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.
In an effort to test whether a transition state analog is an inhibitor of the metallo-β-lactamases, a phospholactam analog of carbapenem has been synthesized and characterized. The phospholactam 1 proved to be a weak, time-dependent inhibitor of IMP-1 (70%), CcrA (70%), L1 (70%), NDM-1 (53%), and Bla2 (94%) at an inhibitor concentration of 100 μM. The phospholactam 1 activated ImiS and BcII at the same concentration. Docking studies were used to explain binding and to offer suggestions for modifications to the phospholactam scaffold to improve binding affinities.  相似文献   

2.
The first report of slow-tight inhibition of xylanase by a bifunctional inhibitor alkalo-thermophilic Bacillus inhibitor (ATBI), from an extremophilic Bacillus sp. is described. ATBI inhibits aspartic protease (Dash, C., and Rao, M. (2001) J. Biol. Chem., 276, 2487-2493) and xylanase (Xyl I) from a Thermomonospora sp. The steady-state kinetics revealed time-dependent competitive inhibition of Xyl I by ATBI, consistent with two-step inhibition mechanism. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI*, and the dissociation of EI* were 13 +/- 1 x 10(-6) s(-1) and 5 +/- 0.5 x 10(-8) s(-1), respectively. The K(i) value for the formation of EI complex was 2.5 +/- 0.5 microm, whereas the overall inhibition constant K(i)* was 7 +/- 1 nm. The conformational changes induced in Xyl I by ATBI were monitored by fluorescence spectroscopy and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. ATBI binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde (OPTA)-labeled Xyl I. Our results revealed that the inactivation of Xyl I is due to the disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis and a model depicting the probable interaction between ATBI or OPTA with Xyl I has been proposed.  相似文献   

3.
The kinetics of slow onset inhibition of Proteinase K by a proteinaceous alkaline protease inhibitor (API) from a Streptomyces sp. is presented. The kinetic analysis revealed competitive inhibition of Proteinase K by API with an IC50 value 5.5 +/- 0.5 x 10-5 m. The progress curves were time-dependent, consistent with a two-step slow tight binding inhibition. The first step involved a rapid equilibrium for formation of reversible enzyme-inhibitor complex (EI) with a Ki value 5.2 +/- 0.6 x 10-6 m. The EI complex isomerized to a stable complex (EI*) in the second step because of inhibitor-induced conformational changes, with a rate constant k5 (9.2 +/- 1 x 10-3 s-1). The rate of dissociation of EI* (k6) was slower (4.5 +/- 0.5 x 10-5 s-1) indicating the tight binding nature of the inhibitor. The overall inhibition constant Ki* for two-step inhibition of Proteinase K by API was 2.5 +/- 0.3 x 10-7 m. Time-dependent dissociation of EI* revealed that the complex failed to dissociate after a time point and formed a conformationally altered, irreversible complex EI**. These conformational states of enzyme-inhibitor complexes were characterized by fluorescence spectroscopy. Tryptophanyl fluorescence of Proteinase K was quenched as a function of API concentration without any shift in the emission maximum indicating a subtle conformational change in the enzyme, which is correlated to the isomerization of EI to EI*. Time-dependent shift in the emission maxima of EI* revealed the induction of gross conformational changes, which can be correlated to the irreversible conformationally locked EI** complex. API binds to the active site of the enzyme as demonstrated by the abolished fluorescence of 5-iodoacetamidofluorescein-labeled Proteinase K. The chemoaffinity labeling experiments lead us to hypothesize that the inactivation of Proteinase K is because of the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the catalytic triad and other residues involved in catalysis.  相似文献   

4.
The present article reports a low molecular weight aspartic protease inhibitor, API, from a newly isolated thermo-tolerant Bacillus licheniformis. The inhibitor was purified to homogeneity as shown by rp-HPLC and SDS-PAGE. API is found to be stable over a broad pH range of 2-11 and at temperature 90 degrees C for 2 1/2h. It has a Mr (relative molecular mass) of 1363 Da as shown by MALDI-TOF spectra and 1358 Da as analyzed by SDS-PAGE .The amino acid analysis of the peptide shows the presence of 12 amino acid residues having Mr of 1425 Da. The secondary structure of API as analyzed by the CD spectra showed 7% alpha-helix, 49% beta-sheet and 44% aperiodic structure. The Kinetic studies of Pepsin-API interactions reveal that API is a slow-tight binding competitive inhibitor with the IC(50) and Ki values 4.0 nM and (3.83 nM-5.31 nM) respectively. The overall inhibition constant Ki* value is 0.107+/-0.015 nM. The progress curves are time-dependent and consistent with slow-tight binding inhibition: E+I -->/<-- (k(4), k(5)) EI -->/<-- (k(6), k(7)) EI*. Rate constant k(6)=2.73+/-0.32 s(-1) reveals a fast isomerization of enzyme-inhibitor complex and very slow dissociation as proved by k(7)=0.068+/-0.009 s(-1). The Rate constants from the intrinsic tryptophanyl fluorescence data is in agreement with those obtained from the kinetic analysis; therefore, the induced conformational changes were correlated to the isomerization of EI to EI*.  相似文献   

5.
This is the first report that describes the inhibition mechanism of xylanase from Thermomonospora sp. by pepstatin A, a specific inhibitor toward aspartic proteases. The kinetic analysis revealed competitive inhibition of xylanase by pepstatin A with an IC50 value 3.6 +/- 0.5 microm. The progress curves were time-depended, consistent with a two-step slow tight binding inhibition. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI* and the dissociation of EI* were 15 +/- 1 x 10(-5) and 3.0 +/- 1 x 10(-8) s(-1), respectively. The Ki value for the formation of EI complex was 1.5 +/- 0.5 microm, whereas the overall inhibition constant Ki* was 28.0 +/- 1 nm. The conformational changes induced in Xyl I by pepstatin A were monitored by fluorescence spectroscopy, and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. Pepstatin A binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde-labeled xylanase. Our results revealed that the inactivation of xylanase is due to the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis, and a model depicting the probable interaction between pepstatin A with xylanase has been proposed.  相似文献   

6.
A Betz  P W Wong  U Sinha 《Biochemistry》1999,38(44):14582-14591
Recently, peptidylketothiazoles have been shown to be potent inhibitors of proteases, but the details of the interaction have not yet been studied. In the work presented here, the interaction of factor Xa, a coagulation protease, with the transition state inhibitor BnSO(2)-D-Arg-Gly-Arg-ketothiazole (C921-78) is characterized. C921-78 is a tight and selective inhibitor of the coagulation protease factor Xa (K(d) = 14 pM). The hydrolytic activity of factor Xa was inhibited by C921-78 in a time-dependent manner. The rate-limiting step of the bimolecular combination of inhibitor and enzyme was competitive with the substrate. Conversely, the inhibitor could be displaced from the active site of the enzyme after exposure of the preformed complex to an excess of substrate or to the active site inhibitor dansyl-Glu-Gly-Arg-chloromethyl ketone (DEGR-CMK) in a slow reaction. The formation of the C921-78-factor Xa complex resulted in a 60% increase in the magnitude of the fluorescence emission spectrum. Rapid mixing of the enzyme and inhibitor produces a monophasic fluorescence increase, compatible with spectral transition in a single step. The rate constant for this reaction increased hyperbolically with the concentration of C921-78, but the amplitude remained constant. These results are consistent with the initial formation of an enzyme-inhibitor complex (EI), followed by a unimolecular conversion of EI to EI linked to a spectral transition. The rate constants of the isomerization provide an estimate of 300000-fold stabilization. Thus, the inhibition of factor Xa by C921-78 follows a mechanism similar to that described classically for slow tight binding inhibitors. However, the two steps of the reaction cannot be kinetically separated by the rapid equilibrium assumption, and therefore, the formation of EI is partially rate-limiting, too. The driving energy for the unusually fast isomerization step may result from the highly favorable interactions of the inhibitor in the primary binding site.  相似文献   

7.
Polypeptide deformylase (PDF) is an essential bacterial metalloenzyme responsible for the removal of the N-formyl group from the N-terminal methionine of nascent polypeptides. Inhibition of bacterial PDF enzymes by actinonin, a naturally occurring antibacterial agent, has been characterized using steady-state and transient kinetic methods. Slow binding of actinonin to these enzymes is observed under steady-state conditions. Progress curve analysis is consistent with a two-step binding mechanism, in which tightening of the initial encounter complex (EI) results in a final complex (EI*) with an extremely slow, but observable, off-rate (t(1/2) for inhibitor dissociation >or=0.77 days). Stopped-flow measurement of PDF fluorescence confirms formation of EI and provides a direct measurement of the association rate. Rapid dilution studies establish that the potency of actinonin is enhanced by more than 2000-fold upon tightening of EI to form EI*, from K(i) = 530 nM (EI) to Ki*相似文献   

8.
Dash C  Phadtare S  Deshpande V  Rao M 《Biochemistry》2001,40(38):11525-11532
We present here the first report of a hydrophilic peptidic inhibitor, ATBI, from an extremophilic Bacillus sp. exhibiting a two-step inhibition mechanism against the aspartic proteases, pepsin and F-prot from Aspergillus saitoi. Kinetic analysis shows that these proteases are competitively inhibited by ATBI. The progress curves are time-dependent and consistent with slow-tight binding inhibition: E + I right arrow over left arrow (k(3), k(4)) EI right arrow over left arrow (k(5), k(6)) EI. The K(i) values for the first reversible complex (EI) of ATBI with pepsin and F-prot were (17 +/- 0.5) x 10(-9) M and (3.2 +/- 0.6) x 10(-6) M, whereas the overall inhibition constant K(i) values were (55 +/- 0.5) x 10(-12) M and (5.2 +/- 0.6) x 10(-8) M, respectively. The rate constant k(5) revealed a faster isomerization of EI for F-prot [(2.3 +/- 0.4) x 10(-3) s(-1)] than pepsin [(7.7 +/- 0.3) x 10(-4) s(-1)]. However, ATBI dissociated from the tight enzyme-inhibitor complex (EI) of F-prot faster [(3.8 +/- 0.5) x 10(-5) s(-1)] than pepsin [(2.5 +/- 0.4) x 10(-6) s(-1)]. Comparative analysis of the kinetic parameters with pepstatin, the known inhibitor of pepsin, revealed a higher value of k(5)/k(6) for ATBI. The binding of the inhibitor with the aspartic proteases and the subsequent conformational changes induced were monitored by exploiting the intrinsic tryptophanyl fluorescence. The rate constants derived from the fluorescence data were in agreement with those obtained from the kinetic analysis; therefore, the induced conformational changes were correlated to the isomerization of EI to EI. Chemical modification of the Asp or Glu by WRK and Lys residues by TNBS abolished the antiproteolytic activity and revealed the involvement of two carboxyl groups and one amine group of ATBI in the enzymatic inactivation.  相似文献   

9.
Metallo-beta-lactamases have attracted considerable attention due to their role in microbial resistance to beta-lactam antibiotics. IMP-1, the binuclear Zn-dependent beta-lactamase produced by Pseudomonas aeruginosa and other microorganisms, is of particular interest in view of its increasing prevalence. An examination of the susceptibility of IMP-1 to inactivation by six different divalent metal ion chelators has revealed that all except Zincon cause inhibition by forming a complex with the holoenzyme. Exposure of the enzyme to dipicolinic acid (DPA), the most potent inhibitor, results in the production of the mononuclear Zn form of the protein as determined by electrospray ionization mass spectrometry (ESI-MS) under nondenaturing conditions. This mononuclear Zn species was found to be catalytically competent. Studies with the chromophoric chelator 4-(2-pyridylazo)resorcinol (PAR) show that the two zinc centers in IMP-1 differ in their accessibility, a feature that could be overcome in the presence of guanidine hydrochloride (GdnHCl, 1.5 M).  相似文献   

10.
1,4-benzoquinone (BQ) and 2,5-dimethyl-1,4-benzoquinone (DMBQ) were studied as inhibitors of jack bean urease in 50 mM phosphate buffer, pH 7.0. The mechanisms of inhibition were evaluated by progress curves studies and steady-state approach to data achieved by preincubation of the enzyme with the inhibitor. The obtained reaction progress curves were time-dependent and characteristic of slow-binding inhibition. The effects of different concentrations of BQ and DMBQ on the initial and steady-state velocities as well as the apparent first-order velocity constants obeyed the relationships of two-step enzyme-inhibitor interaction, qualified as mechanism B. The rapid formation of an initial BQ-urease complex with an inhibition constant of Ki = 0.031 mM was followed by a slow isomerization into the final BQ-urease complex with the overall inhibition constant of Ki* = 4.5 x 10(-5) mM. The respective inhibition constants for DMBQ were Ki = 0.42 mM, Ki* = 1.2 x 10(-3) mM. The rate constants of the inhibitor-urease isomerization indicated that forward processes were rapid in contrast to slow reverse reactions. The overall inhibition constants obtained by the steady-state analysis were found to be 5.1 x 10(-5) mM for BQ and 0.98 x 10(-3) mM for DMBQ. BQ was found to be a much stronger inhibitor of urease than DMBQ. A test, based on reaction with L-cysteine, confirmed the essential role of the sulfhydryl group in the inhibition of urease by BQ and DMBQ.  相似文献   

11.
1,4-benzoquinone (BQ) and 2,5-dimethyl-1,4-benzoquinone (DMBQ) were studied as inhibitors of jack bean urease in 50 mM phosphate buffer, pH 7.0. The mechanisms of inhibition were evaluated by progress curves studies and steady-state approach to data achieved by preincubation of the enzyme with the inhibitor. The obtained reaction progress curves were time-dependent and characteristic of slow-binding inhibition. The effects of different concentrations of BQ and DMBQ on the initial and steady-state velocities as well as the apparent first-order velocity constants obeyed the relationships of two-step enzyme-inhibitor interaction, qualified as mechanism B. The rapid formation of an initial BQ-urease complex with an inhibition constant of K i =0.031 mM was followed by a slow isomerization into the final BQ-urease complex with the overall inhibition constant of K*i=4.5 × 10 ?5 mM. The respective inhibition constants for DMBQ were K i =0.42 mM, K*i =1.2 × 10 ?3 mM. The rate constants of the inhibitor-urease isomerization indicated that forward processes were rapid in contrast to slow reverse reactions. The overall inhibition constants obtained by the steady-state analysis were found to be 5.1 × 10 ?5 mM for BQ and 0.98 × 10 ?3 mM for DMBQ. BQ was found to be a much stronger inhibitor of urease than DMBQ. A test, based on reaction with L-cysteine, confirmed the essential role of the sulfhydryl group in the inhibition of urease by BQ and DMBQ.  相似文献   

12.
Metallo-β-lactamases have attracted considerable attention due to their role in microbial resistance to β-lactam antibiotics. IMP-1, the binuclear Zn-dependent β-lactamase produced by Pseudomonas aeruginosa and other microorganisms, is of particular interest in view of its increasing prevalence. An examination of the susceptibility of IMP-1 to inactivation by six different divalent metal ion chelators has revealed that all except Zincon cause inhibition by forming a complex with the holoenzyme. Exposure of the enzyme to dipicolinic acid (DPA), the most potent inhibitor, results in the production of the mononuclear Zn form of the protein as determined by electrospray ionization mass spectrometry (ESI-MS) under nondenaturing conditions. This mononuclear Zn species was found to be catalytically competent. Studies with the chromophoric chelator 4-(2-pyridylazo)resorcinol (PAR) show that the two zinc centers in IMP-1 differ in their accessibility, a feature that could be overcome in the presence of guanidine hydrochloride (GdnHCl, 1.5 M).  相似文献   

13.
Lavendustin-A was reported to be a potent tyrosine kinase inhibitor of the epidermal growth factor (EGF) receptor (Onoda, T., Iinuma, H., Sasaki, Y., Hamada, M., Isshibi, K., Naganawa, H., Takeuchi, T., Tatsuta, K., and Umezawa, K. (1989) J. Nat. Prod. 52, 1252-1257). Its inhibition kinetics was studied in detail using the baculovirus-expressed recombinant intracellular domain of the EGF receptor (EGFR-IC). Lavendustin-A (RG 14355) is a slow and tight binding inhibitor of the receptor tyrosine kinase. The pre-steady state kinetic analysis demonstrates that the inhibition corresponds to a two-step mechanism in which an initial enzyme-inhibitor complex (EI) is rapidly formed followed by a slow isomerization step to form a tight complex (EI*). The dissociation constant for the initial rapid forming complex is 370 nM, whereas the overall dissociation constant is estimated to be less than or equal to 1 nM. The difference between the two values is due to the tight binding nature of the inhibitor to the enzyme in EI*. The kinetic analysis using a preincubation protocol to pre-equilibrate the enzyme with the inhibitor in the presence of one substrate showed that Lavendustin-A is a hyperbolic mixed-type inhibitor with respect to both ATP and the peptide substrate, with a major effect on the binding affinities for both substrates. An analogue of Lavendustin-A (RG 14467) showed similar inhibition kinetics to that of Lavendustin-A. The results of the pre-steady state analysis are also consistent with the proposed two-step mechanism. The dissociation constant for the initial fast forming complex in this case is 3.4 microM, whereas the overall dissociation constant is estimated to be less than or equal to 30 nM. It is a partial (hyperbolic) competitive inhibitor with respect to ATP. Its inhibition is reduced to different extents by different peptide substrates, when the peptide is added to the enzyme simultaneously with the inhibitor. When studied with the least protective peptide, K1 (a peptide containing the major autophosphorylation site of the EGF receptor), RG 14467 acts as a hyperbolic noncompetitive inhibitor with respect to the peptide.  相似文献   

14.
In a preceding study, 4,4'-dithiodipyridine (DTDP) was shown to be superior to 5,5'-dithio-bis(2-nitrobenzoic acid) (Ellman's reagent) for spectrophotometric measurement of thiol groups in aqueous solution. (i) Sensitivity is higher because a larger absorbance increase is seen at a given thiol concentration. (ii) The intrinsic reactivity of thiolate anions for DTDP is much higher than for Ellman's reagent; thus, the reaction can be carried out at pH > or = 4.5 instead of at pH 8.0. In the present study, these advantages of DTDP were exploited for spectrophotometric measurement of thiols in organic solvent. DTDP was found to quantitatively react with nonpolar thiols when triethylamine was used as catalyst, intense light absorption (between 344 and 360 nm) was seen when the reaction was terminated with acetic acid, and the spectrophotometric responses were independent of the nonthiol portions of the mercaptans. The determination limit (10x the standard deviation of the reagent blank) was 3 microM, and the upper limit was approximately 40 microM on a typical spectrophotometer. The thiol contents of the mercaptans were independently verified by a modification of standard iodometry in which toluene/butanol or chloroform/butanol was included to dissolve nonpolar mercaptans.  相似文献   

15.
Bovine kidney aldose reductase (ALR2) displays substrate inhibition by aldehyde substrates that is uncompetitive versus NADPH when allowance is made for nonenzymic reaction of the aldehyde with the adenine moiety of NADPH. A time-dependent increase in substrate inhibition observed in product versus time plots for reduction of short-chain aldoses containing an enolizable alpha-proton, but not for p-nitrobenzaldehyde, is shown to be consistent with a model in which rapidly reversible inhibition due to formation of the dead-end E-NADP-glycolaldehyde complex is combined with the formation at the enzyme active site of a tightly-bound covalent NADP-glycolaldehyde adduct. Quantitative analysis of reaction time courses for ALR2-catalyzed reduction of glycolaldehyde using NADPH or the 3-acetylpyridine analogue, (AP)ADPH, yields values of the forward and reverse rate constants for ALR2-mediated adduct formation that agree with the values determined in the absence of glycolaldehyde turnover. Substrate inhibition is only partial, indicating that reaction can occur via an alternate pathway at high [glycolaldehyde]. Kinetic evidence for a slow isomerization of the E-NADP complex at pH 8.0 is used to explain the similar V/Et values observed for glycolaldehyde reduction at pH 7.0 using NADPH, (AP)ADPH, and the hypoxanthine analogue N(Hx)DPH. The practical implications of these results for kinetics studies of aldose reductase are discussed.  相似文献   

16.
The interaction of tetracycline and oxovanadium(IV) in aqueous solution was studied by potentiometric and spectrophotometric methods. Oxovanadium(IV) ions form both a positively charged 1:1 and a neutral 2:1 metal-ligand complex with tetracycline. When a 1:1 ligand-to-metal ratio mixture is used at about pH 4.5 the 1:1 species predominates, being replaced at pH 6 by the binuclear complex. The binuclear complex has been isolated and fully characterised. Infrared and EPR studies suggest the existence of two distinct vanadyl binding sites. Our results indicate that the first vanadium coordinates to the BCD-ring system and the second one to the A-ring. Biological implications of the existence of a neutral complex at physiological pH are briefly discussed.  相似文献   

17.
R L Blakley  L Cocco 《Biochemistry》1985,24(18):4772-4777
Stopped-flow measurements of protein fluorescence quenching when methotrexate (MTX) binds to dihydrofolate reductase (isoenzyme II) of Streptococcus faecium (SFDHFR II) analyze as the sum of two differentials: a rapid binding phase and a second phase for which the observed rate constant is independent of methotrexate concentration. Analysis of variation of the ratio of the amplitude of the fast and slow phases with methotrexate concentration indicates that the second phase is an isomerization of the initial binary complex. At pH 7.3, the equilibrium constant for this isomerization is 21.9, and the forward and reverse rate constants are 0.57 and 0.026 s-1, respectively. Similar results were obtained for binding of 3-deazamethotrexate to SFDHFR II, but the forward rate constant is greater (2.9 s-1 at pH 7.3). The equilibrium constants for these isomerizations are pH independent, but the rate constants decrease as the pH is raised, probably due to deprotonation of one or more groups on the enzyme. Analysis of progress curves obtained by the development of inhibition when SFDHFR II is added last to reaction mixtures containing dihydrofolate, NADPH, and MTX gives an association constant for initial reactions of 4.3 X 10(7) M-1. Since a preliminary estimate of the association constant for the binding reaction is 7.6 X 10(5) M-1, this suggests an isomerization of the ternary complex(es) with an equilibrium constant of about 56. In addition, analysis of the progress of development of inhibition indicates a further very slow isomerization with equilibrium constant 419 and forward rate constant 2.6 min-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
pH dependence of folding of iso-2-cytochrome c   总被引:4,自引:0,他引:4  
B T Nall  J J Osterhout  L Ramdas 《Biochemistry》1988,27(19):7310-7314
Starting from a standard unfolded state (3.0 M guanidine hydrochloride, pH 7.2), the kinetics of refolding of iso-2-cytochrome c have been investigated as a function of final pH between pH 3 and pH 10. Absorbance in the ultraviolet and visible spectral regions and tryptophan fluorescence are used to monitor folding. Over most of the pH range, fast and slow folding phases are detected by both fluorescence and absorbance probes. Near neutral pH, the rate of fast folding appears to be the same when monitored by absorbance and fluorescence probes. At higher and lower pH, there are two fast folding reactions, with absorbance-detected fast folding occurring in a slightly faster time range than fluorescence-detected fast folding. The rates of both fast folding reactions pass through broad minima near neutral pH, indicating involvement of ionizable groups in rate-limiting steps. The rates of slow folding also depend on the final pH. At acid pH, there appears to be a single slow folding phase for both fluorescence and absorbance probes. At neutral pH, the absorbance-detected and fluorescence-detected slow folding phases separate into distinct kinetic processes which differ in rate and relative amplitude. At high pH, absorbance-detected slow folding is no longer observed, while fluorescence-detected slow folding is decreased in amplitude. In contrast, the equilibrium and kinetic properties of proline imide bond isomerization, believed to be involved in the slow folding reactions, are largely independent of pH. The results suggest that the pH dependence of slow folding involves coupling of pH-sensitive structure to proline imide bond isomerization.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The resonance Raman spectra of the aa3 cytochrome c oxidase from Rhodobacter sphaeroides reveal pH-dependent structural changes in the binuclear site at room temperature. The binuclear site, which is the catalytic center of the enzyme, possesses two conformations at neutral pH, assessed from their distinctly different Fe-CO stretching modes in the resonance Raman spectra of the CO complex of the fully reduced enzyme. The two conformations (alpha and beta) interconvert reversibly in the pH 6-9 range with a pKa of 7.4, consistent with Fourier transform infrared spectroscopy measurements done at cryogenic temperatures (D.M. Mitchell, J.P. Sapleigh, A.M.Archer, J.O. Alben, and R.B.Gennis, 1996, Biochemistry 35:9446-9450). It is postulated that the different structures result from a change in the position of the Cu(B) atom with respect to the CO due to the presence of one or more ionizable groups in the vicinity of the binuclear center. The conserved tyrosine residue (Tyr-288 in R. sphaeroides, Tyr-244 in the bovine enzyme) that is adjacent to the oxygen-binding pocket or one of the histidines that coordinate Cu(B) are possible candidates. The existence of an equilibrium between the two conformers at physiological pH and room temperature suggests that the conformers may be functionally involved in enzymatic activity.  相似文献   

20.
Unfolded ribonuclease A consists of 20% fast refolding (Uf) and 80% slow refolding material (Us). The latter consists of at least two different forms which refold at different rates. We have used absorbance and fluorescence spectrophotometry to compare the kinetics of refolding in aqueous and aqueous-methanol solutions. At 1 degree C and pH 3.0, the addition of increasing concentrations of methanol (to 50%, v/v) had negligible effect on the rates and amplitudes of the slow refolding Us states. The effect of temperature on the two slow phases of refolding was determined in 35 and 50% methanol. From Arrhenius plots the energies of activation were found to be in the vicinity of 20 kcal/mol for both processes. The results suggest that both slow phases correspond to proline isomerization, and that the presence of methanol does not significantly perturb the overall refolding process. It is possible that the faster of the slow refolding phases corresponds to the isomerization of a proline residue which is trans in the folded native state but which undergoes extensive isomerization to the cis conformation in the unfolded state.  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