首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A remarkable correlation has been discovered between fluorescence lifetimes of bound NADPH and rates of hydride transfer among mutants of dihydrofolate reductase (DHFR) from Escherichia coli. Rates of hydride transfer from NADPH to dihydrofolate change by a factor of 1,000 for the series of mutant enzymes. Since binding constants for the initial complex between coenzyme and DHFR change by only a factor of 10, the major portion of the change in hydride transfer must be attributed to losses in transition-state stabilization. The time course of fluorescence decay for NADPH bound to DHFR is biphasic. Lifetimes ranging from 0.3 to 0.5 ns are attributed to a solvent-exposed dihydronicotinamide conformation of bound coenzyme which is presumably not active in catalysis, while decay times (tau 2) in the range of 1.3 to 2.3 ns are assigned to a more tightly bound species of NADPH in which dihydronicotinamide is sequestered from solvent. It is this slower component that is of interest. Ternary complexes with three different inhibitors, methotrexate, 5-deazafolate, and trimethoprim, were investigated, along with the holoenzyme complex; 3-acetylNADPH was also investigated. Fluorescence polarization decay, excitation polarization spectra, the temperature variation of fluorescence lifetimes, fluorescence amplitudes, and wavelength of absorbance maxima were measured. We suggest that dynamic quenching or internal conversion promotes decay of the excited state in NADPH-DHFR. When rates of hydride transfer are plotted against the fluorescence lifetime (tau 2) of tightly bound NADPH, an unusual correlation is observed. The fluorescence lifetime becomes longer as the rate of catalysis decreases for most mutants studied. However, the fluorescence lifetime is unchanged for those mutations that principally alter the binding of dihydrofolate while leaving most dihydronicotinamide interactions relatively undisturbed. The data are interpreted in terms of possible dynamic motions of a flexible loop region in DHFR which closes over both substrate and coenzyme binding sites. These motions could lead to faster rates of fluorescence decay in holoenzyme complexes and, when correlated over time, may be involved in other motions which give rise to enhanced rates of catalysis in DHFR.  相似文献   

2.
This paper reviews the results from hybrid quantum/classical molecular dynamics simulations of the hydride transfer reaction catalysed by wild-type (WT) and mutant Escherichia coli and WT Bacillus subtilis dihydrofolate reductase (DHFR). Nuclear quantum effects such as zero point energy and hydrogen tunnelling are significant in these reactions and substantially decrease the free energy barrier. The donor-acceptor distance decreases to ca 2.7 A at transition-state configurations to enable the hydride transfer. A network of coupled motions representing conformational changes along the collective reaction coordinate facilitates the hydride transfer reaction by decreasing the donor-acceptor distance and providing a favourable geometric and electrostatic environment. Recent single-molecule experiments confirm that at least some of these thermally averaged equilibrium conformational changes occur on the millisecond time-scale of the hydride transfer. Distal mutations can lead to non-local structural changes and significantly impact the probability of sampling configurations conducive to the hydride transfer, thereby altering the free-energy barrier and the rate of hydride transfer. E. coli and B. subtilis DHFR enzymes, which have similar tertiary structures and hydride transfer rates with 44% sequence identity, exhibit both similarities and differences in the equilibrium motions and conformational changes correlated to hydride transfer, suggesting a balance of conservation and flexibility across species.  相似文献   

3.
The variable residue Leu-28 of Escherichia coli dihydrofolate reductase (DHFR) and the corresponding residue Phe-31 in murine DHFR were interchanged, and the impact on catalysis was evaluated by steady-state and pre-steady-state analysis. The E. coli L28F mutant increased the pH-independent kcat from 11 to 50 s-1 but had little effect on Km(H2F). An increase in the rate constant for dissociation of H4F from E.H4F.NH (from 12 to 80 s-1) was found to be largely responsible for the increase in kcat. Unexpectedly, the rate constant for hydride transfer increased from 950 to 4000 s-1 with little perturbation of NADPH and NADP+ binding to E. Consequently, the flux efficiency of the E. coli L28F mutant rose from 15% to 48% and suggests a role in genetic selection for this variable side chain. The murine F31L mutant decreased the pH-independent kcat from 28 to 4.8 s-1 but had little effect on Km(H2F). A decrease in the rate constant for dissociation of H4F from E.H4F.NH (from 40 to 22 s-1) and E.H4F (from 15 to 0.4 s-1) was found to be mainly responsible for the decrease in kcat. The rate constant for hydride transfer decreased from 9000 to 5000 s-1 with minor perturbation of NADPH binding. Thus, the free energy differences along the kinetic pathway were generally similar in magnitude but opposite in direction to those incurred by the E. coli L28F mutant. This conclusion implies that DHFR hydrophobic active-site side chains impart their characteristics individually and not collectively.  相似文献   

4.
Kim HS  Damo SM  Lee SY  Wemmer D  Klinman JP 《Biochemistry》2005,44(34):11428-11439
Dihydrofolate reductase (DHFR) from a moderate thermophilic organism, Bacillus stearothermophilus, has been cloned and expressed. Physical characterization of the protein (BsDHFR) indicates that it is a monomeric protein with a molecular mass of 18,694.6 Da (0.8), coincident with the mass of 18 694.67 Da calculated from the primary sequence. Determination of the X-ray structure of BsDHFR provides the first structure for a monomeric DHFR from a thermophilic organism, indicating a high degree of conservation of structure in relation to all chromosomal DHFRs. Structurally based sequence alignment of DHFRs indicates the following levels of sequence identity and similarity for BsDHFR: 38 and 58% with Escherichia coli, 35 and 56% with Lactobacillus casei, and 23 and 40% with Thermotoga maritima, respectively. Steady state kinetic isotope effect studies indicate an ordered kinetic mechanism at elevated temperatures, with NADPH binding first to the enzyme. This converts to a more random mechanism at reduced temperatures, reflected in a greatly reduced K(m) for dihydrofolate at 20 degrees C in relation to that at 60 degrees C. A reduction in either temperature or pH reduces the degree to which the hydride transfer step is rate-determining for the second-order reaction of DHF with the enzyme-NADPH binary complex. Transient state kinetics have been used to study the temperature dependence of the isotope effect on hydride transfer at pH 9 between 10 and 50 degrees C. The data support rate-limiting hydride transfer with a moderate enthalpy of activation (E(a) = 5.5 kcal/mol) and a somewhat greater temperature dependence for the kinetic isotope effect than predicted from classical behavior [A(H)/A(D) = 0.57 (0.15)]. Comparison of kinetic parameters for BsDHFR to published data for DHFR from E. coli and T. maritima shows a decreasing trend in efficiency of hydride transfer with increasing thermophilicity of the protein. These results are discussed in the context of the capacity of each enzyme to optimize H-tunneling from donor (NADPH) to acceptor (DHF) substrates.  相似文献   

5.
The structure and folding of dihydrofolate reductase (DHFR) from Escherichia coli and the mutant G121V-DHFR, in which glycine 121 in the exterior FG loop was replaced with valine, were studied by molecular dynamics simulations and CD and fluorescence spectroscopy. The importance of residue 121 for the chemical step during DHFR catalysis had been demonstrated previously. High-temperature MD simulations indicated that while DHFR and G121V-DHFR followed similar unfolding pathways, the strong contacts between the M20 loop and the FG loop in DHFR were less stable in the mutant. These contacts have been proposed to be involved in a coupled network of interactions that influence the protein dynamics and promote catalysis [Benkovic, S. J., and Hammes-Schiffer, S. (2003) Science 301, 1196-1202]. CD spectroscopy of DHFR and G121V-DHFR indicated that the two proteins existed in different conformations at room temperature. While the thermally induced unfolding of DHFR was highly cooperative with a midpoint at 51.6 +/- 0.7 degrees C, G121V-DHFR exhibited a gradual decrease in its level of secondary structure without a clear melting temperature. Temperature-induced unfolding and renaturation from the urea-denatured state revealed that both proteins folded via highly fluorescent intermediates. The formation of these intermediates occurred with relaxation times of 149 +/- 4.5 and 256 +/- 13 ms for DHFR and G121V-DHFR, respectively. The fluorescence intensity for the intermediates formed during refolding of G121V-DHFR was approximately twice that of the wild-type. While the fluorescence intensity then slowly decayed for DHFR toward a state representing the native protein, G121V-DHFR appeared to be trapped in a highly fluorescent state. These results suggest that the reduced catalytic activity of G121V-DHFR is the consequence of nonlocal structural effects that may result in a perturbation of the network of promoting motions.  相似文献   

6.
The complete nucleotide and encoded amino acid sequences were determined for the dihydrofolate reductase (DHFR) from the bacteria Enterobacter aerogenes and Citrobacter freundii. These were compared with the closely related Escherichia coli DHFR sequence. The ancestral DHFR sequence common to these three species was reconstructed. Since that ancestor there have been seven, nine, and one amino acid replacements in E. coli, E. aerogenes, and C. freundii, respectively. In E. coli, five of its seven replacements were located in the beta-sheet portion of the protein, and all seven were located in a single restricted region of the protein. In E. aerogenes, all nine of its replacements were located within surface residues, with five clustered in a region topologically distinct from the E. coli cluster. The replaced side chains are sometimes in direct contact but more often are separated by an intervening side chain. It is argued that the temporal clustering of replacements is typical for the evolution of most proteins and that the associated topological clustering gives a picture of how evolutionary change is accommodated by protein structure.  相似文献   

7.
Protein-mediated transfer of phosphatidylcholine (PC) by bovine liver phosphatidylcholine transfer protein (PC-TP) was examined using a vesicle-vesicle assay system. Donor and acceptor membranes were prepared from Escherichia coli phospholipids and limiting amounts of egg yolk PC. PC transfer between vesicles of E. coli lipid/egg PC was markedly higher than transfer of PC from vesicles of E. coli lipid/egg PC to vesicles of E. coli lipid. Kinetic parameters of the interaction between PC-TP and E. coli lipid vesicles with or without PC was investigated. The apparent dissociation constants of the complex formed between PC-TP and these vesicles were determined kinetically and from double-reciprocal plots of intrinsic PC-TP fluorescence intensity increase versus vesicle concentration. The magnitude of the dissociation constant decreased as the PC content of the vesicles increased from 0 to 5 mol%. In addition, kinetic analysis revealed that the presence of PC in acceptor vesicles increased both the association and dissociation of PC-TP from vesicles. The effect of membrane PC molecules on transfer rates was examined using bis-phosphatidylcholine, a dimeric PC molecule which is not transferred by PC-TP. Rates of PC transfer to acceptor vesicles comprised of E. coli lipid/bis-PC were virtually identical to rates observed with acceptors vesicles prepared from E. coli lipid. The results suggest that transfer of PC by PC-TP is enhanced only when insertion of protein-bound PC occurs concurrently with the extraction of a molecule of membrane PC, i.e., a concerted, one-step catalytic mechanism for phospholipid exchange.  相似文献   

8.
Ensemble kinetics and single-molecule fluorescence microscopy were used to study conformational transitions associated with enzyme catalysis by dihydrofolate reductase (DHFR). The active site loop of DHFR was labeled with a fluorescence quencher, QSY35, at amino acid position 17, and the fluorescent probe, Alexa555, at amino acid 37, by introducing cysteines at these sites with site-specific mutagenesis. The distance between the probes was such that approximately 50% fluorescence resonance energy transfer (FRET) occurred. The double-labeled enzyme retained essentially full catalytic activity, and stopped-flow studies of both the forward and reverse reactions revealed that the distance between probes increased prior to hydride transfer. A fluctuation in fluorescence intensity of single molecules of DHFR was observed in an equilibrium mixture of substrates but not in their absence. Ensemble rate constants were derived from the distributions of lifetimes observed and attributed to a reversible conformational change. Studies were carried out with both NADPH and NADPD as substrates, with no measurable isotope effect. Similar studies with a G121V mutant DHFR resulted in smaller rate constants. This mutant DHFR has reduced catalytic activity, so that the collective data for the conformational change suggest that the conformational change being observed is associated with catalysis and probably represents a conformational change prior to hydride transfer. If the change in fluorescence is attributed to a change in FRET, the distance change associated with the conformational change is approximately 1-2 A. These results are correlated with other measurements related to conformation coupled catalysis.  相似文献   

9.
Riboflavin is biosynthesized by most microorganisms and plants, while mammals depend entirely on the absorption of this vitamin from the diet to meet their metabolic needs. Therefore, riboflavin biosynthesis appears to be an attractive target for drug design, since appropriate inhibitors of the pathway would selectively target the microorganism. We have cloned and solubly expressed the bifunctional ribD gene from Escherichia coli, whose three-dimensional structure was recently determined. We have demonstrated that the rate of deamination (370 min (-1)) exceeds the rate of reduction (19 min (-1)), suggesting no channeling between the two active sites. The reductive ring opening reaction occurs via a hydride transfer from the C 4- pro-R hydrogen of NADPH to C'-1 of ribose and is the rate-limiting step in the overall reaction, exhibiting a primary kinetic isotope effect ( (D) V) of 2.2. We also show that the INH-NADP adduct, one of the active forms of the anti-TB drug isoniazid, inhibits the E. coli RibD. On the basis of the observed patterns of inhibition versus the two substrates, we propose that the RibD-catalyzed reduction step follows a kinetic scheme similar to that of its structural homologue, DHFR.  相似文献   

10.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  相似文献   

11.
Thorpe IF  Brooks CL 《Proteins》2004,57(3):444-457
The energy barrier for hydride transfer in wild-type G121V and G121S variants of Escherichia coli dihydrofolate reductase (DHFR) fluctuates in a time-dependent manner. This fluctuation may be attributed to structural changes in the protein that modulate the site of chemistry. Despite being far from the active site, mutations at position 121 of DHFR reduce the hydride transfer rate of the enzyme. This occurrence has been suggested to arise from modifications to the conformational ensemble of the protein. We elucidate the effects of the G121S and G121V mutations on the hydride transfer barrier by identifying structural changes in the protein that correlate with lowered barriers. The effect of these structural parameters on the hydride transfer barrier may be rationalized by simple considerations of the geometric constraints of the hydride transfer reaction. Fluctuations of these properties are associated with specific backbone dihedral angles of residues within the Methione-20 (M20) loop. The dihedral angle preferences are mediated by interactions with the region of the enzyme in the vicinity of residue 121 and are translated into distinct ligand conformations. We predict mutations within the M20 loop that may alter the conformational space explored by DHFR. Such mutational changes are anticipated to adjust the hydride transfer efficacy of DHFR by modifying equilibrium distributions of hydride transfer barriers found in the enzyme.  相似文献   

12.
The unfolded state of a protein is an ensemble of a large number of conformations ranging from fully extended to compact structures. To investigate the effects of the difference in the unfolded-state ensemble on protein folding, we have studied the structure, stability, and folding of "circular" dihydrofolate reductase (DHFR) from Escherichia coli in which the N and C-terminal regions are cross-linked by a disulfide bond, and compared the results with those of disulfide-reduced "linear" DHFR. Equilibrium studies by circular dichroism, difference absorption spectra, solution X-ray scattering, and size-exclusion chromatography show that whereas the native structures of both proteins are essentially the same, the unfolded state of circular DHFR adopts more compact conformations than the unfolded state of the linear form, even with the absence of secondary structure. Circular DHFR is more stable than linear DHFR, which may be due to the decrease in the conformational entropy of the unfolded state as a result of circularization. Kinetic refolding measurements by stopped-flow circular dichroism and fluorescence show that under the native conditions both proteins accumulate a burst-phase intermediate having the same structures and both fold by the same complex folding mechanism with the same folding rates. Thus, the effects of the difference in the unfolded state of circular and linear DHFRs on the refolding reaction are not observed after the formation of the intermediate. This suggests that for the proteins with close termini in the native structure, early compaction of a protein molecule to form a specific folding intermediate with the N and C-terminal regions in close proximity is a crucial event in folding. If there is an enhancement in the folding reflecting the reduction in the breadth of the unfolded-state ensemble for circular DHFR, this acceleration must occur in the sub-millisecond time-range.  相似文献   

13.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

14.
The Bacillus subtilis dihydrofolate reductase (DHFR) gene was expressed in Escherichia coli. The gene product was purified to homogeneity by Butyl-Toyopearl, Toyopearl HW55, and DEAE-Toyopearl column chromatographies, and its molecular properties were compared to those of E. coli DHFR. The specific enzyme activity of the B. subtilis DHFR was 240 units/mg under the standard assay conditions, being about four times higher than that of the E. coli DHFR. Km for coenzyme NADPH was 20.7 microM, a value about three times larger than that of E. coli, whereas Km (1.5 microM) for the substrate, dihydrofolate, was similar to that of E. coli DHFR. This seems to reflect the low homology of the amino acid sequence in residues 61-88 of the two DHFRs where one of the NADPH binding sites is located [Bystrof, C. & Kraut, J. (1991) Biochemistry 30, 2227-2239]. Similar to the E. coli DHFR [Iwakura, M. et al. (1992) J. Biochem. 111, 37-45], the extension of amino acid sequences at the C-terminal end of the B. subtilis DHFR could be attained without loss of the enzyme function or decrease of the protein yield. Thus, the DHFR is useful as a carrier protein for expressing small polypeptides, such as leucine enkephalin, bradykinin, and somatostatin.  相似文献   

15.
Achieving molecular recognition of NADPH binding sites is a compelling strategy to control many redox biological processes. The NADPH sites recognize the ubiquitous NADPH cofactor via highly conserved binding interactions, despite differences in the regulation of the hydride transfer in redox active proteins. We recently developed a photoactive NADPH substitute, called nanotrigger NT synchronizing the initiation of enzymatic catalysis of the endothelial NO‐synthase (eNOS) with a laser pulse. Spatial and temporal control of enzymatic activity by such a designed light‐driven activator would benefit from achieving molecular selectivity, i.e. activation of a single NADPH‐mediated enzyme. In this work, we probe the ability of NT to discriminate between two NADPH sites with light. The selected NADPH sites belong to dihydrofolate reductase dihydrofolate reductase enzyme (DHFR) and endothelial NO‐synthase (eNOS). Ultrafast kinetics showed that NT could not activate DHFR catalysis with a laser pulse in contrast with the observed trigger of eNOS catalysis leading to NO formation. Homology modelling, molecular dynamics simulations showed that NT discriminated between the two NADPH sites by different donor to acceptor distances and by local steric effects hindering light activation of DHFR catalysis. The data suggested that the narrow NADPH site required a tight fit of the nanotrigger at a suitable distance/angle to the electron acceptor for a specific activation of the catalysis. The ability of the nanotrigger to activate eNOS combined with a low reactivity in unfavourable NADPH sites makes NT a highly promising tool for targeting eNOS in endothelial cells with a laser pulse. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The technique of hydrogen-deuterium exchange coupled to mass spectrometry (HDX-MS) has been applied to a mesophilic (E. coli) dihydrofolate reductase under conditions that allow direct comparison to a thermophilic (B. stearothermophilus) ortholog, Ec-DHFR and Bs-DHFR, respectively. The analysis of hydrogen-deuterium exchange patterns within proteolytically derived peptides allows spatial resolution, while requiring a series of controls to compare orthologous proteins with only ca. 40% sequence identity. These controls include the determination of primary structure effects on intrinsic rate constants for HDX as well as the use of existing 3-dimensional structures to evaluate the distance of each backbone amide hydrogen to the protein surface. Only a single peptide from the Ec-DHFR is found to be substantially more flexible than the Bs-DHFR at 25 °C in a region located within the protein interior at the intersection of the cofactor and substrate-binding sites. The surrounding regions of the enzyme are either unchanged or more flexible in the thermophilic DHFR from B. stearothermophilus. The region with increased flexibility in Ec-DHFR corresponds to one of two regions previously proposed to control the enthalpic barrier for hydride transfer in Bs-DHFR [Oyeyemi et al. (2010) Proc. Natl. Acad. Sci. U.S.A. 107, 10074].  相似文献   

17.
As a continuation to our studies on the importance of interloop interactions in the Escherichia coli DHFR catalytic cycle, we have investigated the role of the betaG-betaH loop in modulating the closed and occluded conformations of the Met20 loop during the DHFR catalytic cycle. Specifically, to assess the importance of the hydrogen bond formed between Ser148 in the betaG-betaH loop and the Met20 loop, Ser148 was independently substituted with aspartic acid, alanine, and lysine. Moreover, the betaG-betaH loop was deleted entirely to yield the Delta(146-148) DHFR mutant. Steady-state turnover rates for all mutants were at most 3-fold lower than the wild-type rate. Lack of an isotope effect on this rate indicated the chemistry step does not contribute to the steady-state turnover. Consistent with this finding, hydride transfer rates for the DHFR mutants were at least 10-fold greater than the observed steady-state rates. The values ranged from a 30% decrease (Ser148Ala and Ser148Lys) to a 50% increase (Ser148Asp) in rate relative to that of the wild type. Modifications of the betaG-betaH loop enhanced the affinity for the cofactor and decreased the affinity for pterin, as determined by the K(D) values of the mutant proteins. Further analysis of Ser148Ala and Delta(146-148) DHFRs indicated these effects were manifest mainly in ligand off rates, although in some cases the on rate was affected. The Ser148Asp and Delta(146-148) mutations perturbed the preferred catalytic cycle through the introduction of branching at key intermediates. Rather than following the single WT pathway which involves loss of NADP(+) and rebinding of NADPH to precede loss of the product H4F (negative cooperativity), the mutants can reenter the catalytic cycle through different pathways. These findings suggest that the role of the interloop interaction between the betaG-betaH loop and the Met20 loop is to modulate ligand off rates allowing for proper cycling through the preferred kinetic pathway.  相似文献   

18.
Cyclophilins accelerate slow protein folding reactions in vitro by catalyzing the cis/trans isomerization of peptidyl-prolyl bonds. Cyclophilins were reported to be involved in a variety of cellular functions, including the promotion of protein folding by use of the substrate mouse dihydrofolate reductase (DHFR). The interaction of cyclophilin with DHFR has only been studied under limited conditions so far, not taking into account that native DHFR exists in equilibrium with a non-native late-folding intermediate. Here we report a systematic analysis of catalysis of DHFR folding by cyclophilins. The specific ligand methotrexate traps DHFR in its native state, permitting a specific analysis of the action of cyclophilin on both denatured DHFR with non-native prolyl bonds and denatured DHFR with all-native prolyl bonds. Cyclophilins from yeast and Neurospora crassa as well as the related prolyl isomerase b from Escherichia coli promote the folding of different forms of DHFR to the enzymatically active form, demonstrating the generality of cyclophilin-catalyzed folding of DHFR. The slow equilibrium between the late-folding intermediate and native DHFR suggests that prolyl isomerization may be required for this final phase of conversion to native DHFR. However, by reversible trapping of the intermediate, we analyze the slow interconversion between native and late-folding conformations in the backward and forward reactions and show a complete independence of cyclophilin. We conclude that cyclophilin catalyzes folding of DHFR, but surprisingly not in the last slow folding step.  相似文献   

19.
Alkaline phosphatase (AP) from a North Atlantic marine Vibrio bacterium was previously characterized as being kinetically cold-adapted. It is still unknown whether its characteristics originate locally in the active site or are linked to more general structural factors. There are three metal-binding sites in the active site of APs, and all three metal ions participate in catalysis. The amino acid residues that bind the two zinc ions most commonly present are conserved in all known APs. In contrast, two of the residues that bind the third metal ion (numbered 153 and 328 in Escherichia coli AP) are different in various APs. This may explain their different catalytic efficiencies, as the Mg2+ most often present there is important for both structural stability and the reaction mechanism. We have mutated these key residues to the corresponding residues in E. coli AP to obtain the double mutant Asp116/Lys274, and both single mutants. All these mutants displayed reduced substrate affinity and lower overall reaction rates. The Lys274 and Asp116/Lys274 mutants also displayed an increase in global heat stability, which may be due to the formation of a stabilizing salt bridge. Overall, the results show that a single amino acid substitution in the active site is sufficient to alter the structural stability of the cold-active Vibrio AP both locally and globally, and this influences kinetic properties.  相似文献   

20.
The E. coli chaperonin GroEL and its cofactor GroES promote protein folding by sequestering nonnative polypeptides in a cage-like structure. Here we define the contribution of this system to protein folding across the entire E. coli proteome. Approximately 250 different proteins interact with GroEL, but most of these can utilize either GroEL or the upstream chaperones trigger factor (TF) and DnaK for folding. Obligate GroEL-dependence is limited to only approximately 85 substrates, including 13 essential proteins, and occupying more than 75% of GroEL capacity. These proteins appear to populate kinetically trapped intermediates during folding; they are stabilized by TF/DnaK against aggregation but reach native state only upon transfer to GroEL/GroES. Interestingly, substantially enriched among the GroEL substrates are proteins with (betaalpha)8 TIM-barrel domains. We suggest that the chaperonin system may have facilitated the evolution of this fold into a versatile platform for the implementation of numerous enzymatic functions.  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