Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 20
Filtrar
Mais filtros











Base de dados
Intervalo de ano de publicação
1.
Methods Enzymol ; 463: 57-71, 2009.
Artigo em Inglês | MEDLINE | ID: mdl-19892167

RESUMO

To study and understand the nature of living cells, scientists have continually employed traditional biochemical techniques aimed to fractionate and characterize a designated network of macromolecular components required to carry out a particular cellular function. At the most rudimentary level, cellular functions ultimately entail rapid chemical transformations that otherwise would not occur in the physiological environment of the cell. The term enzyme is used to singularly designate a macromolecular gene product that specifically and greatly enhances the rate of a chemical transformation. Purification and characterization of individual and collective groups of enzymes has been and will remain essential toward advancement of the molecular biological sciences; and developing and utilizing enzyme reaction assays is central to this mission. First, basic kinetic principles are described for understanding chemical reaction rates and the catalytic effects of enzymes on such rates. Then, a number of methods are described for measuring enzyme-catalyzed reaction rates, which mainly differ with regard to techniques used to detect and quantify concentration changes of given reactants or products. Finally, short commentary is given toward formulation of reaction mixtures used to measure enzyme activity. Whereas a comprehensive treatment of enzymatic reaction assays is not within the scope of this chapter, the very core principles that are presented should enable new researchers to better understand the logic and utility of any given enzymatic assay that becomes of interest.


Assuntos
Ensaios Enzimáticos/métodos , Enzimas/metabolismo , Animais , Catálise , Enzimas/isolamento & purificação , Humanos , Cinética , Modelos Químicos
2.
Biochemistry ; 40(23): 6805-18, 2001 Jun 12.
Artigo em Inglês | MEDLINE | ID: mdl-11389594

RESUMO

Methylglyoxal synthase (MGS) and triosephosphate isomerase (TIM) share neither sequence nor structural similarities, yet the reactions catalyzed by both enzymes are similar, in that both initially convert dihydroxyacetone phosphate to a cis-enediolic intermediate. This enediolic intermediate is formed from the abstraction of the pro-S C3 proton of DHAP by Asp-71 of MGS or the pro-R C3 proton of DHAP by Glu-165 of TIM. MGS then catalyzes the elimination of phosphate from this enediolic intermediate to form the enol of methylglyoxal, while TIM catalyzes proton donation to C2 to form D-glyceraldehyde phosphate. A competitive inhibitor of TIM, phosphoglycolohydroxamic acid (PGH) is found to be a tight binding competitive inhibitor of MGS with a K(i) of 39 nM. PGH's high affinity for MGS may be due in part to a short, strong hydrogen bond (SSHB) from the NOH of PGH to the carboxylate of Asp-71. Evidence for this SSHB is found in X-ray, 1H NMR, and fractionation factor data. The X-ray structure of the MGS homohexamer complexed with PGH at 2.0 A resolution shows this distance to be 2.30-2.37 +/- 0.24 A. 1H NMR shows a PGH-dependent 18.1 ppm signal that is consistent with a hydrogen bond length of 2.49 +/- 0.02 A. The D/H fractionation factor (phi = 0.43 +/- 0.02) is consistent with a hydrogen bond length of 2.53 +/- 0.01 A. Further, 15N NMR suggests a significant partial positive charge on the nitrogen atom of bound PGH, which could strengthen hydrogen bond donation to Asp-71. Both His-98 and His-19 are uncharged in the MGS-PGH complex on the basis of the chemical shifts of their Cdelta and C(epsilon) protons. The crystal structure reveals that Asp-71, on the re face of PGH, and His-19, on the si face of PGH, both approach the NO group of the analogue, while His-98, in the plane of PGH, approaches the carbonyl oxygen of the analogue. The phosphate group of PGH accepts nine hydrogen bonds from seven residues and is tilted out of the imidate plane of PGH toward the re face. Asp-71 and phosphate are thus positioned to function as the base and leaving group, respectively, in a concerted suprafacial 1,4-elimination of phosphate from the enediolic intermediate in the second step of the MGS reaction. Combined, these data suggest that Asp-71 is the one base that initially abstracts the C3 pro-S proton from DHAP and subsequently the 3-OH proton from the enediolic intermediate. This mechanism is compared to an alternative TIM-like mechanism for MGS, and the relative merits of both mechanisms are discussed.


Assuntos
Carbono-Oxigênio Liases/química , Inibidores Enzimáticos/química , Ácidos Hidroxâmicos/química , Substituição de Aminoácidos/genética , Asparagina/genética , Ácido Aspártico/genética , Sítios de Ligação/genética , Ligação Competitiva/genética , Carbono-Oxigênio Liases/antagonistas & inibidores , Carbono-Oxigênio Liases/genética , Fracionamento Químico , Cristalografia por Raios X , Escherichia coli/enzimologia , Escherichia coli/genética , Glicolatos/química , Cinética , Substâncias Macromoleculares , Ressonância Magnética Nuclear Biomolecular/métodos , Prótons , Proteínas Recombinantes/química , Triose-Fosfato Isomerase/química
3.
Biochemistry ; 40(7): 1984-95, 2001 Feb 20.
Artigo em Inglês | MEDLINE | ID: mdl-11329265

RESUMO

The amino-terminal proline of 4-oxalocrotonate tautomerase (4-OT) functions as the general base catalyst in the enzyme-catalyzed isomerization of beta,gamma-unsaturated enones to their alpha,beta-isomers because of its unusually low pK(a) of 6.4 +/- 0.2, which is 3 units lower than that of the model compound, proline amide. Recent studies show that this abnormally low pK(a) is not due to the electrostatic effects of nearby cationic residues (Arg-11, Arg-39, and Arg-61) [Czerwinski, R. M., Harris, T. K., Johnson, Jr., W. H., Legler, P. M., Stivers, J. T., Mildvan, A. S., and Whitman, C. P. (1999) Biochemistry 38, 12358-12366]. Hence, it may result solely from a low local dielectric constant of 14.7 +/- 0.8 at the otherwise hydrophobic active site. Support for this mechanism comes from the study of mutants of the active site Phe-50, which is 5.8 A from Pro-1 and is one of 12 apolar residues within 9 A of Pro-1. Replacing Phe-50 with Tyr does not significantly alter k(cat) or K(m) and results in a pK(a) of 6.0 +/- 0.1 for Pro-1 as determined by (15)N NMR spectroscopy, comparable to that observed for wild type. (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY HSQC spectra of the F50Y mutant demonstrate its conformation to be very similar to that of the wild-type enzyme. In the F50Y mutant, the pK(a) of Tyr-50 is increased by two units from that of a model compound N-acetyl-tyrosine amide to 12.2 +/- 0.3, as determined by UV and (1)H NMR titrations, yielding a local dielectric constant of 13.4 +/- 1.7, in agreement with the value of 13.7 +/- 0.3 determined from the decreased pK(a) of Pro-1 in this mutant. In the F50A mutant, the pK(a) of Pro-1 is 7.3 +/- 0.1 by (15)N NMR titration, comparable to the pK(a) of 7.6 +/- 0.2 found in the pH vs k(cat)/K(m) rate profile, and is one unit greater than that of the wild-type enzyme, indicating an increase in the local dielectric constant to a value of 21.2 +/- 2.6. A loss of structure of the beta-hairpin from residues 50 to 57, which covers the active site, and is the site of the mutation, is indicated by the disappearance in the F50A mutant of four interstrand NOEs and one turn NOE found in wild-type 4-OT. (1)H-(15)N HSQC spectra of the F50A mutant reveal widespread and large changes in the backbone (15)N and NH chemical shifts including those of Gly residues 48, 51, 53, and 54 causing their loss of dispersion at 23 degrees C and their disappearance at 43 degrees C due to rapid exchange with solvent. These observations confirm that the active site of the F50A mutant is more accessible to the external aqueous environment, causing an increase in the local dielectric constant and in the pK(a) of Pro-1. In addition, the F50A mutation decreased k(cat) 167-fold and increased K(m) 11-fold from those of the wild-type enzyme, suggesting an important role for the hydrophobic environment in catalysis, beyond that of decreasing the pK(a) of Pro-1. The F50I and F50V mutations destabilize the protein and decrease k(cat) by factors of 58 and 1.6, and increase K(m) by 3.3- and 3.8-fold, respectively.


Assuntos
Isomerases/química , Isomerases/genética , Mutagênese Sítio-Dirigida , Fenilalanina/genética , Alanina/genética , Sequência de Aminoácidos , Catálise , Dicroísmo Circular , Escherichia coli/enzimologia , Escherichia coli/genética , Concentração de Íons de Hidrogênio , Isomerases/metabolismo , Cinética , Modelos Químicos , Modelos Moleculares , Dados de Sequência Molecular , Isótopos de Nitrogênio , Ressonância Magnética Nuclear Biomolecular , Fenilalanina/química , Fenilalanina/metabolismo , Reação em Cadeia da Polimerase , Prolina/metabolismo , Proteínas Recombinantes/biossíntese , Proteínas Recombinantes/química , Proteínas Recombinantes/metabolismo , Relação Estrutura-Atividade , Titulometria , Tirosina/genética , Tirosina/metabolismo
4.
Biochemistry ; 39(51): 16200-5, 2000 Dec 26.
Artigo em Inglês | MEDLINE | ID: mdl-11123949

RESUMO

Cholinesterases (ChE), use a Glu-His-Ser catalytic triad to enhance the nucleophilicity of the catalytic serine. It has been shown that serine proteases, which employ an Asp-His-Ser catalytic triad for optimal catalytic efficiency, decrease the hydrogen bonding distance between the Asp-His pair to form a short, strong hydrogen bond (SSHB) upon binding mechanism-based inhibitors, which form tetrahedral Ser-adducts, analogous to the tetrahedral intermediates in catalysis, or at low pH when the histidine is protonated [Cassidy, C. S., Lin, J., Frey, P. A. (1997) Biochemistry 36, 4576-4584]. Two types of mechanism-based inhibitors were bound to pure equine butyrylcholinesterase (BChE), a 364 kDa homotetramer, and the complexes were studied by (1)H NMR at 600 MHz and 25-37 degrees C. The downfield region of the (1)H NMR spectrum of free BChE at pH 7.5 showed a broad, weak, deshielded resonance with a chemical shift, delta = 16.1 ppm, ascribed to a small amount of the histidine-protonated form. Upon addition of a 3-fold excess of diethyl 4-nitrophenyl phosphate (paraoxon) and subsequent dealkylation, the broad 16.1 ppm resonance increased in intensity 4.7-fold, and yielded a D/H fractionation factor phi = 0.72+/-0.10 consistent with a SSHB between Glu and His of the catalytic triad. From an empirical correlation of delta with hydrogen-bond length in small crystalline compounds, the length of this SSBH is 2.64+/-0.04 A, in agreement with the length of 2.62+/-0.02 A independently obtained from phi. The addition of a 3-fold excess of m-(N,N, N-trimethylammonio)trifluoroacetophenone to BChE yielded no signal at 16.1 ppm, and a 640 Hz broad, highly deshielded proton resonance with a chemical shift delta = 18.1 ppm and a D/H fractionation factor phi = 0.63+/-0.10, also consistent with a SSHB. The length of this SSHB is calculated to be 2.62+/-0.04 A from delta and 2.59+/-0.03 A from phi. These NMR-derived distances agree with those found in the X-ray structures of the homologous acetylcholinesterase complexed with the same mechanism-based inhibitors, 2.60+/-0.22 and 2.66+/-0.28 A. However, the order of magnitude greater precision of the NMR-derived distances establish the presence of SSHBs. We suggest that ChEs achieve their remarkable catalytic power in ester hydrolysis, in part, due to the formation of a SSHB between Glu and His of the catalytic triad.


Assuntos
Butirilcolinesterase/química , Acetofenonas/metabolismo , Acetofenonas/farmacologia , Animais , Sítios de Ligação/efeitos dos fármacos , Butirilcolinesterase/metabolismo , Inibidores da Colinesterase/metabolismo , Inibidores da Colinesterase/farmacologia , Ativação Enzimática/efeitos dos fármacos , Cavalos , Ligação de Hidrogênio/efeitos dos fármacos , Cinética , Ressonância Magnética Nuclear Biomolecular/métodos , Paraoxon/metabolismo , Paraoxon/farmacologia
5.
Biochemistry ; 39(7): 1655-74, 2000 Feb 22.
Artigo em Inglês | MEDLINE | ID: mdl-10677214

RESUMO

The MutT enzyme catalyzes the hydrolysis of nucleoside triphosphates (NTP) to NMP and PP(i) by nucleophilic substitution at the rarely attacked beta-phosphorus. The solution structure of the quaternary E-M(2+)-AMPCPP-M(2+) complex indicated that conserved residues Glu-53, -56, -57, and -98 are at the active site near the bound divalent cation possibly serving as metal ligands, Lys-39 is positioned to promote departure of the NMP leaving group, and Glu-44 precedes helix I (residues 47-59) possibly stabilizing this helix which contributes four catalytic residues to the active site [Lin, J. , Abeygunawardana, C., Frick, D. N., Bessman, M. J., and Mildvan, A. S. (1997) Biochemistry 36, 1199-1211]. To test these proposed roles, the effects of mutations of each of these residues on the kinetic parameters and on the Mn(2+), Mg(2+), and substrate binding properties were examined. The largest decreases in k(cat) for the Mg(2+)-activated enzyme of 10(4.7)- and 10(2.6)-fold were observed for the E53Q and E53D mutants, respectively, while 97-, 48-, 25-, and 14-fold decreases were observed for the E44D, E56D, E56Q, and E44Q mutations, respectively. Smaller effects on k(cat) were observed for mutations of Glu-98 and Lys-39. For wild type MutT and its E53D and E44D mutants, plots of log(k(cat)) versus pH exhibited a limiting slope of 1 on the ascending limb and then a hump, i.e., a sharply defined maximum near pH 8 followed by a plateau, yielding apparent pK(a) values of 7.6 +/- 0.3 and 8.4 +/- 0.4 for an essential base and a nonessential acid catalyst, respectively, in the active quaternary MutT-Mg(2+)-dGTP-Mg(2+) complex. The pK(a) of 7.6 is assigned to Glu-53, functioning as a base catalyst in the active quaternary complex, on the basis of the disappearance of the ascending limb of the pH-rate profile of the E53Q mutant, and its restoration in the E53D mutant with a 10(1.9)-fold increase in (k(cat))(max). The pK(a) of 8.4 is assigned to Lys-39 on the basis of the disappearance of the descending limb of the pH-rate profile of the K39Q mutant, and the observation that removal of the positive charge of Lys-39, by either deprotonation or mutation, results in the same 8.7-fold decrease in k(cat). Values of k(cat) of both wild type MutT and the E53Q mutant were independent of solvent viscosity, indicating that a chemical step is likely to be rate-limiting with both. A liganding role for Glu-53 and Glu-56, but not Glu-98, in the binary E-M(2+) complex is indicated by the observation that the E53Q, E53D, E56Q, and E56D mutants bound Mn(2+) at the active site 36-, 27-, 4.7-, and 1.9-fold weaker, and exhibited 2.10-, 1.50-, 1.12-, and 1.24-fold lower enhanced paramagnetic effects of Mn(2+), respectively, than the wild type enzyme as detected by 1/T(1) values of water protons, consistent with the loss of a metal ligand. However, the K(m) values of Mg(2+) and Mn(2+) indicate that Glu-56, and to a lesser degree Glu-98, contribute to metal binding in the active quaternary complex. Mutations of the more distant but conserved residue Glu-44 had little effect on metal binding or enhancement factors in the binary E-M(2+) complexes. Two-dimensional (1)H-(15)N HSQC and three-dimensional (1)H-(15)N NOESY-HSQC spectra of the kinetically damaged E53Q and E56Q mutants showed largely intact proteins with structural changes near the mutated residues. Structural changes in the kinetically more damaged E44D mutant detected in (1)H-(15)N HSQC spectra were largely limited to the loop I-helix I motif, suggesting that Glu-44 stabilizes the active site region. (1)H-(15)N HSQC titrations of the E53Q, E56Q, and E44D mutants with dGTP showed changes in chemical shifts of residues lining the active site cleft, and revealed tighter nucleotide binding by these mutants, indicating an intact substrate binding site. (ABSTRACT TRUNCATED)


Assuntos
Proteínas de Bactérias/genética , Proteínas de Bactérias/metabolismo , Proteínas de Escherichia coli , Ácido Glutâmico/genética , Ácido Glutâmico/metabolismo , Lisina/genética , Mutagênese Sítio-Dirigida , Monoéster Fosfórico Hidrolases/genética , Monoéster Fosfórico Hidrolases/metabolismo , Trifosfato de Adenosina/análogos & derivados , Trifosfato de Adenosina/química , Trifosfato de Adenosina/genética , Ácido Aspártico/genética , Proteínas de Bactérias/química , Ligação Competitiva/genética , Catálise , Cloretos/química , Sequência Conservada , Análise Mutacional de DNA , Ácido Glutâmico/química , Glutamina/genética , Cinética , Lisina/química , Magnésio/química , Magnésio/metabolismo , Cloreto de Magnésio/química , Manganês/química , Manganês/metabolismo , Compostos de Manganês/química , Isótopos de Nitrogênio , Ressonância Magnética Nuclear Biomolecular , Monoéster Fosfórico Hidrolases/química , Ligação Proteica/genética , Conformação Proteica , Estrutura Terciária de Proteína/genética , Prótons , Pirofosfatases , Proteínas Recombinantes/química , Proteínas Recombinantes/metabolismo , Solventes
7.
Biochemistry ; 38(38): 12343-57, 1999 Sep 21.
Artigo em Inglês | MEDLINE | ID: mdl-10493802

RESUMO

Three arginine residues (Arg-11, Arg-39, Arg-61) are found at the active site of 4-oxalocrotonate tautomerase in the X-ray structure of the affinity-labeled enzyme [Taylor, A. B., Czerwinski, R. M., Johnson, R. M., Jr., Whitman, C. P., and Hackert, M. L. (1998) Biochemistry 37, 14692-14700]. The catalytic roles of these arginines were examined by mutagenesis, kinetic, and heteronuclear NMR studies. With a 1,6-dicarboxylate substrate (2-hydroxymuconate), the R61A mutation showed no kinetic effects, while the R11A mutation decreased k(cat) 88-fold and increased K(m) 8.6-fold, suggesting both binding and catalytic roles for Arg-11. With a 1-monocarboxylate substrate (2-hydroxy-2,4-pentadienoate), no kinetic effects of the R11A mutation were found, indicating that Arg-11 interacts with the 6-carboxylate of the substrate. The stereoselectivity of the R11A-catalyzed protonation at C-5 of the dicarboxylate substrate decreased, while the stereoselectivity of protonation at C-3 of the monocarboxylate substrate increased in comparison with wild-type 4-OT, indicating the importance of Arg-11 in properly orienting the dicarboxylate substrate by interacting with the charged 6-carboxylate group. With 2-hydroxymuconate, the R39A and R39Q mutations decreased k(cat) by 125- and 389-fold and increased K(m) by 1.5- and 2.6-fold, respectively, suggesting a largely catalytic role for Arg-39. The activity of the R11A/R39A double mutant was at least 10(4)-fold lower than that of the wild-type enzyme, indicating approximate additivity of the effects of the two arginine mutants on k(cat). For both R11A and R39Q, 2D (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY-HSQC spectra showed chemical shift changes mainly near the mutated residues, indicating otherwise intact protein structures. The changes in the R39Q mutant were mainly in the beta-hairpin from residues 50 to 57 which covers the active site. HSQC titration of R11A with the substrate analogue cis, cis-muconate yielded a K(d) of 22 mM, 37-fold greater than the K(d) found with wild-type 4-OT (0.6 mM). With the R39Q mutant, cis, cis-muconate showed negative cooperativity in active site binding with two K(d) values, 3.5 and 29 mM. This observation together with the low K(m) of 2-hydroxymuconate (0.47 mM) suggests that only the tight binding sites function catalytically in the R39Q mutant. The (15)Nepsilon resonances of all six Arg residues of 4-OT were assigned, and the assignments of Arg-11, -39, and -61 were confirmed by mutagenesis. The binding of cis,cis-muconate to wild-type 4-OT upshifts Arg-11 Nepsilon (by 0.05 ppm) and downshifts Arg-39 Nepsilon (by 1.19 ppm), indicating differing electronic delocalizations in the guanidinium groups. A mechanism is proposed in which Arg-11 interacts with the 6-carboxylate of the substrate to facilitate both substrate binding and catalysis and Arg-39 interacts with the 1-carboxylate and the 2-keto group of the substrate to promote carbonyl polarization and catalysis, while Pro-1 transfers protons from C-3 to C-5. This mechanism, together with the effects of mutations of catalytic residues on k(cat), provides a quantitative explanation of the 10(7)-fold catalytic power of 4-OT. Despite its presence in the active site in the crystal structure of the affinity-labeled enzyme, Arg-61 does not play a significant role in either substrate binding or catalysis.


Assuntos
Arginina/genética , Isomerases/química , Isomerases/genética , Mutagênese Sítio-Dirigida , Alanina/genética , Sítios de Ligação/genética , Catálise , Glutamina/genética , Isomerases/biossíntese , Isomerases/metabolismo , Cinética , Modelos Moleculares , Isótopos de Nitrogênio , Ressonância Magnética Nuclear Biomolecular , Pseudomonas putida/enzimologia , Pseudomonas putida/genética , Proteínas Recombinantes/biossíntese , Proteínas Recombinantes/síntese química , Proteínas Recombinantes/química , Ácido Sórbico/análogos & derivados , Ácido Sórbico/metabolismo , Estereoisomerismo , Titulometria
8.
Biochemistry ; 38(38): 12358-66, 1999 Sep 21.
Artigo em Inglês | MEDLINE | ID: mdl-10493803

RESUMO

The unusually low pK(a) value of the general base catalyst Pro-1 (pK(a) = 6.4) in 4-oxalocrotonate tautomerase (4-OT) has been ascribed to both a low dielectric constant at the active site and the proximity of the cationic residues Arg-11 and Arg-39 [Stivers, J. T., Abeygunawardana, C., Mildvan, A. S., Hajipour, G., and Whitman, C. P. (1996) Biochemistry 35, 814-823]. In addition, the pH-rate profiles in that study showed an unidentified protonated group essential for catalysis with a pK(a) of 9.0. To address these issues, the pK(a) values of the active site Pro-1 and lower limit pK(a) values of arginine residues were determined by direct (15)N NMR pH titrations. The pK(a) values of Pro-1 and of the essential acid group were determined independently from pH-rate profiles of the kinetic parameters of 4-OT in arginine mutants of 4-OT and compared with those of wild type. The chemical shifts of all of the Arg Nepsilon resonances in wild-type 4-OT and in the R11A and R39Q mutants were found to be independent of pH over the range 4.9-9.7, indicating that no arginine is responsible for the kinetically determined pK(a) of 9.0 for an acidic group in free 4-OT. With the R11A mutant, where k(cat)/K(m) was reduced by a factor of 10(2.9), the pK(a) of Pro-1 was not significantly altered from that of the wild-type enzyme (pK(a) = 6.4 +/- 0.2) as revealed by both direct (15)N NMR titration (pK(a) = 6.3 +/- 0.1) and the pH dependence of k(cat)/K(m) (pK(a) = 6.4 +/- 0.2). The pH-rate profiles of both k(cat)/K(m) and k(cat) for the reaction of the R11A mutant with the dicarboxylate substrate, 2-hydroxymuconate, showed humps, i.e., sharply defined maxima followed by nonzero plateaus. The humps disappeared in the reaction with the monocarboxylate substrate, 2-hydroxy-2,4-pentadienoate, indicating that, unlike the wild-type enzyme which reacts only with the dianionic form of the dicarboxylic substrate, the R11A mutant reacts with both the 6-COOH and 6-COO(-) forms, with the 6-COOH form being 12-fold more active. This reversal in the preferred ionization state of the 6-carboxyl group of the substrate that occurs upon mutation of Arg-11 to Ala provides strong evidence that Arg-11 interacts with the 6-carboxylate of the substrate. In the R39Q mutant, where k(cat)/K(m) was reduced by a factor of 10(3), the kinetically determined pK(a) value for Pro-1 was 4.6 +/- 0.2, while the ionization of Pro-1 showed negative cooperativity with an apparent pK(a) of 7.1 +/- 0.1 determined by 1D (15)N NMR. From the Hill coefficient of 0.54, it can be shown that the apparent pK(a) value of 7.1 could result most simply from the averaging of two limiting pK(a) values of 4.6 and 8.2. Mutation of Arg-39, by altering the structure of the beta-hairpin which covers the active site, could result in an increase in the solvent exposure of Pro-1, raising its upper limit pK(a) value to 8.2. In the R39A mutant, the kinetically determined pK(a) of Pro-1 was also low, 5.0 +/- 0.2, indicating that in both the R39Q and R39A mutants, only the sites with low pK(a) values were kinetically operative. With the fully active R61A mutant, the kinetically determined pK(a) of Pro-1 (pK(a) = 6.5 +/- 0.2) agreed with that of wild-type 4-OT. It is concluded that the unusually low pK(a) of Pro-1 shows little contribution from electrostatic effects of the nearby cationic Arg-11, Arg-39, and Arg-61 residues but results primarily from a site of low local dielectric constant.


Assuntos
Arginina/genética , Arginina/metabolismo , Isomerases/genética , Isomerases/metabolismo , Mutagênese Sítio-Dirigida , Alanina/genética , Sítios de Ligação/genética , Catálise , Glutamina/metabolismo , Concentração de Íons de Hidrogênio , Cinética , Isótopos de Nitrogênio , Ressonância Magnética Nuclear Biomolecular , Prolina/metabolismo , Pseudomonas putida/enzimologia , Pseudomonas putida/genética , Especificidade por Substrato , Titulometria
9.
Proteins ; 35(3): 275-82, 1999 May 15.
Artigo em Inglês | MEDLINE | ID: mdl-10328262

RESUMO

We have compared hydrogen bond lengths on enzymes derived with high precision (< or = +/- 0.05 A) from both the proton chemical shifts (delta) and the fractionation factors (phi) of the proton involved with those obtained from protein X-ray crystallography. Hydrogen bond distances derived from proton chemical shifts were obtained from a correlation of 59 O--H....O hydrogen bond lengths, measured by small molecule high-resolution X-ray crystallography, with chemical shifts determined by solid-state nuclear magnetic resonance (NMR) in the same crystals (McDermott A, Ridenour CF, Encyclopedia of NMR, Sussex, U.K.: Wiley, 1996:3820-3825). Hydrogen bond distances were independently obtained from fractionation factors that yield distances between the two proton wells in quartic double minimum potential functions (Kreevoy MM, Liang TM, J Am Chem Soc, 1980;102:3315-3322). The high-precision hydrogen bond distances derived from their corresponding NMR-measured proton chemical shifts and fractionation factors agree well with each other and with those reported in protein X-ray structures within the larger errors (+/-0.2-0.8 A) in distances obtained by protein X-ray crystallography. The increased precision in measurements of hydrogen bond lengths by NMR has provided insight into the contributions of short, strong hydrogen bonds to catalysis for several enzymatic reactions.


Assuntos
Proteínas/química , Ligação de Hidrogênio , Espectroscopia de Ressonância Magnética , Prótons , Serina Endopeptidases/química , Esteroide Isomerases/química , Triose-Fosfato Isomerase/química
10.
Biochemistry ; 37(47): 16828-38, 1998 Nov 24.
Artigo em Inglês | MEDLINE | ID: mdl-9843453

RESUMO

Triosephosphate isomerase (TIM) catalyzes the reversible interconversion of dihydroxyacetone phosphate (DHAP) and glyceraldehyde 3-phosphate (GAP), with Glu-165 removing the pro-R proton from C1 of DHAP and neutral His-95 polarizing the carbonyl group of the substrate. During the TIM reaction, approximately 2% of the pro-R tritium from C1 of DHAP is conserved and appears at C2 of GAP [Nickbarg, E. B., and Knowles, J. R. (1988) Biochemistry 27, 5939]. In the "classical" mechanism, 98% of the pro-R tritium exchanges with solvent from Glu-165 at the intermediate state and the remaining 2% is transferred by Glu-165 to C2 of the same substrate molecule. This intramolecular transfer of tritium is therefore predicted to be independent of DHAP concentration. On the basis of NMR detection of a strong hydrogen bond between Glu-165 and the 1-OH of an analogue of the enediol intermediate [Harris, T. K., Abeygunawardana, C., and Mildvan, A. S. (1997) Biochemistry 36, 14661], we have suggested a "criss-cross" mechanism for TIM in which Glu-165 transfers a proton from C1 of DHAP to O2 of the enediol, and subsequently from O1 of the enediol to C2 of the product GAP. Since the pro-R proton is transferred to O2 instead of C2 in the criss-cross mechanism, no intramolecular transfer of label from substrate to product would be expected to occur. However, intermolecular transfer of label could occur if the label exchanges from O2 into a group on the protein and is transferred to GAP in subsequent turnovers. The extent of intermolecular tritium transfer in the criss-cross mechanism would be predicted to be dependent on DHAP concentration. The extent of tritium transfer was studied as a function of initial DHAP concentration using DHAP highly tritiated at the pro-R position. At 50% conversion to GAP, triphasic tritium transfer behavior was found. For phase 1, between 0.03 and 0.3 mM DHAP, a constant extent of tritium transfer of 1.19 +/- 0.03% occurred. For phase 2, between 0.3 and 1.0 mM DHAP, the extent of transfer progressively increased as a function of DHAP concentration to 2.17 +/- 0.15%. For phase 3, between 1.0 and 7.0 mM DHAP, the extent of transfer slightly decreased to 1.68 +/- 0.17%. In a direct test for intermolecular isotope transfer, doubly labeled [1(R)-D, 13C3]DHAP and 13C-depleted [1(R)-H,12C3]DHAP were synthesized, mixed in equal amounts, and incubated at 1 mM total DHAP with TIM, GAP dehydrogenase, NAD+, and arsenate until 50% conversion to 3-phosphoglycerate occurred. Electrospray ionization mass spectral analysis of the stable 3-phosphoglycerate product detected an extent of 1.4 +/- 0.4% of intramolecular D transfer from [13C3]DHAP to the 13C3 product, but no intermolecular transfer (

Assuntos
Prótons , Triose-Fosfato Isomerase/química , Animais , Deutério/química , Deutério/metabolismo , Fosfato de Di-Hidroxiacetona/química , Fosfato de Di-Hidroxiacetona/metabolismo , Transporte de Elétrons , Gliceraldeído 3-Fosfato/química , Gliceraldeído 3-Fosfato/metabolismo , Músculo Esquelético/enzimologia , Coelhos , Especificidade por Substrato , Temperatura , Triose-Fosfato Isomerase/metabolismo , Trítio/química , Trítio/metabolismo
11.
Biochemistry ; 36(48): 14661-75, 1997 Dec 02.
Artigo em Inglês | MEDLINE | ID: mdl-9398185

RESUMO

Triosephosphate isomerase (TIM) catalyzes the reversible interconversion of dihydroxyacetone phosphate (DHAP) and glyceraldehyde-3-phosphate (GAP), with Glu-165 removing the pro-R proton from C1 of DHAP and neutral His-95 polarizing the carbonyl group of the substrate. TIM and its complexes with the reactive intermediate analogs, phosphoglycolic acid (PGA) and phosphoglycolohydroxamic acid (PGH), were studied by 1H NMR at 600 MHz and at low temperature (-4.8 degrees C). His-95 shows an N epsilon H resonance at 13.1 ppm which shifts to 13.3 ppm in the TIM-PGA complex and to 13.5 ppm in the TIM-PGH complex. In the TIM-PGH complex, His-95 N epsilon H shows a slow, pH-independent exchange rate with water (kex = 80 s-1 at 30 degrees C, Eact = 19 kcal/mol), which is 44-fold slower than that of an exposed histidine suggesting partial shielding from bulk solvent, and a fractionation factor phi = 0.71 +/- 0.02 consistent with its donation of a normal hydrogen bond. The formation of the TIM-PGH complex results in the appearance of several deshielded proton resonances, including one at 14.9 ppm and one at 10.9 ppm which overlaps with another resonance. The resonance at 14.9 ppm is absent and the resonance at 10.9 ppm is much weaker in the TIM complex of PGA, which lacks the hydroxamic acid (-NHOH) moiety. 15N-labeled PGH was synthesized and the NH proton of free [15N]PGH shows a single 1H-15N HMQC cross peak with delta (1H) = 10.3 ppm and delta (15N) = 168 ppm which shifts to delta (1H) = 10.9 ppm and delta (15N) = 174 ppm in the TIM complex of [15N]PGH. The 15N-1H coupling in the complex indicates covalent N-H bonding, and the deshielded delta (15N) indicates a significant contribution of the imidate resonance form of PGH. The 14.9 ppm resonance is assigned to the NOH proton of bound PGH. This resonance shows a pH-independent exchange rate with water (kex = 3900 s-1 at 30 degrees C, Eact = 8.9 kcal/mol) which may reflect the dissociation of the TIM-PGH complex, and meets the criteria for a low-barrier hydrogen bond on the basis of the significant downfield shift of 6.2 ppm from the NOH proton of the model compound acetohydroxamic acid, and a very low fractionation factor phi = 0.38 +/- 0.06. In the X-ray structure of the TIM-PGH complex [Davenport, R. C., Bash, P. A., Seaton, B. A., Karplus, M., Petsko, G. A., and Ringe, D. (1991) Biochemistry 30, 5821], the NOH proton of bound PGH is hydrogen bonded to Glu-165. A low-barrier hydrogen bond from PGH NOH to Glu-165 suggests a dual role for Glu-165 in catalysis of proton transfer not only between the C1 and C2 carbons but also between the O1 and O2 oxygens in the interconversion of DHAP and GAP in wild type TIM. Such a mechanism, together with the measured exchange rate of the His-95 N epsilon H proton with solvent protons can accommodate the classical measurements of tritium incorporation from DHAP into GAP.


Assuntos
Triose-Fosfato Isomerase/metabolismo , Simulação por Computador , Histidina/química , Histidina/metabolismo , Ligação de Hidrogênio , Ácidos Hidroxâmicos/química , Ácidos Hidroxâmicos/metabolismo , Modelos Químicos , Isótopos de Nitrogênio , Ressonância Magnética Nuclear Biomolecular , Titulometria , Triose-Fosfato Isomerase/química , Leveduras/enzimologia
12.
Biochemistry ; 36(17): 5212-22, 1997 Apr 29.
Artigo em Inglês | MEDLINE | ID: mdl-9136883

RESUMO

The Vaccinia type I topoisomerase catalyzes site-specific DNA strand cleavage and religation by forming a transient phosphotyrosyl linkage between the DNA and Tyr-274, resulting in the release of DNA supercoils. For type I topoisomerases, two mechanisms have been proposed for supercoil release: (I) a coupled mechanism termed strand passage, in which a single supercoil is removed per cleavage/religation cycle, resulting in multiple topoisomer intermediates and late product formation, or (2) an uncoupled mechanism termed free rotation, where multiple supercoils are removed per cleavage/religation cycle, resulting in few intermediates and early product formation. To determine the mechanism, single-turnover experiments were done with supercoiled plasmid DNA under conditions in which the topoisomerase cleaves predominantly at a single site per DNA molecule. The concentrations of supercoiled substrate, intermediate topoisomers, and relaxed product vs time were measured by fluorescence imaging, and the rate constants for their interconversion were determined by kinetic simulation. Few intermediates and early product formation were observed. From these data, the rate constants for cleavage (0.3 s(-1)), religation (4 s(-1)), and the cleavage equilibrium constant on the enzyme (0.075) at 22 degrees C are in reasonable agreement with those obtained with small oligonucleotide substrates, while the rotation rate of the cleaved DNA strand is fast (approximately 20 rotations/s). Thus, the average number of supercoils removed for each cleavage event greatly exceeds unity (delta n = 5) and depends on kinetic competition between religation and supercoil release, establishing a free rotation mechanism. This free rotation mechanism for a type I topoisomerase differs from the strand passage mechanism proposed for the type II enzymes.


Assuntos
DNA Topoisomerases Tipo I/metabolismo , DNA Super-Helicoidal/metabolismo , Conformação de Ácido Nucleico , Vacínia/enzimologia , Cinética , Modelos Químicos , Concentração Osmolar
13.
Biochemistry ; 34(43): 13994-4000, 1995 Oct 31.
Artigo em Inglês | MEDLINE | ID: mdl-7577997

RESUMO

The distribution of tritium derived from enzyme-bound [thiazole-2-T]thiamin diphosphate (TDP) during the reaction of pyruvate to form acetaldehyde catalyzed by pyruvate decarboxylase isozymes (PDC; EC 4.1.1.1) from Saccharomyces carlsbergensis was determined under single-turnover conditions ([E] > [S]) in the presence of the nonsubstrate allosteric effector pyruvamide. The specific radioactivity of the [1-L]acetaldehyde product and solvent ([L]H2O) was 43 +/- 4% and 54 +/- 2%, respectively, of the initial specific radioactivity of PDC-bound [thiazole-2-T]TDP and was independent of the extent of the single-turnover reaction. There is little (< or = 3%) or no return of the abstracted C(2)-hydron to the C(2) position of PDC-bound TDP. This provides evidence that the abstracted C(2)-hydron is involved in the specific protonation of the C(alpha) position of the PDC-bound intermediate 2-(1-hydroxyethyl)thiamin diphosphate (HETDP), which is cleaved to form [1-L]acetaldehyde and PDC-bound [thiazole-2-H]TDP. The partial exchange of C(2)-derived tritium into solvent requires that (1) hydron transfer from C(2) occurs to a catalytic-base in which the conjugate catalytic acid is partially shielded from hydron exchange with the solvent, (2) the conjugate catalytic acid transfers the C(2)-derived hydron to the C(alpha) position of HETDP, and (3) hydron transfer to C(2) to regenerate the coenzyme occurs either from solvent directly or from a second catalytic acid of the enzyme that undergoes rapid hydron exchange with the solvent.(ABSTRACT TRUNCATED AT 250 WORDS)


Assuntos
Acetaldeído/metabolismo , Isoenzimas/metabolismo , Prótons , Piruvato Descarboxilase/metabolismo , Saccharomyces/enzimologia , Tiamina Pirofosfato/metabolismo , Catálise , Ativação Enzimática
14.
Biochemistry ; 34(43): 14001-11, 1995 Oct 31.
Artigo em Inglês | MEDLINE | ID: mdl-7577998

RESUMO

Catalysis of proton transfer to thiamin diphosphate (TDP) and 2-(1-hydroxyethyl)thiamin diphosphate (HETDP) by pyruvate decarboxylase isozymes (PDC; EC 4.1.1.1) from Saccharomyces carlsbergensis was investigated by determining the solvent discrimination tritium isotope effect, (kH/kT)disc, on the reaction of pyruvate to form acetaldehyde in the presence of the nonsubstrate allosteric effector pyruvamide. The fractionation factors for TDP C(2)-L (phi C(2) = 0.98 +/- 0.06) and HETDP C(alpha)-L (phi C(alpha) = 1.01 +/- 0.07) (L = H or D) do not contribute significantly to observed enzymic isotopic discrimination. The value of (kH/kT)disc = 1.0 for reprotonation of TDP C(2)-L under single-turnover conditions ([E] > [S]) is consistent with C(2)-hydron transfer via a catalytic group (phi = 1) equilibrated with solvent. [1-L]Acetaldehyde formation under transient steady-state ([E] < [S]) conditions shows solvent discrimination tritium isotope effects that increase over the range (kH/kT)disc = 0.39 (single turnover) to 0.86 (ten turnovers). The 2-fold increase in the value of (kH/kT)disc for the [1-L]acetaldehyde product under steady-state compared to single-turnover conditions is attributed to a fractionation factor of phi 1 = 0.88 +/- 0.06 for the residue(s) involved in C(alpha)-hydron transfer to form HETDP. This provides evidence that catalysis of acetaldehyde formation by PDC involves specific protonation of both HETDP C(alpha)-L and TDP C(2)-L (phi 2 = 1.0 +/- 0.1) and requires at least two catalytic groups. Values of phi < or = 1 for protonation of TDP C(2)-L and HETDP C(alpha)-L provide no evidence that the exocyclic 4'-amino or -imino group (phi > or = 1.2) provides significant intramolecular catalysis in the enzyme-bound coenzyme.


Assuntos
Isoenzimas/metabolismo , Prótons , Piruvato Descarboxilase/metabolismo , Saccharomyces/enzimologia , Solventes , Acetaldeído/metabolismo , Catálise , Cinética , Concentração Osmolar
15.
Biochemistry ; 34(39): 12636-44, 1995 Oct 03.
Artigo em Inglês | MEDLINE | ID: mdl-7548014

RESUMO

Reconstitution of apo-pyruvate decarboxylase isozymes (PDC, EC 4.1.1.1) from Saccharomyces carlsbergensis was investigated by determination of the steady-state kinetics of the reaction with thiamin diphosphate (TDP) and Mg2+ in the presence and absence of substrate (pyruvate) or allosteric effector (pyruvamide). Reconstitution of the PDC isozyme mixture and alpha 4 isozyme (alpha 4-PDC) exhibits biphasic kinetics with 52 +/- 11% of the PDC reacting with k1 = (1.0 +/- 0.3) x 10(-2) s-1 and 48 +/- 12% of the PDC reacting with k2 = (1.1 +/- 0.6) x 10(-1) s-1 when TDP (KTDP = 0.5 +/- 0.2 mM) is added to apo-PDC equilibrated with saturating Mg2+. PDC reconstitution exhibits first-order kinetics with k1 = (1.6 +/- 0.5) x 10(-2) s-1 upon addition of Mg2+ (KMg2+ = 0.2 +/- 0.1 mM) to apo-PDC equilibrated with saturating TDP. Biphasic kinetics for the PDC isozymes provides evidence that apo-PDC reconstitution with TDP and Mg2+ involves two pathways, TDP binding followed by Mg2+ (k1) or Mg2+ binding followed by TDP (k2). This is supported by a change in reconstitution pathway with the order of cofactor addition and is inconsistent with a single pathway involving ordered binding of the metal ion followed by TDP. The presence of pyruvamide has no significant effect on the rate constants for apo-PDC reconstitution and favors the k2 pathway; pyruvate decreases the value of k2 < or = 3-fold and has no effect on the value of k1.(ABSTRACT TRUNCATED AT 250 WORDS)


Assuntos
Isoenzimas/química , Magnésio/química , Piruvato Descarboxilase/química , Saccharomyces cerevisiae/enzimologia , Tiamina Pirofosfato/química , Cinética
16.
Biochem J ; 303 ( Pt 1): 141-5, 1994 Oct 01.
Artigo em Inglês | MEDLINE | ID: mdl-7945232

RESUMO

Methanol dehydrogenase (MEDH) possesses tightly bound Ca2+ in addition to its pyrroloquinoline quinone prosthetic group. Ca2+ was replaced with Sr2+ by growing the host bacterium, Paracoccus denitrificans, in media in which Ca2+ was replaced with Sr2+. At temperatures in the transition region for stability, the rate constants for inactivation of MEDH purified from these cells (Sr-MEDH) were 2-fold lower than those for MEDH. However, Arrhenius plots yielded an activation energy (Ea) of 699 kJ (167 kcal)/mol for MEDH compared with 640 kJ (153 kcal)/mol for Sr-MEDH. Further analysis by transition-state theory yielded values for the activation enthalpy (delta H*) and activation entropy (delta S*) of 696 kJ (166 kcal)/mol and 1.73 kJ (414 cal)/mol per K for MEDH and 637 kJ (152 kcal)/mol and 1.55 kJ (371 cal)/mol per K for Sr-MEDH. The higher rate of inactivation of MEDH than Sr-MEDH at higher temperatures is a consequence of a more favourable net gain in entropy. This positive entropy contribution increases at high temperatures, and reduces the more favourable stability obtained from the enthalpy contribution for the free energy (delta G*) of inactivation. The differences in these thermodynamic data are discussed in relation to the recently determined crystal structure of MEDH as well as 1H electron-nuclear double resonance studies of the influence of Sr2+ substitution on the structure of the pyrroloquinoline quinone-derived radical in MEDH.


Assuntos
Oxirredutases do Álcool/química , Cálcio/metabolismo , Paracoccus denitrificans/enzimologia , Estrôncio/metabolismo , Oxirredutases do Álcool/efeitos dos fármacos , Oxirredutases do Álcool/metabolismo , Cálcio/farmacologia , Estabilidade Enzimática , Cinética , Ligação Proteica , Espectrofotometria , Estrôncio/farmacologia , Termodinâmica
17.
Biochemistry ; 33(42): 12600-8, 1994 Oct 25.
Artigo em Inglês | MEDLINE | ID: mdl-7918485

RESUMO

The quinoprotein methanol dehydrogenase and cytochrome c-551i are two soluble acidic proteins that form a physiological complex in which electrons are transferred from pyrroloquinoline quinone to heme. The oxidation of methanol dehydrogenase by the cytochrome was studied as a function of ionic strength using stopped-flow spectroscopy. The dissociation constant (Kd) for complex formation decreased 2-fold with increasing ionic strength from 0.21 to 1.3 M and increased at higher ionic strengths. The rate constant for the electron transfer reaction (kET) increased 2-fold with increasing ionic strength from 0.21 to 1.3 M and decreased at higher ionic strengths. The variation of Kd and kET over this range of ionic strengths was described by Van Leeuwen theory, which takes into account monopole-dipole and dipole-dipole forces, in addition to the monopole-monopole force, to predict the interactions between large molecules. Analysis of the kinetic results in terms of these electrostatic interactions indicated the probable orientations for protein-protein binding and electron transfer. To explain the ionic strength dependence of the observed kET, a model is presented in which the true kET is reduced by a factor Kc, an equilibrium constant that describes some rearrangement of the proteins after a nonoptimal collision to produce the most efficient orientation for electron transfer. This model is consistent with the notion that the large reorganizational energy obtained from temperature-dependence studies of this electron transfer reaction [Harris, T. K., & Davidson, V. L. (1993) Biochemistry 32, 14145-14150] is due to such an intracomplex rearrangement.(ABSTRACT TRUNCATED AT 250 WORDS)


Assuntos
Oxirredutases do Álcool/metabolismo , Grupo dos Citocromos c/metabolismo , Proteínas de Bactérias/metabolismo , Transporte de Elétrons/fisiologia , Cinética , Modelos Químicos , Modelos Moleculares , Concentração Osmolar , Paracoccus denitrificans/metabolismo , Conformação Proteica
18.
Biochem J ; 300 ( Pt 1): 175-82, 1994 May 15.
Artigo em Inglês | MEDLINE | ID: mdl-8198531

RESUMO

Methanol dehydrogenase (MEDH) possesses tightly bound Ca2+ in addition to its pyrroloquinoline quinone (PQQ) prosthetic group. Ca2+ was replaced with Sr2+ by growing the host bacterium, Paracoccus denitrificans, in media in which Ca2+ was replaced with Sr2+. MEDH, which was purified from these cells (Sr-MEDH), exhibited an increased absorption coefficient for the PQQ chromophore, and displayed certain kinetic properties which were different from those of native MEDH. Native MEDH exhibits an endogenous activity which is not stimulated by substrate and which is inhibited by cyanide. Sr-MEDH exhibited lower endogenous activity which was stimulated by substrate, and was much less sensitive to inhibition by cyanide. The Vmax. for the methanol-dependent activity of Sr-MEDH was 3-fold greater than that of the native enzyme, and the Ks for methanol was altered. Cyanide also acts as an obligatory activator and competitive inhibitor of methanol-dependent activity in native MEDH from P. denitrificans [Harris and Davidson (1993) Biochemistry 32, 4362-4368]. Sr-MEDH exhibited a similar K1 for cyanide inhibition of methanol-dependent activity, but the KA for cyanide activation of this activity was 17-fold greater than that for the native enzyme. The activation energy of Sr-MEDH was 13.4 kJ (3.2 kcal)/mol lower than that of the native enzyme. These data confirm and significantly extend the conclusions from genetic [Richardson and Anthony (1992) Biochem. J. 287, 709-715] and crystallographic [White, Boyd, Mathews, Xia, Dai, Zhang and Davidson (1993) Biochemistry 32, 12955-12958] studies that suggest an apparently unique role for Ca2+ in MEDH compared with other Ca(2+)-dependent proteins and enzymes.


Assuntos
Oxirredutases do Álcool/metabolismo , Cálcio/metabolismo , Estrôncio/metabolismo , Oxirredutases do Álcool/antagonistas & inibidores , Oxirredutases do Álcool/química , Amônia/química , Cinética , Paracoccus denitrificans/enzimologia , Cianeto de Potássio/química , Espectrofotometria Ultravioleta , Especificidade por Substrato , Temperatura
19.
Biochemistry ; 32(51): 14145-50, 1993 Dec 28.
Artigo em Inglês | MEDLINE | ID: mdl-8260498

RESUMO

The quinoprotein methanol dehydrogenase and cytochrome c-551i form a physiologic complex in which electrons are transferred from pyrroloquinoline quinone to heme. The reoxidation of methanol dehydrogenase by the cytochrome was studied by stopped-flow spectroscopy. The rate constant for the electron transfer reaction and the dissociation constant for complex formation were each determined at temperatures ranging from 20 to 50 degrees C. The electron transfer rates varied from 1.4 to 4.6 s-1. Analysis of the electron transfer reaction by Marcus theory yielded values of 1.9 eV for the reorganizational energy and 0.071 cm-1 for the electronic coupling and predicted a theoretical distance between redox centers of 15 A. Kinetically determined dissociation constants correlated well with a Kd of 375 microM which was determined in a direct ultrafiltration binding assay. Thermodynamic analysis of the dissociation constants indicated the importance of the hydrophobic effect in complex formation.


Assuntos
Oxirredutases do Álcool/química , Grupo dos Citocromos c/química , Oxirredutases do Álcool/metabolismo , Proteínas de Bactérias/química , Grupo dos Citocromos c/metabolismo , Cinética , Oxirredução , Paracoccus denitrificans/enzimologia , Análise Espectral , Temperatura , Termodinâmica
20.
Biochemistry ; 32(16): 4362-8, 1993 Apr 27.
Artigo em Inglês | MEDLINE | ID: mdl-8386543

RESUMO

The reactions of methanol dehydrogenase from Paracoccus denitrificans with artificial electron acceptors, ammonia, cyanide, and substrates have been characterized by steady-state kinetic analysis. Phenazine ethosulfate, a commonly used electron acceptor for this enzyme, was shown to exhibit pronounced substrate inhibition with a K(i) value approximately 20-fold lower than its Km. Wurster's Blue exhibited only relatively mild substrate inhibition and was deemed a more appropriate electron acceptor. Ammonia was an obligatory activator of the enzyme at low concentrations and inhibited a high concentrations. The K(i) value for this inhibition correlated closely with the Kd calculated from a titration of perturbations of the absorption spectrum of methanol dehydrogenase which were caused by the addition of ammonia. Cyanide, which suppressed the peculiar endogenous reaction of methanol dehydrogenase, was also both an activator of substrate-dependent activity and a competitive inhibitor with respect to methanol. Kinetic analysis indicated that the latter two activities corresponded to two distinct binding sites for cyanide. The Ka for cyanide activation correlated closely with the concentration required to inhibit 50% of the endogenous reaction, suggesting that a single binding event is responsible for both of these effects. A model is presented to describe the effects of ammonia and cyanide in the reaction cycle of methanol dehydrogenase, and the physiological relevance of the activation and inhibition by these compounds in vitro is discussed.


Assuntos
Oxirredutases do Álcool/metabolismo , Matemática , Modelos Teóricos , Paracoccus denitrificans/enzimologia , Coenzimas/metabolismo , Cinética , Cofator PQQ , Cianeto de Potássio/farmacologia , Quinolonas/metabolismo , Especificidade por Substrato
SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA