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.
p-Hydroxyphenylacetate (HPA) 3-hydroxylase is a two-component flavoprotein monooxygenase that catalyzes the hydroxylation of p-hydroxyphenylacetate to form 3,4-dihydroxyphenylacetate. Based on structures of the oxygenase component (C2), both His-120 and Ser-146 are located ∼2.8 Å from the hydroxyl group of HPA. The variants H120N, H120Q, H120Y, H120D, and H120E can form C4a-hydroperoxy-FMN (a reactive intermediate necessary for hydroxylation) but cannot hydroxylate HPA. The impairment of H120N is not due to substrate binding because the variant can still bind HPA. In contrast, the H120K variant catalyzes hydroxylation with efficiency comparable with that of the wild-type enzyme; the hydroxylation rate constant for H120K is 5.7 ± 0.6 s−1, and the product conversion ratio is 75%, compared with values of 16 s−1 and 90% for the wild-type enzyme. H120R can also catalyze hydroxylation, suggesting that a positive charge on residue 120 can substitute for the hydroxylation function of His-120. Because the hydroxylation reaction of wild-type C2 is pH-independent between pH 6 and 10, the protonation status of key components required for hydroxylation likely remains unchanged in this pH range. His-120 may be positively charged for selective binding to the phenolate form of HPA, i.e. to form the Hisδ+·HPAδ− complex, which in turn promotes oxygen atom transfer via an electrophilic aromatic substitution mechanism. Analysis of Ser-146 variants revealed that this residue is necessary for but not directly engaged in hydroxylation. Product formation in S146A is pH-independent and constant at ∼70% over a pH range of 6–10, whereas product formation for S146C decreased from ∼65% at pH 6.0 to 27% at pH 10.0. These data indicate that the ionization of Cys-146 in the S146C variant has an adverse effect on hydroxylation, possibly by perturbing formation of the Hisδ+·HPAδ− complex needed for hydroxylation.  相似文献   

2.
To investigate the effect of pH on sequence-specific binding, a thermodynamic characterization of the interaction of the protein MunI with a specific, and a nonspecific, oligonucleotide was performed. MunI is a type II restriction endonuclease which is able to bind specifically, but loses its enzymatic activity in the absence of magnesium ions. Comparison of the specific and nonspecific interactions at 10 and 25 degrees C shows that the latter is accompanied by a small change in enthalpy, and a negligible change in constant pressure heat capacity. On going through the pH range 5.75-9.0 at 25 degrees C, the affinity of specific complex formation is reduced by 20-fold. The interaction is accompanied by the protonation of groups assumed to be on the protein. Based on the simplest model that will fit the data, two distinct protonation events are observed. At low pH, two groups per protein molecule undergo protonation with a pK(a) of 6.0 and 6.9 in the free and bound forms, respectively. At high pH, a further independent protonation occurs involving two groups with pK(a) values of 8.9 and approximately 10.7 in the free and bound forms, respectively. The change in heat capacity ranges from -2.7 to -1.7 kJ mol(-1) K(-1) in going from pH 6.5 to 8.5. This range of variation of change in heat capacity can be accounted for by the effects of protonation of the interacting molecules. The change in heat capacity, calculated from surface area burial using a previously established relationship (1.15 kJ mol(-1) K(-1)), does not correlate well with the experimentally determined values.  相似文献   

3.
The effect of pH on the (1)H NMR spectrum, reduction potential, and self-exchange rate constant of the novel plastocyanin (PCu) from the fern plant Dryopteris crassirhizoma has been studied. The results are compared with those for the higher-plant PCu from parsley. In the (1)H NMR spectrum of D. crassirhizoma PCu(I), there is no sign that either of the His ligands is protonated at pH* down to 5.4. The reduction potentials of D. crassirhizoma and parsley PCu are 382 and 379 mV, respectively, at pH 7.4. When the pH value is decreased, the reduction potential of parsley PCu is seen to increase quite dramatically, consistent with protonation at His87 in PCu(I). A pK(a) of 5.8 is obtained from the electrochemistry data, consistent with a value of 5.6 determined by NMR. The reduction potential of D. crassirhizoma PCu exhibits a much less pronounced dependence on pH. The self-exchange rate constant of D. crassirhizomaPCu(I) is 3.4 x 10(3) M(-1) s(-1) at pH* 7.9. This is the smallest self-exchange rate constant reported to date for a PCu and can be rationalized by considering the altered distribution of charged residues on the surface of the D. crassirhizoma protein compared to the charge distributions of other higher-plant PCus. The self-exchange rate constant increases to 9 x 10(3) M(-1) s(-1) at pH* 5.4, consistent with enhanced protein-protein association at lower pH*, and the absence of His87 protonation in D. crassirhizoma PCu(I) in the accessible pH range.  相似文献   

4.
The gene coding for thermophilic xylose (glucose) isomerase of Clostridium thermosulfurogenes was isolated and its complete nucleotide sequence was determined. The structural gene (xylA) for xylose isomerase encodes a polypeptide of 439 amino acids with an estimated molecular weight of 50,474. The deduced amino acid sequence of thermophilic C. thermosulfurogenes xylose isomerase displayed higher homology with those of thermolabile xylose isomerases from Bacillus subtilis (70%) and Escherichia coli (50%) than with those of thermostable xylose isomerases from Ampullariella (22%), Arthrobacter (23%), and Streptomyces violaceoniger (24%). Several discrete regions were highly conserved throughout the amino acid sequences of all these enzymes. To identify the histidine residue of the active site and to elucidate its function during enzymatic xylose or glucose isomerization, histidine residues at four different positions in the C. thermosulfurogenes enzyme were individually modified by site-directed mutagenesis. Substitution of His101 by phenylalanine completely abolished enzyme activity whereas substitution of other histidine residues by phenylalanine had no effect on enzyme activity. When His101 was changed to glutamine, glutamic acid, asparagine, or aspartic acid, approximately 10-16% of wild-type enzyme activity was retained by the mutant enzymes. The Gln101 mutant enzyme was resistant to diethylpyrocarbonate inhibition which completely inactivated the wild-type enzyme, indicating that His101 is the only essential histidine residue involved directly in enzyme catalysis. The constant Vmax values of the Gln101, Glu101, Asn101, and Asp101 mutant enzymes over the pH range of 5.0-8.5 indicate that protonation of His101 is responsible for the reduced Vmax values of the wild-type enzyme at pH below 6.5. Deuterium isotope effects by D-[2-2H]glucose on the rate of glucose isomerization indicated that hydrogen transfer and not substrate ring opening is the rate-determining step for both the wild-type and Gln101 mutant enzymes. These results suggest that the enzymatic sugar isomerization does not involve a histidine-catalyzed proton transfer mechanism. Rather, essential histidine functions to stabilize the transition state by hydrogen bonding to the C5 hydroxyl group of the substrate and this enables a metal-catalyzed hydride shift from C2 to C1.  相似文献   

5.
Self-cleavage of the genomic and antigenomic ribozymes from hepatitis delta virus (HDV) requires divalent cation for optimal activity. Recently, the HDV genomic ribozyme has been shown to be active in NaCl in the absence of added divalent metal ion at low pH (apparent pKa 5.7). However, we find that the antigenomic ribozyme is 100 to 1000-fold less active under similar conditions. With deletion of a three-nucleotide sequence (C41-A42-A43) unique to the genomic ribozyme, the rate constant for cleavage decreased substantially, while activity of the antigenomic ribozyme was enhanced by introducing a CAA sequence. From the crystal structure, it has been proposed that C41 in this sequence is protonated. To investigate a possible connection between activity at low pH and protonation of C41, mutations were made that were predicted to either eliminate protonation or alter the nature of the tertiary interaction upon protonation. In the absence of added Mg2+, these mutations reduced activity and eliminated the observed pH-rate dependence. Thermal denaturation studies revealed a pH-sensitive structural feature in the genomic ribozyme, while unfolding of the mutant ribozymes was pH-independent. We propose that, in the absence of added Mg2+, protonation of C41 contributes to enhanced activity of the HDV genomic ribozyme at low pH.  相似文献   

6.
Samples CR  Howard T  Raushel FM  DeRose VJ 《Biochemistry》2005,44(33):11005-11013
Phosphotriesterase (PTE) is a binuclear metalloenzyme that catalyzes the hydrolysis of organophosphates, including pesticides and chemical warfare agents, at rates approaching the diffusion controlled limit. The catalytic mechanism of this enzyme features a bridging solvent molecule that is proposed to initiate nucleophilic attack at the phosphorus center of the substrate. X-band EPR spectroscopy is utilized to investigate the active site of Mn/Mn-substituted PTE. Simulation of the dominant EPR spectrum from the coupled binuclear center of Mn/Mn-PTE requires slightly rhombic zero-field splitting parameters. Assuming that the signal arises from the S = 2 manifold, an exchange coupling constant of J = -2.7 +/- 0.2 cm(-)(1) (H(ex) = -2JS(1) x S(2)) is calculated. A kinetic pK(a) of 7.1 +/- 0.1 associated with loss in activity at low pH indicates that a protonation event is responsible for inhibition of catalysis. Analysis of changes in the EPR spectrum as a function of pH provides a pK(a) of 7.3 +/- 0.1 that is assigned as the protonation of the hydroxyl bridge. From the comparison of kinetic and spectral pK(a) values, it is concluded that the loss of catalytic activity at acidic pH results from the protonation of the hydroxide that bridges the binuclear metal center.  相似文献   

7.
Deaza analogues of nucleobases are potential drugs against infectious diseases caused by parasites. A caveat is that apart from binding their target parasite enzymes, they also bind and inhibit enzymes of the host. In order to design derivatives of deaza analogues which specifically bind target enzymes, knowledge of their molecular structure, protonation state, and predominant tautomers at physiological conditions is essential. We have employed resonance Raman spectroscopy at an excitation wavelength of 260 nm, to decipher solution structure of 9-deazaguanine (9DAG) and 9-deazahypoxanthine (9DAH). These are analogues of guanine and hypoxanthine, respectively, and have been exploited to study static complexes of nucleobase binding enzymes. Such enzymes are known to perturb pKa of their ligands, and thus, we also determined solution structures of these analogues at two, acidic and alkaline, pH. Structure of each possible protonation state and tautomer was computed using density functional theoretical calculations. Species at various pHs were identified based on isotopic shifts in experimental wavenumbers and by comparing these shifts with corresponding computed isotopic shifts. Our results show that at physiological pH, N1 of pyrimidine ring in 9DAG and 9DAH bears a proton. At lower pH, N3 is place of protonation, and at higher pH, deprotonation occurs at N1 position. The proton at N7 of purine ring remains intact even at pH 12.5. We have further compared these results with naturally occurring nucleotides. Our results identify key vibrational modes which can report on hydrogen bonding interactions, protonation and deprotonation in purine rings upon binding to the active site of enzymes.  相似文献   

8.
R L Blakley  L Cocco 《Biochemistry》1985,24(17):4704-4709
Stopped-flow spectrophotometry and stopped-flow fluorometry have been used to study the binding of methotrexate (MTX) and 3-deazamethotrexate (3-deazaMTX) to dihydrofolate reductase (DHFR) isoenzymes from Streptococcus faecium and from Lactobacillus casei. The absorbance change and fluorescence quenching that occur when MTX binds to DHFR isoenzyme II from S. faecium (SFDHFR II) are both biphasic and give similar apparent rate constants for both phases. The faster phase has an apparent rate constant that is dependent on MTX concentration and therefore corresponds to the initial binding reaction. From the concentration dependence it has been calculated that the association rate constant is 3.0 X 10(5) M-1 s-1 at 20 degrees C and pH 7.3, and the association constant (equilibrium constant) under these conditions is 5.8 X 10(5) M-1. By examination of the amplitude of the fast-phase absorbance change at various wavelengths, it has been determined that the absorbance change occurring in the fast phase is due to MTX protonation. Within the limits of the method it was thus not possible to detect a difference in the rates of binding and of protonation of MTX. The MTX association rate constant is pH dependent, decreasing 330-fold as the pH is decreased from 5.0 to 9.0. The data fit well to a curve generated by assuming a single ionization with a pKa of 6.0 and a pH-independent association rate constant 1000-fold greater for binding of protonated MTX to SFDHFR II than for binding of unprotonated MTX.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
We studied the dependence of activity and stability of proteolytic enzymes in salmon roe on pH and temperature. The activity of proteolytic enzymes in roe was primarily determined by proteinases. These enzymes were active at acid pH and had an optimum of 3.6. A study of subclasses of proteolytic enzymes in salmon roe and the published data suggest that the activity of proteinases may be related to the presence of aspartyl proteinases (cathepsin D). Serine proteinases and metalloenzymes were not found in roe. The activity of cysteine proteinases was low. The proposed conditions of pasteurization favored the complete inactivation of salmon roe at pH 6.0-6.4.  相似文献   

10.
We studied the dependence of activity and stability of proteolytic enzymes in salmon roe on pH and temperature. The activity of proteolytic enzymes in roe was primarily determined by proteinases. These enzymes were active at acid pH and had an optimum of 3.6. A study of subclasses of proteolytic enzymes in salmon roe and the published data suggest that the activity of proteinases may be related to the presence of aspartyl proteinases (cathepsin D). Serine proteinases and metalloenzymes were not found in roe. The activity of cysteine proteinases was low. The proposed conditions of pasteurization favored the complete inactivation of salmon roe at pH 6.0–6.4.  相似文献   

11.
The acid-base properties of ionophore A23187 in methanol-water solutions (0--95% w/w) and bound to unilamellar vesicles of dimyristoylphosphatidylcholine were examined by ultraviolet and fluorescence spectroscopy, and the spectral properties for the acidic and basic forms were defined under these conditions. Standard mixed-solvent buffers were employed to calibrate pH measurement in the methanol-water solvents. In 65% methanol-water, two protonation equilibria were observed, the most basic of which displayed a value for the logarithm of the protonation constant (log KH) of 7.19 +/- 0.05 at 25 degrees C and 0.05 M ionic strength. Instability of A23187 was encountered below pH approximately 4; however, decomposition was slow enough to allow log KH for the more acidic equilibrium to be estimated as 1.28. Comparison of these results to those obtained with the methyl ester of A23187 (log KH = 1.32) and literature values for other model compounds allowed assignment of the more basic equilibrium to the carboxylic acid moiety and the more acidic one to the N-methylamino substituent of the benzoxazole ring. log KH of the carboxylic acid increased from 5.69 +/- 0.05 to 9.37 +/- 0.05 over the range of solvent polarity encompassed by water to 95% methanol-water. Values for the ground state (absorption) and first excited state (fluorescence) were equal within experimental error. The logarithm of the protonation constant for the membrane-bound ionophore, measured under conditions where the surface potential generated by ionization did not significantly alter the equilibrium, was found to be 7.85 +/- 0.05 at 25 degrees C and at ionic strength of 0.05 M in the aqueous phase. The value agrees with that observed in 80% methanol-water, as does the wavelength of maximum fluorescence emission for the membrane-bound free acid. An interfacial location for the monoprotonated form of the benzoxazolate moiety is proposed, both above and below the membrane phase transition temperature. The location of other regions of the A23187 molecule could not be assessed from these data.  相似文献   

12.
Due to its elevated cellulolytic activity, the filamentous fungus Trichoderma harzianum (T. harzianum) has considerable potential in biomass hydrolysis application. Cellulases from Trichoderma reesei have been widely used in studies of cellulose breakdown. However, cellulases from T. harzianum are less-studied enzymes that have not been characterized biophysically and biochemically as yet. Here, we examined the effects of pH and temperature on the secondary and tertiary structures, compactness, and enzymatic activity of cellobiohydrolase Cel7A from T. harzianum (Th Cel7A) using a number of biophysical and biochemical techniques. Our results show that pH and temperature perturbations affect Th Cel7A stability by two different mechanisms. Variations in pH modify protonation of the enzyme residues, directly affecting its activity, while leading to structural destabilization only at extreme pH limits. Temperature, on the other hand, has direct influence on mobility, fold, and compactness of the enzyme, causing unfolding of Th Cel7A just above the optimum temperature limit. Finally, we demonstrated that incubation with cellobiose, the product of the reaction and a competitive inhibitor, significantly increased the thermal stability of Th Cel7A. Our studies might provide insights into understanding, at a molecular level, the interplay between structure and activity of Th Cel7A at different pH and temperature conditions.  相似文献   

13.
Two homologous Delta5-3-ketosteroid isomerases from Comamonas testosteroni (TI-WT) and Pseudomonas putida biotype B (PI-WT) exhibit different pH activity profiles. TI-WT loses activity below pH 5.0 due to the protonation of the conserved catalytic base, Asp-38, while PI-WT does not. Based on the structural analysis of PI-WT, the critical catalytic base, Asp-38, was found to form a hydrogen bond with the indole ring NH of Trp-116, which is homologously replaced with Phe-116 in TI-WT. To investigate the role of Trp-116, we prepared the F116W mutant of TI-WT (TI-F116W) and the W116F mutant of PI-WT (PI-W116F) and compared kinetic parameters of those mutants at different pH levels. PI-W116F exhibited significantly decreased catalytic activity at acidic pH like TI-WT, whereas TI-F116W maintained catalytic activity at acidic pH like PI-WT and increased the kcat/Km value by 2.5- to 4.7-fold compared with TI-WT at pH 3.8. The crystal structure of TI-F116W clearly showed that the indole ring NH of Trp-116 could form a hydrogen bond with the carboxyl oxygen of Asp-38 like that of PI-WT. The present results demonstrate that the activities of both PI-WT and TI-F116W at low pH were maintained by a tryptophan, which was able not only to lower the pKa value of the catalytic base but also to increase the substrate affinity. This is one example of the strategy nature can adopt to evolve the diversity of the catalytic function in the enzymes. Our results provide insight into deciphering the molecular evolution of the enzyme and creating novel enzymes by protein engineering.  相似文献   

14.
Murine melanoma melanosomal tyrosinase, solubilised at pH 6.8 and 1% Igepal, exhibits a lag in cresolase activity which increases with increasing concentration of tyrosine. The enzyme, solubilised at pH 5.0 and assayed at pH 5.0, does not exhibit lag even at inhibitory concentrations of tyrosine while the same enzyme when assayed at pH 6.8 exhibits characteristic lag. When the enzyme was solubilised from a melanosomal fraction with detergent/water without any buffer, significant linear activity for 2 h was seen at an inhibitory concentration of tyrosine, indicating for the first time the presence of a form of tyrosinase without lag and inhibition by excess tyrosine. Exposure of the enzyme solubilised in buffer/detergent at pH 6.8 to rapid decrease in pH to 5.0 or 4.7 makes the enzyme remain irreversibly in the form without characteristic lag, even at an inhibitory concentration of tyrosine and at pH 6.8. These results may be interpreted as follows. The enzyme at pH 6.8 exists in the E form with an allosteric site for tyrosine. Decrease of the pH of the enzyme solution from 6.8 to 5.0 or 4.7 by dialysis results in the reversible protonation of the enzyme, which no longer binds tyrosine at its allosteric site and consequently inhibition by excess tyrosine and lag were not observed at acidic pH. However, if the enzyme was rapidly brought to pH 5.0 from 6.8 it remains irreversibly in the protonated form even at pH 6.8. Ascorbic acid acts as an effective reductant for the hydroxylation of tyrosine by tyrosinase, while 3,4-dihydroxyphenylalanine is both an effective reductant and counteracts the inhibition by tyrosine at pH 6.8.  相似文献   

15.
Reconstitution of factor VIII from isolated heavy chain (HC) and light chain (LC) shows pH-dependence. In the presence of Ca2+, up to 80% of native factor VIII activity was recovered over a wide range of pH. In contrast, affinity of HC and LC was maximal at pH 6.5-6.75 (Kd approximately 4 nM), whereas a Kd approximately 20 nM was observed at physiological pH (7.25). The effect of Cu2+ (0.5 microM total Cu2+) on maximal activity regenerated was negligible at pH 6.25-8.0. However, this level of Cu2+ increased the inter-chain affinity by approximately 5-fold at pH 7.25. This effect resulted from an approximately 1.5-fold increased association rate constant (k(on)) and an approximately 3-fold reduced dissociation rate constant (k(off)). High affinity (Kd=5.3 fM) of the factor VIII heterodimer for Cu2+ was estimated by increases in cofactor activity. No significant increase in inter-chain affinity was observed when either isolated chain was reacted with Cu2+ followed by addition of the complementary chain. Together, these results suggest that the protonation state of specific residues modulates inter-chain affinity. Furthermore, copper ion contributes to the maintenance of the heterodimer at physiologic pH by a mechanism consistent with bridging the two chains.  相似文献   

16.
In batch culture for Poly(vinyl alcohol) (PVA)-degrading enzyme (PVAase) production by a mixed culture, higher pH (pH 7.5) was favorable for PVAase production at the prophase of cultivation, but lower pH (pH 7.0) was favorable at the anaphase. This situation was caused by the fact that the optimum pH for different key enzymes [PVA dehydrogenase (PVADH) and oxidized PVA hydrolase (OPH)] production is various. The activity and average specific production rate of PVADH reached the highest values at constant pH 7.5, whereas those of OPH appeared at pH 7.0. A two-stage pH control strategy was therefore developed and compared for its potential in improving PVAase production. By using this strategy, the maximal PVAase activity reached 2.05 U/ml, which increased by 15.2% and 24.2% over the fermentation at constant pH 7.5 and 7.0.  相似文献   

17.
Heparin binding to rabbit histidine-rich glycoprotein (HRG) was studied in a purified system, allowing for determination of a heparin dissociation constant of approximately 5.5 X 10(-8) M for the interaction with HRG at pH 7.0. The strong interaction between heparin and HRG was demonstrated to be competitive with the binding of both antithrombin and thrombin to the heparin chain. HRG was further tested as a modulator of the anticoagulant activity of heparin by comparing rates of the heparin-catalyzed reaction between antithrombin and thrombin in the presence and absence of added HRG. The heparin-antithrombin-thrombin reaction was modeled using the formalism of a two-substrate enzyme-catalyzed reaction with heparin as the enzyme and HRG analyzed as an enzyme inhibitor. HRG was shown to compete with both antithrombin and thrombin for binding to heparin by this kinetic analysis. Thus, both the kinetic and heparin-binding data indicate that the mechanism by which HRG modulates heparin anticoagulant activity involves competition for heparin with both the inhibitor and the protease. Inhibition by HRG of the heparin-catalyzed reaction was found to be highly dependent on pH, with a sharp increase in inhibition from about 15% to greater than 90% observed as pH was lowered from 7.4 to 7.0. Since little change in the rate of the heparin-catalyzed inhibition of thrombin by antithrombin occurs in this pH region, the dramatic change in HRG inhibition seen upon pH titration may reflect increasing ionic interaction between heparin and HRG due to the protonation of histidine residues which occurs in this pH region.  相似文献   

18.
I Vass  S Styring 《Biochemistry》1991,30(3):830-839
The effect of protonation events on the charge equilibrium between tyrosine-D and the water-oxidizing complex in photosystem II has been studied by time-resolved measurements of the EPR signal IIslow at room temperature. The flash-induced oxidation of YD by the water-oxidizing complex in the S2 state is a monophasic process above pH 6.5 and biphasic at lower pHs, showing a slow and a fast phase. The half-time of the slow phase increases from about 1 s at pH 8.0 to about 20 s at pH 5.0, whereas the half-time of the fast phase is pH independent (0.4-1 s). The dark reduction of YD+ was followed by measuring the decay of signal IIslow at room temperature. YD+ decays in a biphasic way on the tens of minutes to hours time scale. The minutes phase is due to the electron transfer to YD+ from the S0 state of the water-oxidizing complex. The half-time of this process increases from about 5 min at pH 8.0 to 40 min at pH 4.5. The hours phase of YD+ has a constant half-time of about 500 min between pH 4.7 and 7.2, which abruptly decreases above pH 7.2 and below pH 4.7. This phase reflects the reduction of YD+ either from the medium or by an unidentified redox component of PSII in those centers that are in the S1 state. The titration curve of the half-times for the oxidation of YD reveals a proton binding with a pK around 7.3-7.5 that retards the electron transfer from YD to the water-oxidizing complex. We propose that this monoprotic event reflects the protonation of an amino acid residue, probably histidine-190 on the D2 protein, to which YD is hydrogen bonded. The titration curves for the oxidation of YD and for the reduction of YD+ show a second proton binding with pK approximately 5.8-6.0 that accelerates the electron transfer from YD to the water-oxidizing complex and retards the process in the opposite direction. This protonation most probably affects the water-oxidizing complex. From the measured kinetic parameters, the lowest limits for the equilibrium constants between the S0YD+ and the S1YD as well as between the S1YD+ and S2YD states were estimated to be 5 and 750-1000, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Ultrafast, laser-induced pH jump with time-resolved photoacoustic detection has been used to investigate the early protonation steps leading to the formation of the compact acid intermediate (I) of apomyoglobin (ApoMb). When ApoMb is in its native state (N) at pH 7.0, rapid acidification induced by a laser pulse leads to two parallel protonation processes. One reaction can be attributed to the binding of protons to the imidazole rings of His24 and His119. Reaction with imidazole leads to an unusually large contraction of -82 +/- 3 ml/mol, an enthalpy change of 8 +/- 1 kcal/mol, and an apparent bimolecular rate constant of (0.77 +/- 0.03) x 10(10) M(-1) s(-1). Our experiments evidence a rate-limiting step for this process at high ApoMb concentrations, characterized by a value of (0. 60 +/- 0.07) x 10(6) s(-1). The second protonation reaction at pH 7. 0 can be attributed to neutralization of carboxylate groups and is accompanied by an apparent expansion of 3.4 +/- 0.2 ml/mol, occurring with an apparent bimolecular rate constant of (1.25 +/- 0.02) x 10(11) M(-1) s(-1), and a reaction enthalpy of about 2 kcal/mol. The activation energy for the processes associated with the protonation of His24 and His119 is 16.2 +/- 0.9 kcal/mol, whereas that for the neutralization of carboxylates is 9.2 +/- 0.9 kcal/mol. At pH 4.5 ApoMb is in a partially unfolded state (I) and rapid acidification experiments evidence only the process assigned to carboxylate protonation. The unusually large contraction and the high energetic barrier observed at pH 7.0 for the protonation of the His residues suggests that the formation of the compact acid intermediate involves a rate-limiting step after protonation.  相似文献   

20.
Seventy-five fungal strains from different groups of basidiomycetes, newly isolated from rotten wood, were screened for pectinolytic activity. Despite the fact that basidiomycetes are scarcely referred to as pectinase producers, the polygalacturonase (PG) activity was detected in 76% of the strains; 16% with activity higher than 40 nkat/g, 40% between 13.3 and 40 nkat/g, and 44% with activity lower than 13.3 nkat/g. The highest productions were obtained among the fungi from order Aphyllophorales, family Polyporaceae. The characterization of the enzymes from the highest PG producers (Lentinus sp., Gloeophyllum striatum, Pycnoporus sanguineus, Schizophyllum commune) showed optimum temperature for catalytic activity at 60-70 degrees C and two peaks of pH optimum (3.5-4.5 and 8.5-9.5). The enzymes exhibited high pH stability (3.0-11.0) but after incubation at 40 degrees C for 1 h their activity dropped by 18-73%.  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