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 359 毫秒
1.
C A Gates  D B Northrop 《Biochemistry》1988,27(10):3834-3842
Aminoglycoside nucleotidyltransferase 2'-I follows a Theorell-Chance kinetic mechanism in which turnover is controlled by the rate-limiting release of the final product (Q), a nucleotidylated aminoglycoside [Gates, C. A., & Northrop, D. B. (1988) Biochemistry (second of three papers in this issue)]. The effects of viscosity on the kinetic constants of netilmicin, gentamicin C1, and sisomicin aminoglycoside substrates are as follows: no change in the substrate inhibition constants of all three antibiotics, a small but significant and highly unusual increase in Vmax/Km for netilmicin but large, normal decreases for gentamicin C1 and sisomicin, and marked decreases in the maximal velocities for all three. The lack of effect on substrate inhibition provides essential control experiments, signifying that glycerol does not interfere with binding of aminoglycosides to EQ and that the steady-state distribution of EQ does not increase as the release of Q is slowed by a viscosogen. The decrease in the Vmax/Km of better substrates indicates dominance by a diffusion-controlled component in the catalytic segment, attributed to the release of pyrophosphate. The presence of an increase in the Vmax/Km of the poor substrate, however, is inexplicable in terms of either single or multiple diffusion-controlled steps. Instead, it is here attributed to an equilibrium between conformers of the enzyme-nucleotide complex in which glycerol favors the conformation necessary for binding of aminoglycosides. The decrease in Vmax is consistent with the diffusion-controlled release of the final product determining enzymatic turnover.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
C A Gates  D B Northrop 《Biochemistry》1988,27(10):3820-3825
Aminoglycoside nucleotidyltransferase 2'-I (formerly gentamicin adenylyltransferase) conveys antibiotic resistance to Gram-negative bacteria by transfer of AMP to the 2'-hydroxyl group of 4,6-substituted deoxystreptamine-containing aminoglycosides. The kinetics constants of thirteen aminoglycoside antibiotics and the magnesium chelates of eight nucleotide triphosphates were determined with purified enzyme. Eleven of the antibiotics exhibit substrate inhibition attributed to secondary binding of the aminoglycoside to an enzyme-AMP-aminoglycoside complex. Maximal velocities vary by only 4-fold, versus variation of values of Vmax/Km for the aminoglycosides of nearly 4000-fold, consistent with a Theorell-Chance kinetic mechanism as proposed for this enzyme [Gates, C. A., & Northrop, D. B. (1988) Biochemistry (second of three papers in this issue)] with the added specification that the binding of aminoglycosides is in rapid equilibrium. Under these conditions, Vmax/Km becomes kcat/Kd, where kcat is the net rate constant for catalysis (but not turnover) and Kd is the dissociation constant of aminoglycosides from a complex with enzyme and nucleotide. Values of kcat fall closely together into three distinct sets, with the 3',4'-dideoxygentamicins greater than gentamicins greater than kanamycins. These sets reflect unusual structure-activity correlations which are specific for catalysis but have nothing to do with the maximal velocity of this enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Pyruvate has previously been shown to slow down the rate of intramolecular electron transfer from the flavosemiquinone (Fs) to the cytochrome b2 moiety of flavocytochrome b2 [Tegoni, M., Silvestrini, M. C., Labeyrie, F. & Brunori, M. (1984) Eur. J. Biochem. 140, 39-45] and to stabilize markedly the Fs state of the prosthetic flavin, relative to the oxidized (Fo) and the reduced (Fh) states [Tegoni, M., Janot, J. M. & Labeyrie, F. (1986) Eur. J. Biochem. 155, 491-503]. In the present study, we have determined the dissociation constants of pyruvate for the three redox forms of the prosthetic flavin and demonstrated that the Fs-pyruvate complex is actually much more stable than the Fo-pyruvate and Fh-pyruvate complexes. The inhibition produced by pyruvate has been characterized under steady-state conditions using both ferricytochrome c and ferricyanide as external acceptor. A detailed analysis and simulations of the suitable reaction scheme, taking into consideration all data from rapid kinetic studies of partial reactions previously published, show that the experimental noncompetitive inhibition results from the sum of a competitive effect due to binding of pyruvate to Fo and an uncompetitive effect due to binding to the Fs intermediate in a dead-end complex. Pyruvate binding to the semiquinone transient results in a marked loss of the reactivity of this donor in electron transfers to its specific partner, the cytochrome b2 present in the same active site, as to ferricyanide, an external acceptor. A critical evaluation of the parameters involved in the control of such reactivities is presented.  相似文献   

4.
Kinetic data for the 6'-aminoglycoside-modifying enzyme, AAC(6')-4, also named kanamycin acetyltransferase, have been collected for six aminoglycoside antibiotics (amikacin, gentamicin C1a, kanamycin A, neomycin B, sisomicin, and tobramycin) and three coenzymes (acetyl-CoA, n-propionyl-CoA, and n-butyryl-CoA). The initial velocity pattern using acetyl-CoA favors a ping-pong kinetic mechanism (Kia = -0.34 +/- 0.34 microM), but the pattern using n-propionyl-CoA supports a sequential one (Kia = 2.7 +/- 0.8 microM). Kinetic analyses using alternative substrates confirm the sequential mechanism and, moreover, clearly identify a random order of addition of antibiotic and coenzyme because V/K values of antibiotics varied 40-fold as the identity of the coenzyme was changed and V/K values of coenzyme varied 13-fold as the identity of the antibiotic was changed. One or both sets of values would have remained unchanged if the mechanism were either ordered sequential or ping-pong. Combining these results with structure-activity data which argue for a rapid rate of release of substrates and products relative to the rate of enzymatic turnover (Radika, K., and Northrop, D. B. (1984) Biochemistry 25, in press) establishes the kinetic mechanism as rapid equilibrium random sequential.  相似文献   

5.
Aminoglycoside nucleotidyltransferase (2')-Ia [ANT (2')-Ia] was cloned from Pseudomonas aeruginosa and purified from overexpressing Escherichia coli BL21(DE3) cells. The first enzyme-bound conformation of an aminoglycoside antibiotic in the active site of an aminoglycoside nucleotidyltransferase was determined using the purified aminoglycoside nucleotidyltransferase (2' ')-Ia. The conformation of the aminoglycoside antibiotic isepamicin, a psuedo-trisaccharide, bound to aminoglycoside nucleotidyltransferase (2' ')-Ia has been determined using NMR spectroscopy. Molecular modeling, employing experimentally determined interproton distances, resulted in two different enzyme-bound conformations (conformer 1 and conformer 2) of isepamicin. Conformer 1 was by far the major conformer defined by the following average glycosidic dihedral angles: PhiBC = -65.26 +/- 1.63 degrees and PsiBC = -54.76 +/- 4.64 degrees. Conformer 1 was further subdivided into one major (conformer 1a) and two minor components (conformers 1b and 1c) based on the comparison of glycosidic dihedral angles PhiAB and PsiAB. The arrangement of substrates in the enzyme.metal-ATP.isepamicin complex was determined on the basis of the measured effect of the paramagnetic substrate analogue Cr(H2O)4ATP on the relaxation rates of substrate protons which were used to determine relative distances of isepamicin protons to the Cr3+. Both conformers of isepamicin yielded arrangements that satisfied the NOE restraints and the observed paramagnetic effects of Cr(H2O)4ATP. It has been suggested that aminoglycosides use both electrostatic interactions and hydrogen bonds in binding to RNA and that the contacts made by the A and B rings to RNA are the most important for binding [Fourmy, D., Recht, M. I., Blanchard, S. C., and Puglisi, J. D. (1996) Science 274, 1367-1371]. Comparisons based on the determined conformations of enzyme-bound aminoglycoside antibiotics also suggested that interactions of rings A and B with enzymes may be the major determinant in aminoglycoside binding to enzymes [Serpersu, E. H., Cox, J. R., DiGiammarino, E. L., Mohler, M. L., Ekman, D. R., Akal-Strader, A., and Owston, M. (2000) Cell Biochem. Biophys. (in press)]. The conformation of isepamicin bound to the aminoglycoside nucleotidyltransferase (2' ')-Ia, determined in this work, lent further support to this theory. Furthermore, comparison of enzyme-bound conformations of isepamicin to the RNA-bound conformation of gentamycin C1a also showed remarkable similarities between the enzyme-bound and RNA-bound aminoglycoside antibiotic conformations. These studies should aid in the design of effective inhibitors possessing a broad range of aminoglycoside-modifying enzymes as targets.  相似文献   

6.
The GDP/GTP nucleotide exchange of Arf1 is catalyzed by nucleotide exchange factors (GEF), such as Arno, which act through their catalytic Sec7 domain. This exchange is a complex mechanism that undergoes conformational changes and intermediate complex species involving several allosteric partners such as nucleotides, Mg2+, and Sec7 domains. Using a surface plasmon resonance approach, we characterized the kinetic binding parameters for various intermediate complexes. We first confirmed that both GDP and GTP counteract equivalently to the free-nucleotide binary Arf1-Arno complex stability and revealed that Mg2+ potentiates by a factor of 2 the allosteric effect of GDP. Then we explored the uncompetitive inhibitory mechanism of brefeldin A (BFA) that conducts to an abortive pentameric Arf1-Mg2+-GDP-BFA-Sec7 complex. With BFA, the association rate of the abortive complex is drastically reduced by a factor of 42, and by contrast, the 15-fold decrease of the dissociation rate concurs to stabilize the pentameric complex. These specific kinetic signatures have allowed distinguishing the level and nature as well as the fate in real time of formed complexes according to experimental conditions. Thus, we showed that in the presence of GDP, the BFA-resistant Sec7 domain of Arno can also associate to form a pentameric complex, which suggests that the uncompetitive inhibition by BFA and the nucleotide allosteric effect combine to stabilize such abortive complex.  相似文献   

7.
Coupled spectrofluorometric assay for aminoglycoside phosphotransferases   总被引:2,自引:0,他引:2  
In order to make accurate kinetic measurements for the substrates of aminoglycoside (AG) phosphotransferases (APHs), we have developed an assay which overcomes many of the limitations of currently used assays. We have adapted the coupled spectrophotometric assay (P. R. Goldman and D. B. Northrop (1976) Biochem. Biophys. Res. Commun. 68, 230-236) for use in a spectrofluorometer. At an excitation wavelength of 340 nm, NADH will emit an intensity peak at 450 nm; NAD does not emit under these conditions. Our assay can accurately measure differences of 0.25 microM. For the APH(3')-II encoded on Tn5, we have redetermined the Km's for the AGs, amikacin (AK), kanamycin (KM), and ribostamycin (Rib), and for ATP. Our values for AK (76 microM) were lower than those derived from the spectrophotometric assay; for KM and Rib we obtained Km values of 5.1 and 3.6 microM, respectively. These values were well below the limit of accuracy (10 microM) for the spectrophotometric assay. In addition, we have begun characterization of an APH from a clinical isolate with a low Km for AK. Thus far, we have determined that this enzyme has Km's of approximately 1 microM for both AK and KM. These results show that the assay is well suited for accurate determinations of kinetic constants for low Km substrates of APH enzymes.  相似文献   

8.
Bachmann BO  Townsend CA 《Biochemistry》2000,39(37):11187-11193
Streptomyces clavuligerus beta-lactam synthetase (beta-LS) was recently demonstrated to catalyze an early step in clavulanic acid biosynthesis, the ATP/Mg(2+)-dependent intramolecular closure of the beta-amino acid N(2)-(carboxyethyl)-L-arginine (CEA) to the monocyclic beta-lactam deoxyguanidinoproclavaminic acid (DGPC). Here we investigate the steady-state kinetic mechanism of the beta-LS-catalyzed reaction to better understand this unprecedented secondary metabolic enzyme. Initial velocity patterns were consistent with a sequential ordered bi-ter kinetic mechanism. Product inhibition studies with PP(i) and DGPC demonstrated competitive inhibition versus their cognate substrates ATP and CEA, respectively, and noncompetitive inhibition against their noncognate substrates. To clarify the order of substrate binding, the truncated substrate analogue N(2)-(carboxymethyl)-L-arginine was synthesized and demonstrated uncompetitive inhibition versus ATP and competitive patterns versus CEA. These data are consistent with ordered substrate binding, with ATP binding first, an abortive enzyme-DGPC complex, and PP(i) released as the last product. The pH dependence of V and V/K was determined and suggests that residues with a pK of 6.5 and 9.3 must be ionized for optimal activity. These observations were considered in the context of investigations of the homologous primary metabolic enzyme asparagine synthetase B, and a chemical mechanism is proposed that is consistent with the kinetic mechanism.  相似文献   

9.
The kinetic mechanism of pigeon liver fatty acid synthetase action has been studied using steady state kinetic analysis. Initial velocity studies are consistent with an earlier suggestion that the enzyme catalyzes this reaction by a seven-site ping-pong mechanism. Although the range of substrate concentrations that could be used was limited by several factors, the initial velocity patterns showing the relationship between the substrates acetyl coenzyme CoA, malonyl-CoA, and NADPH appear to be a series of parallel lines, regardless of which substrate is varied at fixed levels of a second substrate. However, two of the substrates, acetyl-CoA and malonly-CoA, apparently exhibit a competitive substrate inhibition with respect to each other, but NADPH shows no inhibition of any kind. Product inhibition patterns suggest that free CoA is competitive versus acetyl-CoA and malonyl-CoA and is uncompetitive versus NADPH, and that NADP+ is competitive versus NADPH and uncompetitive versus acetyl-CoA or malonyl-CoA. These results are consistent with a seven-site ping-pong mechanism with intermediates covalently bound to 4'-phosphopantetheine (part of acyl carrier protein). Double competitive substrate inhibition by acetyl-CoA and malonyl-CoA is consistent with the rate equation derived for the over-all mechanism. The kinetic mechanism developed from these results is capable of explaining the formation of fatty acids from malonyl-CoA and NADPH alone (Katiyar, S. S., Briedis, A. V., and Porter, J. W. (1974) Arch. Biochem. Biophys. 162, 412-420) and also the formation of triacetic acid lactone from either malonyl-CoA alone or acetyl-CoA plus malonyl-CoA.  相似文献   

10.
The effects of alpha-, beta- and gamma-cyclodextrins on the amylose and maltopentaose hydrolysis catalysed by porcine pancreatic alpha-amylase (PPA) were investigated. The results of the statistical analysis performed on the kinetic data using the general initial velocity equation of a one-substrate reaction in the presence of one inhibitor indicate that the type of inhibition involved depends on the substrate used: the inhibition of amylose hydrolysis by alpha-, beta- and gamma-cyclodextrin is of the competitive type, while the inhibition of maltopentaose hydrolysis is of the mixed noncompetitive type. Consistently, the Lineweaver-Burk plots intersect on the vertical axis when amylose is used as the substrate, while in the case of maltopentaose, the intersection occurs at a point located in the second quadrant. The inhibition of the hydrolysis therefore involves only one abortive complex, PPA-cyclodextrin, when amylose is used as the substrate, while two abortive complexes, PPA-cyclodextrin and PPA-maltopentaose-cyclodextrin, are involved with maltopentaose. The mixed noncompetitive inhibition thus shows the existence of one accessory binding site. In any case, only one molecule of inhibitor binds to PPA. In line with these findings, the difference spectra of PPA produced by alpha-, beta- and gamma-cyclodextrin indicate that binding occurs at a tryptophan and a tyrosine residue. The corresponding dissociation constants and the inhibition constants obtained using the kinetic approach are in the same range (1.2-7 mM). The results obtained here on the inhibition of maltopentaose hydrolysis by cyclodextrin are similar to those previously obtained with acarbose as the inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Prodanov, E. & Santimone, M. (1998) Eur. J. Biochem. 252, 100-107], but differ from those obtained with amylose as the substrate and acarbose as inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Payan, F., Forest, E. & Santimone, M. (1996) Eur. J. Biochem. 241, 787-796]. It is concluded that the hydrolysis of both long and short chain substrates requires at least one secondary binding site, including a tryptophan residue.  相似文献   

11.
In the direction of reductive condensation of alpha-ketoglutarate and lysine, saccharopine dehydrogenase (N6-(glutar-2-yl)-L-lysine:NAD oxidoreductase (lysine-forming) is inhibited by high concentrations of alpha-ketoglutarate and lysine, but not by NADH. NAD+ and saccharopine show no substrate inhibition in the reverse direction. Substrate inhibition by alpha-ketoglutarate and lysine is linear uncompetitive versus NADH. However, when the inhibition is examined with alpha-ketoglutarate or lysine as the variable substrate, the double reciprocal plots show a family of curved lines concave up. The curvature is more pronounced with increasing concentrations of the inhibitory substrate, suggesting an interaction of variable substrate with the enzyme form carrying the inhibitory substrate. These inhibition patterns, the lack of interaction of structural analogs of lysine such as ornithine and norleucine with the E-NAD+ complex (Fujioka M., and Nakatani, Y. (1972) Eur. J. Biochem. 25, 301-307), the identity of values of inhibition constants of alpha-ketoglutarate and lysine obtained with either one as the substrate inhibitor, and the substrate inhibition data in the presence of a reaction product, NAD+, are consistent with the mechanism that substrate inhibition results from the formation of a dead-end E-NAD+-alpha-ketoglutarate complex followed by the addition of lysine to this abortive complex.  相似文献   

12.
The inhibition of human carbonic anhydrase II (carbonate hydro-lyase, EC 4.2.1.1) by tetrazole, 1,2,4-triazole, 2-nitrophenol, and chloral hydrate has been investigated. These inhibitors, together with phenol which has been studied previously (Simonsson, I., Jonsson, B.-H. and Lindskog, S. (1982) Biochem. Biophys. Res. Commun. 108, 1406-1412), can be classified in three groups depending upon the kinetic patterns of inhibition of CO2 hydration at pH near 9. The first group, represented by tetrazole and 2-nitrophenol, yields predominantly uncompetitive inhibition under these conditions in analogy with simple, inorganic anions. The second group, represented by 1,2,4-triazole and chloral hydrate gives rise to essentially noncompetitive inhibition patterns, whereas phenol, representing the third group, is a competitive inhibitor of CO2 hydration. These diverse inhibition patterns are discussed in terms of the kinetic mechanism scheme originally proposed by Steiner et al. (Steiner, H., Jonsson, B.-H. and Lindskog, S. (1975) Eur. J. Biochem. 59, 253-259.  相似文献   

13.
Ginet N  Lavergne J 《Biochemistry》2001,40(6):1812-1823
The equilibrium and kinetic parameters for the binding of various inhibitors to the Q(B) pocket of the bacterial reaction center were investigated in chromatophores from Rhodobacter capsulatus and Rhodobacter sphaeroides. By monitoring the near-IR absorption changes specific to Q(A)(-) and Q(B)(-), we measured the fraction of inhibited centers in the dark and the kinetics and extent of inhibitor displacement after one flash due to the formation of the Q(A)Q(B)(-) state. The inhibitor release rate was much faster for triazines and o-phenanthroline (t(1/2) in the 50 ms to 1 s range) than for stigmatellin (t(1/2) approximately 20 s). For inhibitors with a rapid release rate, the fast phase of P(+) decay observed in the absence of secondary donor reflects the competition between P(+)Q(A)(-) recombination and inhibitor release: it is thus faster than the P(+)Q(A)(-) recombination, and its relative extent is smaller than the fraction of initially inhibited centers. At appropriate inhibitor concentrations, one can have almost total binding in the dark and almost total inhibitor displacement after one flash. Under such conditions, a pair of closely spaced flashes resets the two-electron gate in a single state (Q(A)Q(B)(-)), irrespective of the initial state. The apparent dissociation constant of terbutryn was significantly increased (by a factor of 4-7) in the presence of Q(A)(-), in agreement with the conclusion of Wraight and co-workers [Stein, R. R., et al. (1984) J. Cell. Biochem. 24, 243-259]. We suggest that this effect is essentially due to a tighter binding of ubiquinone in the Q(A)(-) state.  相似文献   

14.
Porcine pancreatic alpha-amylase (PPA) is inhibited by the red kidney bean (Phaseolus vulgaris) inhibitor alpha-AI1 [Eur. J. Biochem. 265 (1999) 20]. Inhibition kinetics were carried out using DP 4900-amylose and maltopentaose as substrate. As shown by graphical and statistical analysis of the kinetic data, the inhibitory mode is of the mixed noncompetitive type whatever the substrate thus involving the EI, EI2, ESI and ESI2 complexes. This contrast with the E2I complex obtained in the crystal and with biophysical studies. Such difference very likely depends on the [I]/[E] ratio. At low ratio, the E2I complex is favoured; at high ratio the EI, ESI and EI2 complexes are formed. The inhibition model also differs from those previously proposed for acarbose [Eur. J. Biochem. 241 (1996) 787 and Eur. J. Biochem. 252 (1998) 100]. In particular, with alpha-AI1, the inhibition takes place only when PPA and alpha-AI are preincubated together before adding the substrate. This indicates that the abortive PPA-alphaAI1 complex is formed during the preincubation period. One additional carbohydrate binding site is also demonstrated yielding the ESI complex. Also, a second protein binding site is found in EI2 and ESI2 abortive complexes. Conformational changes undergone by PPA upon alpha-AI1 binding are shown by higher sensitivity to subtilisin attack. From X-ray analysis of the alpha-AI1-PPA complex (E2I), the major interaction occurs with two hairpin loops L1 (residues 29-46) and L2 (residues 171-189) of alpha-AI1 protruding into the V-shaped active site of PPA. The hydrolysis of alpha-AI1 that accounts for the inhibitory activity is reported.  相似文献   

15.
K Radika  D B Northrop 《Biochemistry》1984,23(22):5118-5122
Antibiotic resistance caused by the presence of the plasmid pMH67 is mediated by the aminoglycoside acetyltransferase AAC(6')-4, also known as kanamycin acetyltransferase. Bacteria harboring the plasmid are resistant to the kanomycins plus a broad range of other deoxystreptamine-containing aminoglycosides but not to the gentamicins XK62-2 and C1 which are substituted at the 6'-position. Substrate specificity studies on the purified enzyme, however, now show that the enzyme acetylates an even broader range of aminoglycosides, including the gentamicins XK62-2 and C1. The enzyme also accepts several acyl-CoA esters, which differ in nucleotide as well as in acyl chain length. Application of the method of analysis of structure-activity data developed earlier for gentamicin acetyltransferase [Williams, J. W., & Northrop, D. B. (1978) J. Biol. Chem. 253, 5908-5914] to the kinetic data obtained for AAC(6')-4 shows that the turnover of the acylation reaction is limited by catalysis and not by the rate of release of either the acetylated antibiotic or CoA. Most structural changes in aminoglycosides cause changes in rates of release, and only drastic changes, near the 6'-amino group, affect catalysis. The structural requirements on aminoglycosides for enzymatic activity run parallel to the structural requirements for antibacterial activity.  相似文献   

16.
Bacterial resistance to the aminoglycoside antibiotics is manifested primarily by enzymic modification of these drugs. One important mechanism of streptomycin modification is through ATP-dependent O-adenylation, catalyzed by streptomycin adenylyltransferase. Initial velocity patterns deduced from steady state kinetics indicate a sequential mechanism. Dead-end inhibition by tobramycin and neomycin is non-competitive versus streptomycin and uncompetitive versus ATP, indicative of ordered substrate binding where ATP binds first and then streptomycin. These results surmise that streptomycin adenylyltransferase follows an ordered, sequential kinetic mechanism in which one substrate (ATP) binds prior to the antibiotic and pyrophosphate is released prior to formation of AMP-streptomycin.  相似文献   

17.
Saturation curves of activity versus concentration were determined for aspartate transcarbamylase from Escherichia coli (EC 2.1.3.2) for the substrate L-aspartate at saturating carbamyl phosphate (4.8 mM) in buffered solution at pH values from 6.0 to 12.0. Hill coefficients were obtained from the sigmoidal curves. At pH values from 7.8 to 9.1, where substrate inhibition causes difficulties in the Hill approximation, our kinetic scheme includes substrate inhibition and residual activity in the abortive enzyme-substrate complex. The plot of Hill coefficient versus pH has pKalpha values of 7.4 and 9.8 at the half-maximum positions of the curve which has a plateau from pH 8.1 to 9.1. These pKalpha values may be associated with functional groups involved in the allosteric transition which activates the enzyme. A plot of [S]0.5 versus pH shows a pKalpha of 8.5, which may belong to a residue either at or near the aspartate binding site. At 50 mM aspartate concentration the pH-rate profile shows maxima at pH values of 8.8 and 10.0 (cf. Weitzman, P.D.J., and Wilson, I.B.(1966)J. Biol. Chem. 2418 5481-5488, who used 100 mM aspartate). However, when the pH-dependent substrate inhibition is included, the calculated Vmax--H curve is bell-shaped like that of the isolated catalytic subunit.  相似文献   

18.
The intrinsic primary deuterium isotope effect for the O-deethylation of 7-ethoxycoumarin has been estimated by the Northrop method [D. B. Northrop (1977) in Isotope Effects on Enzyme-Catalyzed Reactions (Cleland, W. W., O'Leary, M. H., and Northrop, D. B., eds.), pp. 122-152, University Park Press, Baltimore] for the microsomal cytochrome P-448 system from 3-methylcholanthrene-induced hamster livers. The intrinsic isotope effect (Dk = 5.5) was found to be equivalent to the observed deuterium isotope effect, demonstrating that the isotope effect for this reaction was fully expressed by this cytochrome P-448 system. These data unequivocally demonstrate that C-H bond cleavage is the rate-limiting step in the overall reaction catalyzed by this system. The decrease in the rate of product formation, occurring as a consequence of deuterium substitution, resulted in a reduction in the quantity of substrate metabolized but was not accompanied by the change in regiospecificity observed in previous studies with a hepatic cytochrome P-448 isozyme purified from 3-methylcholanthrene-induced rats. These data demonstrate that the catalytic site of the hamster isozyme(s) offers more constraints to 7-ethoxycoumarin reorientation than does the catalytic site of the rat liver isozyme.  相似文献   

19.
The inhibition of the major form of ox kidney aldehyde reductase (AR 1) by sodium barbitone revealed linear mixed kinetics. This behaviour is distinct from the non-linear intercept effect we reported for valproate [Daly and Mantle (1982) Biochem. J. 205, 381]. 4-Carboxybenzaldehyde exhibits partial uncompetitive substrate inhibition. These results are discussed in terms of a model that involves nucleotide-induced isomerization and an additional flux (with some substrates and inhibitors) through an enzyme.nucleotide.substrate/inhibitor ternary complex.  相似文献   

20.
D Valenzuela  L H Schulman 《Biochemistry》1986,25(16):4555-4561
Four different structural regions of Escherichia coli tRNAfMet have been covalently coupled to E. coli methionyl-tRNA synthetase (MetRS) by using a tRNA derivative carrying a lysine-reactive cross-linker. We have previously shown that this cross-linking occurs at the tRNA binding site of the enzyme and involves reaction of only a small number of the potentially available lysine residues in the protein [Schulman, L. H., Valenzuela, D., & Pelka, H. (1981) Biochemistry 20, 6018-6023; Valenzuela, D., Leon, O., & Schulman, L. H. (1984) Biochem. Biophys. Res. Commun. 119, 677-684]. In this work, four of the cross-linked peptides have been identified. The tRNA-protein cross-linked complex was digested with trypsin, and the peptides attached to the tRNA were separated from the bulk of the tryptic peptides by anion-exchange chromatography. The tRNA-bound peptides were released by cleavage of the disulfide bond of the cross-linker and separated by reverse-phase high-pressure liquid chromatography, yielding five major peaks. Amino acid analysis indicated that four of these peaks contained single peptides. Sequence analysis showed that the peptides were cross-linked to tRNAfMet through lysine residues 402, 439, 465, and 640 in the primary sequence of MetRS. Binding of the tRNA therefore involves interactions with the carboxyl-terminal half of MetRS, while X-ray crystallographic data have shown the ATP binding site to be located in the N-terminal domain of the protein [Zelwer, C., Risler, J. L., & Brunie, S. (1982) J. Mol. Biol. 155, 63-81].(ABSTRACT TRUNCATED AT 250 WORDS)  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