effect does not require smooth muscle tropomyosin, whereas at high ionic strength (approximately 120 mM) tropomyosin enhances the inhibitory activity of caldesmon at low caldesmon concentrations. Cross-linking of actin is not essential for inhibition of ATP hydrolysis to occur since at high ionic strength there is very little cross-linking as determined by a low speed sedimentation assay. Under all conditions examined, the decrease in the rate of ATP hydrolysis is accompanied by a decrease in the binding of myosin subfragment-1 to actin. Furthermore, caldesmon weakens the equilibrium binding of myosin subfragment-1 to actin in the presence of pyrophosphate. We conclude that caldesmon has a general weakening effect on the binding of skeletal muscle myosin subfragment-1 to actin and that this weakening in binding may be responsible for inhibition of ATP hydrolysis.  相似文献   

18.
Cutinase from Fusarium solani pisi is the model-system for a new approach to assess and enhance protein stability based on the use of synthetic triazine-scaffolded affinity ligands as a novel protein-stabilizing tool. The active site of cutinase is excluded from the main surface regions postulated to be involved in early protein's thermal unfolding events. Hence, these regions are suitable targets for binding complementary affinity ligands with a potential stabilizing effect. A random solid-phase combinatorial library of triazine-bisubstituted molecules was screened for binding cutinase by a rapid fluorescence-based method and affinity chromatography. The best binding substituents were combined with those previously selected by screening a rationally designed library. A second-generation solid-phase biased library was designed and synthesized, following a semi-rational methodology. A dual screening of this library enabled the selection of ligands binding cutinase with higher affinity while retaining its functionality. These compounds were utilized for thermostability assessment with adsorbed cutinase at 60 degrees C and pH 8.0. When bound to different types of ligands, the enzyme showed markedly distinct activity retention profiles, with some synthetic affinity ligands displaying a stabilizing effect on cutinase and others a clearly destabilizing effect, when compared with the free enzyme.  相似文献   

19.
The kinetics of ATP hydrolysis and tension responses were studied simultaneously in a permeabilized preparation of cardiac tissue of the guinea pig. This was achieved by combining laserflash photolysis of P3-1-(2-nitrophenyl)ethyladenosine 5'-triphosphate ("caged-ATP") and a rapid freezing technique. In the presence of calcium ions, tension increased following the photolytic production of ATP with a half-time of 0.3 s. The timecourse of ATP hydrolysis consisted of an initial rapid phase followed by a steady-state hydrolysis rate of 0.4 s-1, indicating that the rate-limiting step of the ATPase in isometric fibers is slower and subsequent to the nucleotide hydrolysis step: the isometric steady state intermediate is probably an actomyosin-ADP complex. In the absence of calcium ions, rigor tension decreased upon the photolytic production of ATP with a half-time of 0.45 s. The time course of ATP hydrolysis was biphasic with a rapid initial phase of ATP hydrolysis, followed by a steady-state hydrolysis rate which was too slow to measure over the time scale of these experiments (less than 0.04 s-1). A comparison of the results obtained in this study with those reported for rabbit skeletal muscle reveals qualitative similarities between cardiac and skeletal muscle and also quantitative differences in their physiological and kinetic behavior.  相似文献   

20.
Deactivation and conformational changes of cutinase in reverse micelles   总被引:1,自引:0,他引:1  
Deactivation data and fluorescence intensity changes were used to probe functional and structural stability of cutinase in reverse micelles. A fast deactivation of cutinase in anionic (AOT) reverse micelles occurs due to a reversible denaturation process. The deactivation and denaturation of cutinase is slower in small cationic (CTAB/1-hexanol) reverse micelles and does not occur when the size of the cationic reverse micellar water-pool is larger than cutinase. In both systems, activity loss and denaturation are coupled processes showing the same trend with time. Denaturation is probably caused by the interaction between the enzyme and the surfactant interface of the reversed micelle. When the size of the empty reversed micelle water-pool is smaller than cutinase (at W0 5, with W0 being the water:surfactant concentration ratio) a three-state model describes denaturation and deactivation with an intermediate conformational state existing on the path from native to denaturated cutinase. This intermediate was clearly detected by an increase in activity and shows only minor conformational changes relative to the native state. At W0 20, the size of the empty water-pool was larger than cutinase and the data was well described by a two-state model for both anionic and cationic reverse micelles. For AOT reverse micelles at W0 20, the intermediate state became a transient state and the deactivation and denaturation were described by a two-state model in which only native and denaturated cutinase were present. For CTAB/1-hexanol reverse micelles at W0 20, the native cutinase was in equilibrium with an intermediate state, which did not suffer denaturation. 1-Hexanol showed a stabilizing effect on cutinase in reverse micelles, contributing to the higher stabilities observed in the cationic CTAB/1-hexanol reverse micelles. Copyright 1998 John Wiley & Sons, Inc.  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