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.
A procedure for the preparation of the fully reduced Cu(I) form of galactose oxidase, GOase(red), involving reduction of GOase(semi) (or GOase(ox)) with non-coordinating [Ru(NH(3))(6)](2+) (51 mV vs. nhe) is described. Air-free conditions and a two-fold excess of [Ru(NH(3))(6)](2+) give a stable product with no further UV-Vis changes over >1.5 h. Rate constants for the reduction of GOase(semi) (k(f)=860 M(-1) s(-1)) give a first-order [H(+)]-dependence (pK(1a)=7.9), but the reverse process involving [Ru(NH(3))(6)](3+) oxidation of GOase(red) (k(b)=18.6 M(-1) s(-1)) is independent of pH (5.5 to 9.5). The reduction potential E(2)(o)' (vs. nhe) for the GOase(semi)/GOase(red) (i.e. Cu(II)/Cu(I)) couple is 149 mV at pH 7.5, which varies from 160 mV (pH 5.5) to 120 mV (pH 10.5), suggesting pK(1a) (GOase(semi)) and pK(2a) (GOase(red)) acid dissociation constants both involving Tyr-495. It is concluded that pK(2a) is for acid dissociation of uncoordinated H(+)Tyr-495. Consistent with this interpretation rate constants/M(-1) s(-1) for the GOase(semi) Tyr495 Phe variant, k(f)=1.59x10(3) and k(b)=16.1, respectively, are independent of pH and give a reduction potential of 169 mV. Comparisons are made of reduction potentials (E(1)(o)'/mV pH 7.5) for the GOase(ox)/GOase(semi) (i.e. Tyr(.)/Tyr) couple, and are for the Cys228Gly variant (630), for enzyme with N(3)(-) for H(2)O at the substrate binding exogenous site (393), and for apo-protein (570). These compare with previously reported values for the variants Trp290His (730) and Tyr495Phe (450), and together serve to quantify different contributions to the unusually small E(1)(o)' of 400 mV for the Tyr(.)/Tyr couple. At pH 7.5 the reduction potential for the two-equivalent GOase(ox)/GOase(red) couple is calculated to be 275 mV. The rate constant for the reaction of GOase(red) with GOase(ox) is 4.4x10(3) M(-1) s(-1) at pH 7.5.  相似文献   

2.
Catechol 2,3-dioxygenase [EC 1.13.11.2] from Pseudomonas putida mt-2 (Mpc) catalyzes the extradiol cleavage of catechol to produce 2-hydroxymuconate semialdehyde. The K(m) values for the catecholic substrate (K(mA)) and O(2) (K(mO2)), and catalytic constants (k(cat)) were kinetically determined for eight C3/C4-substituted catechols at 25 degrees C and pH 6.5 or 7.5. The first pK(a) values (pK(1)) were determined for eleven catechols (pK(1) = 7.26-9.47), correlated with Hammett substituent constants, and electron-withdrawing substituents significantly stabilized the monoanionic species of free catechols. Mpc preferred catechols with non-ionic substituents at the C3 or C4 position. 3-Phenylcatechol, a biphenyl, was cleaved, while 4-tert-butylcatechol was not. The logarithm of k(cat)/K(mA) (substrate specificity constant) exhibited a good linear correlation with pK(1), with the exception of those for 4-halocatechols. The logarithm of k(cat)/K(mO2) showed a good linear correlation with pK(1), with the exception of that of 3-phenylcatechol. These results demonstrate that catechol binding to the Mpc active site, the following O(2) binding, and the activation of the bound O(2) are all sensitive to electronic effects of the substituents. However, k(cat) did not correlate significantly with pK(1). The present study distinguishes clearly between the electronic and the steric effects of catecholic substrates in the reactivity of Mpc, and provides important insight into the mechanistic basis for a vast range of substrate specificities of extradiol dioxygenases.  相似文献   

3.
Human matrix metalloproteinase 7 (MMP-7) exhibits a broad bell-shaped pH-dependence with the acidic and alkaline pK(e) (pK(e1) and pK(e2)) values of about 4 and 10. In this study, we estimated the ionizable groups involved in its catalytic mechanism by thermodynamic analysis. pK(a) of side chains of L-Asp, L-Glu, L-His, L-Cys, L-Tyr, L-Lys, and L-Arg at 25-45°C were determined by the pH titration of amino-acid solutions, from which their enthalpy changes, ?H°, of deprotonation were calculated. pK(e1) and pK(e2) of MMP-7 at 15-45°C were determined in the hydrolysis of (7-methoxycoumarin-4-yl)acetyl-L-Pro-L-Leu-Gly-L-Leu-[N(3)-(2,4-dinitrophenyl)-L-2,3-diaminopropionyl]-L-Ala-L-Arg-NH(2), from which ?H(o) for pK(e1) and pK(e2) was calculated. The ?H(o) for pK(e1) (-20.6±6.1kJmol(-1)) was similar to that for L-Glu (-23.6±5.8kJmol(-1)), and the ?H(o) for pK(e2) (89.9±4.0kJmol(-1)) was similar to those for L-Arg (87.6±5.5kJmol(-1)) and L-Lys (70.4±4.4kJmol(-1)). The mutation of the active-site residue Glu198 into Ala completely abolished the activity, suggesting that Glu198 is the ionizable group for pK(e1). On the other hand, no arginine or lysine residues are found in the active site of MMP-7. We proposed a possibility that a protein-bound water is the ionizable group for pK(e2).  相似文献   

4.
The effects of pH on the yield (phi(r)), and on the apparent rise and decay constants (k(r), k(d)), of the O(630) intermediate are important features of the bacteriorhodopsin (bR) photocycle. The effects are associated with three titration-like transitions: 1) A drop in k(r), k(d), and phi(r) at high pH [pK(a)(1) approximately 8]; 2) A rise in phi(r) at low pH [pK(a)(2) approximately 4.5]; and 3) A drop in k(r) and k(d) at low pH [pK(a)(3) approximately 4. 5]. (pK(a) values are for native bR in 100 mM NaCl). Clarification of these effects is approached by studying the pH dependence of phi(r), k(r), and k(d) in native and acetylated bR, and in its D96N and R82Q mutants. The D96N experiments were carried out in the presence of small amounts of the weak acids, azide, nitrite, and thiocyanate. Analysis of the mutant's data leads to the identification of the protein residue (R(1)) whose state of protonation controls the magnitude of phi(r), k(r), and k(d) at high pH, as Asp-96. Acetylation of bR modifies the Lys-129 residue, which is known to affect the pK(a) of the group (XH), which releases the proton to the membrane exterior during the photocycle. The effects of acetylation on the O(630) parameters reveal that the low-pH titrations should be ascribed to two additional protein residues R(2) and R(3). R(2) affects the rise of phi(r) at low pH, whereas the state of protonation of R(3) affects both k(r) and k(d). Our data confirm a previous suggestion that R(3) should be identified as the proton release moiety (XH). A clear identification of R(2), including its possible identity with R(3), remains open.  相似文献   

5.
Most of the esterase properties of human serum albumin (HSA) are the result of multiple irreversible chemical modifications rather than turnover. The HSA-catalyzed hydrolysis of 4-nitrophenyl myristate (NphOMy) is consistent with the minimum three-step mechanism involving the acyl-enzyme intermediate HSA-OMy: Under all the experimental conditions, values of K(s) (= k(-1)/k(+1)), k(+2), and k(+2)/K(s) determined under conditions where [HSA] ≥ 5 × [NphOMy] and [NphOMy] ≥ 5 × [HSA] match very well each other. The deacylation process is rate limiting in catalysis (i.e., k(+3) < k(+2)) and k(-2)~k(-3)~0 s(-1). The pH dependence of k(+2)/K(s), k(+2), and K(s) reflects the acidic pK(a)-shift of one ionizing group from 8.9 ± 0.2 in NphOMy-free HSA to 6.8 ± 0.3 in the HSA:NphOMy adduct. The HSA-catalyzed hydrolysis of NphOMy is inhibited competitively by diazepam, indicating that Tyr411 is the active-site nucleophile.  相似文献   

6.
Alcohol dehydrogenase was purified in 14 h from male Fischer-344 rat livers by differential centrifugation, (NH4)2SO4 precipitation, and chromatography over DEAE-Affi-Gel Blue, Affi-Gel Blue, and AMP-agarose. Following HPLC more than 240-fold purification was obtained. Under denaturing conditions, the enzyme migrated as a single protein band (Mr congruent to 40,000) on 10% sodium dodecyl sulfate-polyacrylamide gels. Under nondenaturing conditions, the protein eluted from an HPLC I-125 column as a symmetrical peak with a constant enzyme specific activity. When examined by analytical isoelectric focusing, two protein and two enzyme activity bands comigrated closely together (broad band) between pH 8.8 and 8.9. The pure enzyme showed pH optima for activity between 8.3 and 8.8 in buffers of 0.5 M Tris-HCl, 50 mM 2-(N-cyclohexylamino)ethanesulfonic acid (CHES), and 50 mM 3-(cyclohexylamino)-1-propanesulfonic acid (CAPS), and above pH 9.0 in 50 mM glycyl-glycine. Kinetic studies with the pure enzyme, in 0.5 M Tris-HCl under varying pH conditions, revealed three characteristic ionization constants for activity: 7.4 (pK1); 8.0-8.1 (pK2), and 9.1 (pK3). The latter two probably represent functional groups in the free enzyme; pK1 may represent a functional group in the enzyme-NAD+ complex. Pure enzyme also was used to determine kinetic constants at 37 degrees C in 0.5 M Tris-HCl buffer, pH 7.4 (I = 0.2). The values obtained were Vmax = 2.21 microM/min/mg enzyme, Km for ethanol = 0.156 mM, Km for NAD+ = 0.176 mM, and a dissociation constant for NAD+ = 0.306 mM. These values were used to extrapolate the forward rate of ethanol oxidation by alcohol dehydrogenase in vivo. At pH 7.4 and 10 mM ethanol, the rate was calculated to be 2.4 microM/min/g liver.  相似文献   

7.
Gasoline-contaminated soil from Isfahan, Iran was selected to isolate a bacterium capable of desulfurizing dibenzothiophene (DBT). The isolated strain was named R1 and identified as Rhodococcus erythropolis through biochemical tests as well as sequencing of 16S rRNA gene. This strain could efficiently produce 2-hydroxybiphenyl (HBP) from DBT via the 4S metabolic pathway. The highest HBP amount was produced at 2 mM DBT with addition of glucose (10 g l(-1)), ethanol (3 g l(-1)), glycerol (2 g l(-1)) or succinate (10 g l(-1)) as carbon sources at pH 7. Highest respiration and growth rates were observed by microplate titration on 0.1 mM HBP, and addition of 0.2 mM HBP to glucose (1 g l(-1)) and DBT (0.3 mM) could inhibite the respiration of the isolate. The isolated strain could grow up to 0.4 mM of HBP when it is used with mineral sulfur as sole sulfur source. To the best of our knowledge this is the first report on a microtiter assay for the production and utilization of HBP by Rhodococcus.  相似文献   

8.
Reaction of rat liver glutathione S-transferase, isozyme 1-1, with 4-(fluorosulfonyl)benzoic acid (4-FSB), a xenobiotic substrate analogue, results in a time-dependent inactivation of the enzyme to a final value of 35% of its original activity when assayed at pH 6.5 with 1-chloro-2,4-dinitrobenzene (CDNB) as substrate. The rate of inactivation exhibits a nonlinear dependence on the concentration of 4-FSB from 0.25 mM to 9 mM, characterized by a KI of 0.78 mM and kmax of 0.011 min-1. S-Hexylglutathione or the xenobiotic substrate analogue, 2,4-dinitrophenol, protects against inactivation of the enzyme by 4-FSB, whereas S-methylglutathione has little effect on the reaction. These experiments indicate that reaction occurs within the active site of the enzyme, probably in the binding site of the xenobiotic substrate, close to the glutathione binding site. Incorporation of [3,5-3H]-4-FSB into the enzyme in the absence and presence of S-hexylglutathione suggests that modification of one residue is responsible for the partial loss of enzyme activity. Tyr 8 and Cys 17 are shown to be the reaction targets of 4-FSB, but only Tyr 8 is protected against 4-FSB by S-hexylglutathione. DTT regenerates cysteine from the reaction product of cysteine and 4-FSB, but does not reactivate the enzyme. These results show that modification of Tyr 8 by 4-FSB causes the partial inactivation of the enzyme. The Michaelis constants for various substrates are not changed by the modification of the enzyme. The pH dependence of the enzyme-catalyzed reaction of glutathione with CDNB for the modified enzyme, as compared with the native enzyme, reveals an increase of about 0.9 in the apparent pKa, which has been interpreted as representing the ionization of enzyme-bound glutathione; however, this pKa of about 7.4 for modified enzyme remains far below the pK of 9.1 for the -SH of free glutathione. Previously, it was considered that Tyr 8 was essential for GST catalysis. In contrast, we conclude that Tyr 8 facilitates the ionization of the thiol group of glutathione bound to glutathione S-transferase, but is not required for enzyme activity.  相似文献   

9.
To elucidate a detailed catalytic mechanism for nitrile hydratases (NHases), the pH and temperature dependence of the kinetic constants k(cat) and K(m) for the cobalt-type NHase from Pseudonocardia thermophila JCM 3095 (PtNHase) were examined. PtNHase was found to exhibit a bell-shaped curve for plots of relative activity versus pH at pH 3.2-11 and was found to display maximal activity between pH 7.2 and 7.8. Fits of these data provided pK(E)(S1) and pK(E)(S2) values of 5.9 +/- 0.1 and 9.2 +/- 0.1 (k(cat)' = 130 +/- 1 s(-1)), respectively, and pK(E)(1) and pK(E)(2) values of 5.8 +/- 0.1 and 9.1 +/- 0.1 (k(cat)'/K(m)' = (6.5 +/- 0.1) x 10(3) s(-1) mm(-1)), respectively. Proton inventory studies indicated that two protons are transferred in the rate-limiting step of the reaction at pH 7.6. Because PtNHase is stable at 60 degrees C, an Arrhenius plot was constructed by plotting ln(k(cat)) versus 1/T, providing E(a) = 23.0 +/- 1.2 kJ/mol. The thermal stability of PtNHase also allowed DeltaH(0) ionization values to be determined, thus helping to identify the ionizing groups exhibiting the pK(E)(S1) and pK(E)(S2) values. Based on DeltaH(0)(ion) data, pK(E)(S1) is assigned to betaTyr(68), whereas pK(E)(S2) is assigned to betaArg(52), betaArg(157), or alphaSer(112) (NHases are alpha(2)beta(2)-heterotetramers). A combination of these data with those previously reported for NHases and synthetic model complexes, along with sequence comparisons of both iron- and cobalt-type NHases, allowed a novel catalytic mechanism for NHases to be proposed.  相似文献   

10.
A role for redox-active tyrosines has been demonstrated in many important biological processes, including water oxidation carried out by photosystem II (PSII) of oxygenic photosynthesis. The rates of tyrosine oxidation and reduction and the Tyr/Tyr reduction potential are undoubtedly controlled by the immediate environment of the tyrosine, with the coupling of electron and proton transfer, a critical component of the kinetic and redox behavior. It has been demonstrated by Faller et al. that the rate of oxidation of tyrosine D (Tyr(D)) at room temperature and the extent of Tyr(D) oxidation at cryogenic temperatures, following flash excitation, dramatically increase as a function of pH with a pK(a) of approximately 7.6 [Faller et al. 2001 Proc. Natl. Acad. Sci. USA 98, 14368-14373; Faller et al. 2001 Biochemistry 41, 12914-12920]. In this work, we investigated, using FTIR difference spectroscopy, the mechanistic reasons behind this large pH dependence. These studies were carried out on Mn-depleted PSII core complexes isolated from Synechocystis sp. PCC 6803, WT unlabeled and labeled with (13)C(6)-, or (13)C(1)(4)-labeled tyrosine, as well as on the D2-Gln164Glu mutant. The main conclusions of this work are that the pH-induced changes involve the reduced Tyr(D) state and not the oxidized Tyr(D)() state and that Tyr(D) does not exist in the tyrosinate form between pH 6 and 10. We can also exclude a change in the protonation state of D2-His189 as being responsible for the large pH dependence of Tyr(D) oxidation. Indeed, our data are consistent with D2-His189 being neutral both in the Tyr(D) and Tyr(D)() states in the whole pH6-10 range. We show that the interactions between reduced Tyr(D) and D2-His189 are modulated by the pH. At pH greater than 7.5, the nu(CO) mode frequency of Tyr(D) indicates that Tyr(D) is involved in a strong hydrogen bond, as a hydrogen bond donor only, in a fraction of the PSII centers. At pH below 7.5, the hydrogen-bonding interaction formed by Tyr(D) is weaker and Tyr(D) could be also involved as a hydrogen bond acceptor, according to calculations performed by Takahashi and Noguchi [J. Phys. Chem. B 2007 111, 13833-13844]. The involvement of Tyr(D) in this strong hydrogen-bonding interaction correlates with the ability to oxidize Tyr(D) at cryogenic temperatures and rapidly at room temperature. A strong hydrogen-bonding interaction is also observed at pH 6 in the D2-Gln164Glu mutant, showing that the residue at position D2-164 regulates the properties of Tyr(D.) The IR data point to the role of a protonatable group(s) (with a pK(a) of approximately 7) other than D2-His189 and Tyr(D), in modifying the characteristics of the Tyr(D) hydrogen-bonding interactions, and hence its oxidation properties. It remains to be determined whether the strong hydrogen-bonding interaction involves D2-His189 and if Tyr(D) oxidation involves the same proton transfer route at low and at high pH.  相似文献   

11.
Human matrix metalloproteinase 7 (MMP-7) exhibits a broad bell-shaped pH-dependence with the acidic and alkaline pK(e) (pK(e1) and pK(e2)) values of about 4 and 10. Its active-site tyrosyl residue, Tyr219, is conserved in all other MMPs, and thus has been thought for the ionizable group responsible for pK(e2). In this study, we examined the mutational effects of Tyr219 on enzyme activity. Five Tyr219 variants, Y219F (Tyr219 is replaced with Phe), Y219D, Y219A, Y219C and Y219S, were constructed by site-directed mutagenesis. In the hydrolysis of (7-methoxycoumarin-4-yl)acetyl-l-Pro-l-Leu-Gly-l-Leu-[N(3)-(2,4-dinitrophenyl)-l-2,3-diaminopropionyl]-l-Ala-l-Arg-NH(2), all five variants retained the activity, indicating that Tyr219 is not the ionizable group responsible for pK(e2). Unexpectedly, all five variants exhibited narrower pH-dependence than the wild-type MMP-7, with the pK(e1) and pK(e2) values in the range of 5.2-5.4 and 8.6-9.4, respectively. Such pH-dependence shifts were not observed in other active-site tyrosyl-residue variants, Y193F and Y216F. These results suggest that Tyr219 is not critical for catalytic activity, but is involved in the broad pH-dependence of the activity.  相似文献   

12.
N-Methyltryptophan oxidase (MTOX) is a flavoenzyme that catalyzes the oxidative demethylation of N-methyl-L-tryptophan and other N-methyl amino acids, including sarcosine, which is a poor substrate. The Escherichia coli gene encoding MTOX (solA) was isolated on the basis of its sequence homology with monomeric sarcosine oxidase, a sarcosine-inducible enzyme found in many bacteria. These studies show that MTOX is expressed as a constitutive enzyme in a wild-type E. coli K-12 strain, providing the first evidence that solA is a functional gene. MTOX expression is enhanced 3-fold by growth on minimal media but not induced by N-methyl-L-tryptophan, L-tryptophan, or 3-indoleacrylate. MTOX forms an anionic flavin semiquinone and a reversible, covalent flavin-sulfite complex (K(d) = 1.7 mM), properties characteristic of flavoprotein oxidases. Rates of formation (k(on) = 5.4 x 10(-3) M(-1) s(-1)) and dissociation (k(off) = 1.3 x 10(-5) s(-1)) of the MTOX-sulfite complex are orders of magnitude slower than observed with most other flavoprotein oxidases. The pK(a) for ionization of oxidized FAD at N(3)H in MTOX (8.36) is two pH units lower than that observed for free FAD. The MTOX active site was probed by characterization of various substrate analogues that act as competitive inhibitors with respect to N-methyl-L-tryptophan. Qualitatively similar perturbations of the MTOX visible absorption spectrum are observed for complexes formed with various aromatic carboxylates, including benzoate, 3-indole-(CH(2))(n)-CO(2)(-) and 2-indole-CO(2)(-). The most stable complex with 3-indole-(CH(2))(n)-CO(2)(-) is formed with 3-indolepropionate (K(d) = 0.79 mM), a derivative with the same side chain length as N-methyl-L-tryptophan. Benzoate binding is enhanced upon protonation of a group in the enzyme-benzoate complex (pK(EL) = 6.87) but blocked by ionization of a group in the free enzyme (pK(E) = 8.41), which is attributed to N(3)H of FAD. Difference spectra observed for the aromatic carboxylate complexes are virtually mirror images of those observed with sarcosine analogues (N,N'-dimethylglycine, N-benzylglycine). Charge-transfer complexes are formed with 3-indoleacrylate, pyrrole-2-carboxylate, and CH(3)XCH(2)CO(2)(-) (X = S, Se, Te).  相似文献   

13.
Heterotetrameric sarcosine oxidase is a flavoprotein that catalyses the oxidative demethylation of sarcosine. It is thought that the dehydrogenated substrate is the anionic form of sarcosine. To verify this assumption, the rate of flavin-adenine dinucleotide (FAD) reduction (k(red)) was analysed using protiated and deuterated sarcosine (N-methyl-d(3)-Gly) at various pH values using stopped-flow method. By increasing the pH from 6.2 to 9.8, k(red) increased for both substrates and reached a plateau, but the pK(a) value (reflecting the ionization of the enzyme-substrate complex) was 6.8 and 7.1 for protiated and deuterated sarcosine, respectively, and the kinetic isotope effect of k(red) decreased from approximately 19 to 8, indicating deprotonation of the bound sarcosine. The k(red)/K(d) (K(d), sarcosine dissociation constant) increased with increasing pH and reached a plateau. The pK (reflecting the ionization of free enzyme or free sarcosine) was 7.0 for both substrates, suggesting deprotonation of the βLys358 residue, which has a pK(a) of 6.7, as the pK(a) of the free sarcosine amine proton was determined to be approximately 10.1. These results indicate that the amine proton of sarcosine is transferred to the unprotonated Lys residue in the enzyme-substrate complex.  相似文献   

14.
Bovine erythrocyte Cu,Zn-superoxide dismutase (BESOD) is a dimeric enzyme composed of identical subunits associated through unusually strong non-covalent interactions. The state of the unique tyrosyl residue (Tyr 108) of BESOD was examined, and the kinetics of subunit exchange was studied using Tyr 108 as a probe. UV-absorption difference spectra demonstrate that Tyr 108 is exposed to the solvent, and that the accessibilities to ethanol, ethylene glycol, and polyethylene glycol 600 are 53.5, 52.0, and 44.6%, respectively. Tyr 108 was fully nitrated by tetranitromethane. The pK(a) values of the phenolic hydroxyl group of native and nitrated Tyr 108 were determined to be 11.3 and 7.9, whereas those of model compounds, L-tyrosine and 3-nitro-L-tyrosine, are 9.8 and 6.8, respectively. When the native and nitrated BESODs of an equal concentration were mixed, the hybrid dimer was formed. From the amount of hybrid dimer formed, the rate constant (k(-1)) of the subunit dissociation at pH 7.8, 25 degrees C was assessed to be (4.17 +/- 0.13) x 10(-6) s(-1). The activation energy of the subunit dissociation at pH 7.8 was determined to be 117 +/- 4 kJ.mol(-1). The k(-1) value remains constant at BESOD concentrations ranging from 0.62 to 6.8 micro M and at pH ranging from 6.0 to 10.0, but increased remarkably with a decrease in the dielectric constant of the reaction mixture. It is suggested that hydrophobic interaction may play a significant role in the subunit interaction.  相似文献   

15.
The difference absorption spectra of hen and turkey lysozymes in the alkaline pH region had three maxima at around 245, 292, and 300 nm and had no isosbestic points. The ratio of the extinction difference at 245 nm to that at 295 nm changed with pH. These spectral features are quite different from those observed when only tyrosyl residues are ionized, and it was impossible to determine precisely the pK values of the tyrosyl residues in lysozyme by spectrophotometric titration. A time-dependent spectral change was observed above about pH 12. This is not due to exposure of a buried tyrosyl residue on alkali denaturation. The disulfide bonds and the peptide bonds in the lysozyme molecule were cleaved by alkali above about pH 11. The intrinsic pK value of Tyr 23 of hen lysozyme was determined to be 10.24 (apparent pK 9.8) at 0.1 ionic strength and 25 degrees C from the CD titration data. Comparison of the CD titration of turkey lysozyme with that of hen lysozyme suggested that Tyr 3 and Tyr 23 in turkey lysozyme have apparent pK values of 11.9 and 9.8, respectively.  相似文献   

16.
The fluorescence and excitation spectra of luliberin (luteinizing hormone-releasing factor) in 0.005 M aqueous ammonium acetate are identical in shape to those of N-acetyltryptophan amide and are related to the indole side chain of Trp3. The change of fluoresecence intensity of luliberin with pH was measured in the range of pH 4-11. The increase of pH from 4 to 7.5 is followed by about 50% increase in fluorescence intensity due to deprotonation of the imidazolium side chain of His2. The fluorimetric titration curve in this pH region reveals a pK value for His2 of 5.95. Increasing of pH from 8 to 11 results in about 40% quenching of the fluorescence due to electronic energy transfer from the excited indole of Trp3 to the phenolate side chain of Tyr5. The pK value of Tyr5, obtained independently from the fluorimetric and photometric titrations indicate that at pH 7-8 luliberin contains only one charged residue, Arg8, which is in close vicinity to both His2 and Tyr5. The side chains of His2, Tyr5 and Arg8 presumably form a combined unit which may play an active role in the hormone action. Trp3 is at a maximal distance from this unit and may thus act as an independent active unit.  相似文献   

17.
Paper electrophoresis has been used over the pH range 1.2 to 10.4 to measure apparent pK(a) values for malvidin-3-O-glucoside of pK(a(1)) 1.76+/-0.07, pK(a(2)) 5.36+/-0.04, and pK(a(3)) 8.39+/-0.07. Using solvent partitioning between buffered aqueous solutions and n-octanol, several micro-pK(a) constants for malvidin-3-O-glucoside were also identified, highlighting the complex nature of malvidin-3-glucoside equilibria. As a nonspectrophotometric procedure, the charge-dependent electrophoretic mobility method provided independent information on the net charge and color of anthocyanin species at wine pH (ca. 3.6). At this pH, the color of malvidin-3-glucoside in red wines is consistent only with the uncharged quinonoidal base as a major colored component of the equilibria.  相似文献   

18.
A soluble F(1)-ATPase was isolated from the mitochondria of crayfish (Orconectes virilis) gill tissue. The maximal mitochondrial disruption rate (95%) was obtained by sonicating for 4 min at pH 8.6. A 15-fold purification was estimated. The properties for both soluble and membrane-bound enzyme were studied. Both enzyme forms were stable at 4 to -70 degrees C when kept in 20% glycerol. Soluble F(1)-ATPase was more stable at room temperature than membrane-bound enzyme. It displayed a narrower pH profile (pK(1) =6.58, pK(2)=7.68) and more acid pH optimum (7.13) than membrane-bound enzyme (pK(1)=6.42, pK(2)=8.55, optimum pH 7.49). The anion-stimulated activities were in the order HCO(3)(-)>SO(4)(2-)>Cl(-). The apparent K(a) values for soluble enzyme were 11.4, 11.2, and 10.9 mM, respectively, but the K(a) of HCO(3)(-) for membrane-bound enzyme (14.9 mM) was higher than for soluble enzyme. Oligomycin and DCCD inhibited membrane-bound F(1)-ATPase with I(50) of 18.6 ng/ml and 2.2 microM, respectively, but were ineffective in inhibiting soluble enzyme. Both enzyme forms shared identical sensitivity to DIDS (I(50)=12.5 microM) and vanadate (I(50)=9.0 mM). Soluble ATPase was significantly more sensitive to pCMB (I(50)=0.15 microM) and NO(3)(-) (I(50)=28.6 mM) than membrane-bound enzyme (I(50)=1.04 microM pCMB and 81.5 mM NO(3)(-)). In addition, soluble F(1)-ATPase was slightly more sensitive to azide (I(50)=91.8 microM) and NBD-Cl (I(50)=9.18 microM) than membrane-bound enzyme (I(50)=111.6 microM azide and 12.88 microM NBD-Cl). These data suggest a conformational change transmission between F(0) and F(1) sectors and slight conformational differences between soluble F(1) and membrane-bound F(1). In addition, an unmodified F(0) stabilizes F(1) and decreases F(1) sensitivities to inhibitors and modulators.  相似文献   

19.
Porter DJ  Short SA 《Biochemistry》2000,39(38):11788-11800
The catalytically active form of monofunctional yeast orotidine-5'-phosphate decarboxylase was a dimer (E(2)). The dimer equilibrium dissociation constant was 0.25 microM in 0.01 M MOPS Na(+) at pH 7.2. The bimolecular rate constant for dimer formation was 1.56 microM(-1) s(-1). The dimeric form of the enzyme was stabilized by NaCl such that the enzyme was E(2) in 100 mM NaCl at all concentrations of enzyme tested. The kinetics of binding of OMP to E(2) was governed by two ionizations (pK(1) = 6.1 and pK(2) = 7.7). From studies with substrate analogues, the higher pK was assigned to a group on the enzyme that interacted with the pyrimidinyl moiety. The value of the lower pK was dependent on the substrate analogue, which suggested that it was not exclusively the result of ionization of the phosphoryl moiety. During the decarboxylation of OMP, the fluorescence of E(2) was quenched over 20%. The enzymatic species with reduced fluorescence was a catalytically competent intermediate that had kinetic properties consistent with it being the initial enzyme-substrate complex. The stoichiometry for binding of OMP to E(2) was one OMP per enzyme monomer. The value of the first-order rate constant for conversion of the enzyme-substrate complex to free enzyme (36 s(-1)) calculated from a single turnover experiment ([E] > [S]) was slightly greater than the value of k(cat), 20 s(-1) (corrected for stoichiometry), calculated from steady-state data. In the single turnover experiments, the enzyme was E(2)*S, whereas in the steady-state turnover the experiment enzyme was E(2)*S(2). The similarity of these values suggested that the subunits were catalytically independent such that E(2)*S(2) could be treated as E*S and that conversion of the enzyme-substrate complex to E was k(cat). Kinetic data for the approach to the steady-state with OMP and E(2) yield a bimolecular association rate complex of 62 microM(-1) s(-1)and a dissociation rate constant for E*S of 60 s(-1). The commitment to catalysis was 0.25. By monitoring the effect of carbonic anhydrase on [H(+)] changes during a single turnover experiment, the initial product of the decarboxylation reaction was shown to be CO(2) not HCO(3-). UMP was released from the enzyme concomitantly with CO(2) during the conversion of E*S to E. Furthermore, the enzyme removed an enzyme equivalent of H(+) from solvent during this step of the reaction. The bimolecular rate constants for association of 6-AzaUMP and 8-AzaXMP, substrate analogues with markedly different nucleobases, had association rate constants of 112 and 130 microM(-1) s(-1), respectively. These results suggested that the nucleobase did not contribute significantly to the success of formation of the initial enzyme-substrate complex.  相似文献   

20.
We assessed whether reactive oxygen-nitrogen intermediates generated by alveolar macrophages (AMs) oxidized and nitrated human surfactant protein (SP) A. SP-A was exposed to lipopolysaccharide (100 ng/ml)-activated AMs in 15 mM HEPES (pH 7.4) for 30 min in the presence and absence of 1.2 mM CO(2). In the presence of CO(2), lipopolysaccharide-stimulated AMs had significantly higher nitric oxide synthase (NOS) activity (as quantified by the conversion of L-[U-(14)C]arginine to L-[U-(14)C]citrulline) and secreted threefold higher levels of nitrate plus nitrite in the medium [28 +/- 3 vs. 6 +/- 1 (SE) nmol. 6.5 h(-1). 10(6) AMs(-1)]. Western blotting studies of immunoprecipitated SP-A indicated that CO(2) enhanced SP-A nitration by AMs and decreased carbonyl formation. CO(2) (0-1.2 mM) also augmented peroxynitrite (0.5 mM)-induced SP-A nitration in a dose-dependent fashion. Peroxynitrite decreased the ability of SP-A to aggregate lipids, and this inhibition was augmented by 1.2 mM CO(2). Mass spectrometry analysis of chymotryptic fragments of peroxynitrite-exposed SP-A showed nitration of two tyrosines (Tyr(164) and Tyr(166)) in the absence of CO(2) and three tyrosines (Tyr(164), Tyr(166), and Tyr(161)) in the presence of 1.2 mM CO(2). These findings indicate that physiological levels of peroxynitrite, produced by activated AMs, nitrate SP-A and that CO(2) increased nitration, at least partially, by enhancing enzymatic nitric oxide production.  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