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.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


2.
We report spectrophotometric equilibrium studies of both the self-association of the new antibiotic iremycin and of its binding to calf thymus DNA in solution (ionic strength 0.2 M; pH 6.0). Iremycin forms dimers in this solution with a dimerization constant K4=(1.19 ± 0.10) × 103 M−1. This equilibrium is taken into account in the evaluation of the interaction of iremycin with DNA. The binding behaviour can be completely described by a single binding mechanism of monomeric iremycin to DNA with allowance both for neighbour exclusion and for cooperativity of interaction. The three intrinsic binding parameters for the homogeneous model were determined simultaneously by a least squares fit of the original titration data: equilibrium constant of cooperative binding K = (2.72 ± 0.66) × 105 M−1 cooperativity parameter σ=0.38±3.27 ± 0.32. The binding parameters of iremycin and adriamycin and their microbial activities are compared.  相似文献   

3.
High-pressure liquid-chromatography and microcalorimetry have been used to determine equilibrium constants and enthalpies of reaction for the disproportionation reaction of adenosine 5′-diphosphate (ADP) to adenosine 5′-triphosphate (ATP) andadenosine 5′-monophosphate (AMP). Adenylate kinase was used to catalyze this reaction. The measurements were carried out over the temperature range 286 to 311 K, at ionic strengths varying from 0.06 to 0.33 mol kg−1, over the pH range 6.04 to 8.87, and over the pMg range 2.22 to 7.16, where pMg = -log a(Mg2+). The equilibrium model developed by Goldberg and Tewari (see the previous paper in this issue) was used for the analysis of the measurements. Thus, for the reference reaction: 2 ADp3− (ao) AMp2− (ao)+ ATp (ao), K° = 0.225 ± 0.010, ΔG° = 3.70 +- 0.11 kJ mol −1, ΔH° = −1.5 ± 1. 5 kJ mol −1, °S ° = −17 ± 5 J mol−1 K−1, and ACPp°≈ = −46 J mo1l−1 K−1 at 298.15 K and 0.1 MPa. These results and the thermodynamic parameters for the auxiliary equilibria in solution have been used to model the thermodynamics of the disproportionation reaction over a wide range of temperature, pH, ionic strength, and magnesium ion morality. Under approximately physiological conditions (311.15 K, pH 6.94, [Mg2+] = 1.35 × 10−3 mol kg−1, and I = 0.23 mol kg−1) the apparent equilibrium constant (KA′ = m(ΣAMP)m(ΣATP)/[ m(ΣADP)]2) for the overall disproportionation reaction is equal to 0.93 ± 0.02. Thermodynamic data on the disproportionation reaction and literature values for this apparent equilibrium constant in human red blood cells are used to calculate a morality of 1.94 × 10−4 mol kg−1 for free magnesium ion in human red blood cells. The results are also discussed in relation to thermochemical cycles and compared with data on the hydrolysis of the guanosine phosphates.  相似文献   

4.
To study structure-activity relationships of growth hormone-releasing hormone (GHRH), a competitive binding assay was developed using cloned porcine adenopituitary GHRH receptors expressed in human kidney 293 cells. Specific binding of [His1,125I-Tyr10,Nle27]hGHRH(1–32)-NH2 increased linearly with protein concentration (10–45 μg protein/tube). Binding reached equilibrium after 90 min at 30°C and remained constant for at least 240 min. Binding was reversible to one class of high-affinity sites (Kd = 1.04 ± 0.19 nM, Bmax = 3.9 ± 0.53 pmol/mg protein). Binding was selective with a rank order of affinity (IC50) for porcine GHRH (2.8 ± 0.51 nM), rat GHRH (3.1 ± 0.69 nM), [N-Ac-Tyr1, -Arg2]hGHRH(3–29)-NH2 (3.9 ± 0.58 nM), and [ -Thr7]GHRH(1–29)-NH2 (189.7 ± 14.3 nM), consistent with their binding to a GHRH receptor. Nonhydrolyzable guanine nucleotides inhibited binding. These data describe a selective and reliable method for a competitive GHRH binding assay that for the first time utilizes rapid filtration to terminate the binding assay.  相似文献   

5.
The study on the interaction of artemisinin with bovine serum albumin (BSA) has been undertaken at three temperatures, 289, 296 and 303 K and investigated the effect of common ions and UV C (253.7 nm) irradiation on the binding of artemisinin with BSA. The binding mode, the binding constant and the protein structure changes in the presence of artemisinin in aqueous solution at pH 7.40 have been evaluated using fluorescence, UV–vis and Fourier transform infrared (FT-IR) spectroscopy. The quenching constant Kq, Ksv and the association constant K were calculated according to Stern–Volmer equation based on the quenching of the fluorescence of BSA. The thermodynamic parameters, the enthalpy (ΔH) and the entropy change (ΔS) were estimated to be −3.625 kJ mol−1 and 107.419 J mol−1 K−1 using the van’t Hoff equation. The displacement experiment shows that artemisinin can bind to the subdomain IIA. The distance between the tryptophan residues in BSA and artemisinin bound to site I was estimated to be 2.22 nm using Föster's equation on the basis of fluorescence energy transfer. The decreased binding constant in the presence of enough common ions and UV C exposure, indicates that common ions and UV C irradiation have effect on artemisinin binding to BSA.  相似文献   

6.
1. The reduction of cytochrome c oxidase by hydrated electrons was studied in the absence and presence of cytochrome c.

2. Hydrated electrons do not readily reduce the heme of cytochrome c oxidase. This observation supports our previous conclusion that heme a is not directly exposed to the solvent.

3. In a mixture of cytochrome c and cytochrome c oxidase, cytochrome c is first reduced by hydrated electrons (k = 4 · 1010 M−1 · s−1 at 22 °C and pH 7.2) after which it transfers electrons to cytochrome c oxidase with a rate constant of 6 · 107 M−1 · s−1 at 22 °C and pH 7.2.

4. It was found that two equivalents of cytochrome c are oxidized initially per equivalent of heme a reduced, showing that one electron is accepted by a second electron acceptor, probably one of the copper atoms of cytochrome c oxidase.

5. After the initial reduction, redistribution of electrons takes place until an equilibrium is reached similar to that found in redox experiments of Tiesjema, R. H., Muijsers, A. O. and Van Gelder, B. F. (1973) Biochim. Biophys. Acta 305, 19–28.  相似文献   


7.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


8.
Richard Maskiewicz  Benon H.J. Bielski   《BBA》1982,680(3):297-303
It has been shown by the pulse radiolysis technique that radiation-generated NADP free radicals (NADP·) first combine with ferredoxin-NADP reductase and then transfer the odd electron by a fast intramolecular process to the enzyme flavin moiety yielding the semiquinone (ferredoxin-NADP reductase, FNR-FADH·). The corresponding first-order rate constant k15 varies with ionic strength from 2.6·103 s−1 at I = 0.66 M to 2.3·104 s−1 at I = 0.005 M In the presence of ferredoxin-NADP reductase-bound oxidized ferredoxin, the electron cascades, thus further reducing the ferredoxin. The transfer of the electron from the flavin semiquinone (ferredoxin-NADP reductase, FNR-FADH·) to the bound oxidized ferredoxin proceeds at a rate of k18 = 2.36 s−1. This process approaches an equilibrium condition which is in favor of the reverse reaction suggesting that k−18 > k18.  相似文献   

9.
1. Fluoride is a mixed-type inhibitor of the cytochrome c oxidase activity with a Ki for the free enzyme of 10 mM and a Ki for the cytochrome c-complexed enzyme of 35 mM.

2. Fluoride shifts the γ-band of the enzyme from 423 to 421 nm and the -band from 597 to 598 nm. The difference spectrum (oxidized enzyme in the presence of fluoride minus oxidized enzyme) has peaks at 400, 453, 482, 605 and 638 nm and troughs at 430, 520, 552 and 674 nm. The changes in absorbance are small (about 3% at absorbance maxima) with respect to those of other hemoproteins.

3. On addition of fluoride to isolated cytochrome c oxidase 3 reactions can be distinguished: (I) a bimolecular binding reaction (Kon = 4 M−1 · s−1 and koff = 2.9 · 10−2s−1 at 25 °C, pH 7.4) contributing at 638 nm and 430 nm; (II) a first-order reaction (k = 2.4 · 10−2) s−1 at 22 °C, pH 7.2) visible mainly at 430 nm and (III) a very slow reaction with a half-time in the order of 10 min.

4. The spectroscopic dissociation constants for the fluoride binding, determined from Hill plots using the absorbance changes at 638 and 430 nm, are similar (7 and 10 mM, respectively, at 22 °C, pH 7.2).

5. A mechanism for the reaction is discussed in which the bimolecular binding reaction is followed by a conformational change of the enzyme-fluoride complex.  相似文献   


10.
Under physiological pH conditions (pH 7.2-7.4) the rate constant of the reaction NO + O2 yielding peroxonitrite (ONOO) was determined as k = (3.7 ± 1.1) × 107 M 1 s 1. The decay of peroxonitrite at this pH follows first order kinetics with a rate constant of 1.4 s 1. At alkaline pH peroxonitrite is practically stable.

Possible consequences of these reactions for the biological lifetime of EDRF will be discussed.  相似文献   

11.
Direct evidence obtained by means of the technique of pulse radiolysis-kinetic spectrometry, with measurements in the time range 10−6 to 1 s, is presented that, consequent upon reaction of a single H-atom with a single molecule of ferricytochrome c, a reducing equivalent is transmitted via the protein structure to the ferriheme moiety. Such transmission accounts for at least 70% of the total reduction of the ferri to the ferro state of cytochrome c. The remainder of the total reduction takes place without stages resolvable on the time scale of these experiments. Reduction brought about by H atoms appears to follow a different course than reduction by hydrated electrons. In the latter case, intramolecular transmission of reducing equivalents could not be demonstrated (Lichtin, N. N., Shafferman, A. and Stein, G. (1973) Biochim. Biophys. Acta 314, 117–135).

Not every H-atom reacts with ferricytochrome c at a site which results in conversion of the Fe(III) state to the Fe(II) state. Approximately half of reacting H-atoms do not produce reduction.

The following second order rate constants have been determined in solutions of low ionic strength at 20±2 °C: k[H+ferricytochrome c] = (1.0±0.2) · 1010 M−1 · s−1 at pH 3.0 and 6.7; k[H+ferrocytochrome c] = (1.3±0.2) · 1010 M−1 · s−1 at pH 3.0; k[eaq + ferrocytochrome c] = (1.9±0.4) · 1010 M−1 · s−1 at pH 6.7.  相似文献   


12.
Studies on the molecular chain morphology of konjac glucomannan   总被引:2,自引:0,他引:2  
The chain geometry and parameters of konjac glucomannan were studied by using laser light scatter (LLS), gel permeation chromatography (GPC) and viscosimetry. The weight-average molecular weight (Mw), root-mean-square ratio of gyration (S21/2), second viral coefficient (A2) and polydispersity index (Mw/Mn) were 1.036×106, 105±0.9 nm, (−1.587±0.283)×10−3  mol ml g−2 and 1.015±0.003 respectively. Mark-Houwink equation was established as , and the molecular chain parameters were as follows: ML=982.82 nm−1, q=27.93 nm, d=0.74 nm, h=0.26 nm, L=1054.11 nm. To confirm the above results, konjac glucomannan was observed by using atomic force microscopy (AFM) and transmission electron microscope (TEM). The physical image showed directly that the konjac glucomannan molecule was an extending semi-flexible linear chain without branches, and than the molecular dimension also conformed to the parameters above. Therefore the image of molecular chain geometry confirmed the deduction drawn by Mark-Houwink equation and molecular chain parameters magnificently.  相似文献   

13.
The stiffness of the sarcomeres was studied during the diastolic interval of 18 stimulated (0.5 Hz) cardiac trabeculae of rat (pH 7.4; temperature = 25°C). Sarcomere length (SL) and force (F) were measured using, respectively, laser diffraction techniques (resolution: 4 nm) and a silicon strain gauge (resolution: 0.63 μN). Sinusoidal perturbations (frequency = 500 Hz) were imposed to the length of the preparation. The stiffness was evaluated from the corresponding F and SL sinusoids by analysis of both signals together either in the time domain or in the frequency domain. A short burst (duration = 30 ms) of sinusoidal perturbations was repeated at 5 predetermined times during diastole providing 5 measurements of stiffness during the time interval separating two twitches. These measurements revealed that stiffness increases by 30% during diastole, while a simultaneous expansion of the sarcomeres (amplitude = 10-60 nm) was detected. Measurements of the fluorescence of fura-2 under the same conditions revealed a continuous exponential decline of [Ca2+]i from 210 to 90 nM (constant of time 300 ms) during diastole. In order to test the possibility that the increase of sarcomere stiffness and the decline of [Ca2+]i were coupled during diastole of intact trabeculae, we studied the effect of different free Ca2+-concentrations ([Ca2+]) between 1 and 430 nM on sarcomere stiffness in rat cardiac trabeculae skinned by saponin (n = 17). Stiffness was studied using 500 Hz sinusoidal perturbations of muscle length (ML). We found that, below 70 nM, the stiffness was independent of [Ca2+]; between 70 and 200 nM, the stiffness declined with increase of [Ca2+]; above 200 nM, the stiffness increased steeply with [Ca2+]. The data fitted accurately to the sum of two sigmoids (Hill functions): (1) at [Ca2+] < 200 nM the stiffness decreased with [Ca2+] (EC50 = 160 ± 13 nM; n = −2.6±0.7) and (2) at [Ca2+] > 200 nM, stiffness increased with [Ca2+] (EC50 = 3.4±0.3 μM; n = 2.1±0.2) due to attachment of cross-bridges. From these results, it was possible to reproduce accurately the time course of diastolic stiffness observed in intact trabeculae and to predict the effect on stiffness of a spontaneous elevation of the diastolic [Ca2+]. Identical stiffness measurements were performed in 4 skinned preparations exposed to a cloned fragment of titin (Ti I-II) which has been shown to exhibit a strong interaction with F-actin in vitro. It was anticipated that Ti I-II would compete with endogenous titin for the same binding site on actin in the I-band. Below 200 nM, Ti I-II (2 μM) eliminated the Ca2+-dependence of stiffness. These results are consistent with the hypothesis that the Ca2+-sensitivity of the sarcomeres at [Ca2+] < 200 nM, i.e. where the myocytes in intact muscle operate during diastole, involves an association between titin molecules and the thin filament.  相似文献   

14.
Vitamin A components, retinol and retinoic acid, are fat-soluble micronutrients and critical for many biological processes, including vision, reproduction, growth, and regulation of cell proliferation and differentiation. The cellular uptake of Vitamin A is through specific interaction of a plasma membrane receptor with serum retinol-binding protein. Human serum albumin (HSA), as a transport protein, is the major target of several micronutrients in vivo. The aim of present study was to examine the interaction of retinol and retinoic acid with human serum albumin in aqueous solution at physiological conditions using constant protein concentration and various retinoid contents. FTIR, UV–vis, CD and fluorescence spectroscopic methods were used to determine retinoid binding mode, the binding constant and the effects of complexation on protein secondary structure.

Structural analysis showed that retinol and retinoic acid bind non-specifically (H-bonding) via protein polar groups with binding constants of Kret = 1.32 (±0.30) × 105 M−1 and Kretac = 3.33 (±0.35) × 105 M−1. The protein secondary structure showed no alterations at low retinoid concentrations (0.125 mM), whereas at high retinoid content (1 mM), an increase of -helix from 55% (free HSA) to 60% and a decrease of β-sheet from 22% (free HSA) to 18% occurred in the retinoid–HSA complexes. The results point to a partial stabilization of protein secondary structure at high retinoid content.  相似文献   


15.
Horse-heart ferrocytochrome c has been labeled with N-(2,2,5,5-tetramethyl-3-pyrrolidinyl-1-oxyl) iodoacetamide at methionine-65. The paramagnetic resonance spectrum of labeled ferricytochrome c indicates a weak immobilization of the radical (τc = 9.3·10−10 sec) which becomes stronger upon binding of labeled cytochrome c to cytochrome c-depleted mitochondrial membranes (τc = 3.3·10−9 sec). The hyperfine coupling constant remains, however, unchanged (16.7 ± 0.1 gauss) indicating that the cytochrome c binding site is highly polar. The region where cytochrome c is bound to the membrane is insensitive to large variations of medium viscosity.  相似文献   

16.
George D. Case  William W. Parson   《BBA》1973,292(3):677-684
The isoionic pH of Chromatium chromatophores is 5.2±0.1. At pH 7.7, the net charge on the chromatophore is approx. −1·104. If a change in this charge accompanies the oxidation of an electron carrier, the midpoint redox potential (Em) of that carrier should be a function of the solution ionic strength (I). of that carrier should be a function of the solution ionic strength (I).

The Em values of P870 and cytochrome c-555 increase strongly with increasing I at low values of I. The Em of cytochrome c-552 also increases with increasing I, though not so strongly. These effects probably cannot be attributed to an influence of I on the activity coefficient of a dissociable ion. We conclude that, when either P870 or cytochrome c-555 loses an electron, no specific ions (including protons) are bound or released in significant amounts, and the absolute value of the charge on the chromatophore decreases.

The Em values of the primary and secondary electron acceptors, X and Y, do not depend on I. Because these Em values have been shown previously to depend on pH, we conclude that the uptake of a proton keeps the charge on the chromatophore constant when either X or Y accepts an electron. This means that the primary and secondary electron transfer reactions in Chromatium result in a net decrease in the charge on the photosynthetic membrane. They do not result in the translocation of protons across the membrane.

The Em of the soluble flavocytochrome c-552 from Chromatium depends only weakly on I, but depends strongly on the pH. The uptake of a proton appears to keep the net charge on this cytochrome constant upon reduction.  相似文献   


17.
A. Mü  hlrad  K. Ajtai  F. F  bi  n 《BBA》1970,205(3):342-354
The specificity and nature of the reaction between salicylaldehyde and myosin and the effect of salicylalation on the molecular parameters of myosin were studied. The following observations were made.

1. 1. The reaction of salicylaldehyde with the lysyl residues of myosin is specific, since no salicylaldehyde is bound if the lysyl residues of myosin are trinitrophenylated.

2. 2. Salicylaldehyde is bound by myosin through the formation of an azomethine linkage (Schiff's base). This was established from the measured difference absorption spectrum of the myosin-salicylaldehyde complex.

3. 3. Three groups of lysyl residues can be distinguished with respect to the reaction with salicylaldehyde, namely, (a) residues with high association constant (Kass = 1.8 ± 0.9·105 M-1), (b) residues with moderate association constant (Kass = 2.2·103 M-1) and (c) residues that react with salicylaldehyde only after the denaturation of the protein. Their numbers could be estimated as 10 ± 5, 130 ± 5 and 260 ± 5 per mole myosin, respectively. The first group of residues was found to be absent from heavy and light meromyosin, the proteolytic fragments of myosin.

4. 4. The reaction is reversible. The complex formation rate constant, evaluated from the formula for second order reaction, is 2.2 sec-1·M-1, and the decomposition rate constant for first order reaction is 1.1·10-3 sec-1 at 22°.

5. 5. The reaction is pH dependent, the reaction yield increasing at higher pH.

6. 6. The solubility of myosin at low ionic strength decreases with increasing degree of salicylalation at slightly alkaline pH.

7. 7. The intrinsic viscosity of myosin does not change on salicylalation.

8. 8. A second peak due to polymerization appears on the sedimentation profile of the protein if more than 70 lysyl residues are salicylalated per mole of myosin.

Abbreviations: TBS, 2,4,6-trinitrobenzenesulphonate  相似文献   


18.
A Bacillus subtilis strain isolated from a hot-spring was shown to produce xylanolytic enzymes. Their associative/synergistic effect was studied using a culture medium with oat spelts xylan as xylanase inducer. Optimal xylanase production of about 12 U ml−1 was achieved at pH 6.0 and 50°C, within 18 h fermentation. At 50°C, xylanase productivity obtained after 11 h in shake-flasks, 96,000 U l−1 h−1, and in reactor, 104,000 U l−1 h−1 was similar. Increasing temperature to 55°C a higher productivity was obtained in the batch reactor 45,000 U l−1 h−1, compared to shake-flask fermentations, 12,000 U l−1 h−1. Optimal xylanolytic activity was reached at 60°C on phosphate buffer, at pH 6.0. The xylanase is thermostable, presenting full stability at 60°C during 3 h. Further increase in the temperature caused a correspondent decrease in the residual activity. At 90°C, 20% relative activity remains after 14 min. Under optimised fermentation conditions, no cellulolytic activity was detected on the extract. Protein disulphide reducing agents, such as DTT, enhanced xylanolytic activity about 2.5-fold. When is used xylan as substrate, xylanase production decreased as function of time in contrast, with trehalose as carbon source, xylanase production in maintained constant for at least 80 h fermentation.  相似文献   

19.
Thor Arnason  John Sinclair 《BBA》1976,430(3):517-523
The modulated oxygen polarograph has been used to study the rate-determining steps of photosynthetic oxygen evolution in spinach chloroplasts. The rate constant, k, of the reaction has a value of 218±10 (S.E.) s−1 at 23 °C and an activation energy of 7±2 (S.E.) kcal · mol−1. A kinetic isotope experiment indicated that this step is probably not the water-splitting reaction. These findings resemble previous results with the unicellular alga Chlorella (Sinclair, J. and Arnason, T. (1974) Biochim. Biophys. Acta 368, 393–400). In other experiments we changed the pH, O2 concentration and osmolarity of the medium, and treated the chloroplasts with 1 mM NH4Cl without detecting any significant change in k. These results suggest that the step is irreversible. However, a significantly lower value of k, 110±20 (S.E.) s−1 was obtained when all salts except 1 mM MgCl2 were removed from the medium bathing the chloroplasts.  相似文献   

20.
A thermostable, alkaline active xylanase was purified to homogeneity from the culture supernatant of an alkaliphilic Bacillus halodurans S7, which was isolated from a soda lake in the Ethiopian Rift Valley. The molecular weight and the pI of this enzyme were estimated to be around 43 kDa and 4.5, respectively. When assayed at 70 °C, it was optimally active at pH 9.0–9.5. The optimum temperature for the activity was 75 °C at pH 9 and 70 °C at pH 10. The enzyme was stable over a broad pH range and showed good thermal stability when incubated at 65 °C in pH 9 buffer. The enzyme activity was strongly inhibited by Mn2+. Partial inhibition was also observed in the presence of 5 mM Cu2+, Co2+ and EDTA. Inhibition by Hg2+ and dithiothreitol was insignificant. The enzyme was free from cellulase activity and degraded xylan in an endo-fashion.  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