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 15 毫秒
1.
The activity of enzyme I (EI), the first protein in the bacterial PEP:sugar phosphotransferase system, is regulated by a monomer-dimer equilibrium where a Mg(2+)-dependent autophosphorylation by PEP requires the homodimer. Using inactive EI(H189A), in which alanine is substituted for the active-site His189, substrate-binding effects can be separated from those of phosphorylation. Whereas 1 mM PEP (with 2 mM Mg(2+)) strongly promotes dimerization of EI(H189A) at pH 7.5 and 20 degrees C, 5 mM pyruvate (with 2 mM Mg(2+)) has the opposite effect. A correlation between the coupling of N- and C-terminal domain unfolding, measured by differential scanning calorimetry, and the dimerization constant for EI, determined by sedimentation equilibrium, is observed. That is, when the coupling between N- and C-terminal domain unfolding produced by 0.2 or 1.0 mM PEP and 2 mM Mg(2+) is inhibited by 5 mM pyruvate, the dimerization constant for EI(H189A) decreases from > 10(8) to < 5 x 10(5) or 3 x 10(7) M(-1), respectively. Incubation of the wild-type, dephospho-enzyme I with the transition-state analog phosphonopyruvate and 2 mM Mg(2+) also increases domain coupling and the dimerization constant approximately 42-fold. With 2 mM Mg(2+) at 15-25 degrees C and pH 7.5, PEP has been found to bind to one site/monomer of EI(H189A) with K(A)' approximately 10(6) M(-1) (deltaG' = -8.05 +/- 0.05 kcal/mole and deltaH = +3.9 kcal/mole at 20 degrees C); deltaC(p) = -0.33 kcal K(-1) mole(-1). The binding of PEP to EI(H189A) is synergistic with that of Mg(2+). Thus, physiological concentrations of PEP and Mg(2+) increase, whereas pyruvate and Mg(2+) decrease the amount of dimeric, active, dephospho-enzyme I.  相似文献   

2.
Enzyme I of the bacterial phosphoenolpyruvate:sugar phosphotransferase system can be phosphorylated by PEP on an active-site histidine residue, localized to a cleft between an alpha-helical domain and an alpha/beta domain on the amino terminal half of the protein. The phosphoryl group on the active-site histidine can be passed to an active-site histidine residue of HPr. It has been proposed that the major interaction between enzyme I and HPr occurs via the alpha-helical domain of enzyme I. The isolated recombinant alpha-helical domain (residues 25-145) with approximately 80% alpha-helices as well as enzyme I deficient in that domain [EI(DeltaHD)] with approximately 50% alpha-helix content from M. capricolum were used to further elucidate the nature of the enzyme I-HPr complex. Isothermal titration calorimetry demonstrated that HPr binds to the alpha-helical domain and intact enzyme I with = 5 x 10(4) and 1.4 x 10(5) M(-)(1) at pH 7.5 and 25 degrees C, respectively, but not to EI(DeltaHD), which contains the active-site histidine of enzyme I and can be autophosphorylated by PEP. In vitro reconstitution experiments with proteins from both M. capricolum and E. coli showed that EI(DeltaHD) can donate its bound phosphoryl group to HPr in the presence of the isolated alpha-helical domain. Furthermore, M. capricolum recombinant C-terminal domain of enzyme I (EIC) was shown to reconstitute phosphotransfer activity with recombinant N-terminal domain (EIN) approximately 5% as efficiently as the HD-EI(DeltaHD) pair. Recombinant EIC strongly self-associates ( approximately 10(10) M(-)(1)) in comparison to dimerization constants of 10(5)-10(7) M(-)(1) measured for EI and EI(DeltaHD).  相似文献   

3.
The bacterial PEP:sugar phosphotransferase system couples the phosphorylation and translocation of specific sugars across the membrane. The activity of the first protein in this pathway, enzyme I (EI), is regulated by a monomer-dimer equilibrium where a Mg(2+)-dependent autophosphorylation by PEP requires the dimer. Dimerization constants for dephospho- and phospho-EI and inactive mutants EI(H189E) and EI(H189A) (in which Glu or Ala is substituted for the active site His189) have been measured under a variety of conditions by sedimentation equilibrium at pH 7.5 and 4 and 20 degrees C. Concurrently, thermal unfolding of these forms of EI has been monitored by differential scanning calorimetry and by changes in the intrinsic tryptophanyl residue fluorescence. Phosphorylated EI and EI(H189E) have 10-fold increased dimerization constants [ approximately 2 x 10(6) (M monomer)(-1)] compared to those of dephospho-EI and EI(H189A) at 20 degrees C. Dimerization is strongly promoted by 1 mM PEP with 2 mM MgCl(2) [K(A)' > or = 10(8) M(-1) at 4 or 20 degrees C], as demonstrated with EI(H189A) which cannot undergo autophosphorylation. Together, 1 mM PEP and 2 mM Mg(2+) also markedly stabilize and couple the unfolding of C- and N-terminal domains of EI(H189A), increasing the transition temperature (T(m)) for unfolding the C-terminal domain by approximately 18 degrees C and that for the N-terminal domain by approximately 9 degrees C to T(max) congruent with 63 degrees C, giving a value of K(D)' congruent with 3 microM PEP at 45 degrees C. PEP alone also promotes the dimerization of EI(H189A) but only increases T(m) approximately 5 degrees C for C-terminal domain unfolding without affecting N-terminal domain unfolding, giving an estimated value of K(D)' congruent with 0.2 mM for PEP dissociation in the absence of Mg(2+) at 45 degrees C. In contrast, the dimerization constant of phospho-EI at 20 degrees C is the same in the absence and presence of 5 mM PEP and 2 mM MgCl(2). Thus, the separation of substrate binding effects from those of phosphorylation by studies with the inactive EI(H189A) has shown that intracellular concentrations of PEP and Mg(2+) are important determinants of both the conformational stability and dimerization of dephospho-EI.  相似文献   

4.
The phosphotransferase system (PTS) is involved in the use of carbon sources in bacteria. It is formed by two general proteins: enzyme I (EI) and the histidine phosphocarrier (HPr), and various sugar-specific permeases. EI is formed by two domains, with the N-terminal domain (EIN) being responsible for the binding to HPr. In low-G+C Gram-positive bacteria, HPr becomes phosphorylated not only by phosphoenolpyruvate (PEP) at the active-site histidine, but also by ATP at a serine. In this work, we have characterized: (i) the stability and binding affinities between the active-site-histidine phosphorylated species of HPr and the EIN from Streptomyces coelicolor; and (ii) the stability and binding affinities of the species involving the phosphorylation at the regulatory serine of HPr(sc). Our results show that the phosphorylated active-site species of both proteins are less stable than the unphosphorylated counterparts. Conversely, the Hpr-S47D, which mimics phosphorylation at the regulatory serine, is more stable than wild-type HPr(sc) due to helical N-capping effects, as suggested by the modeled structure of the protein. Binding among the phosphorylated and unphosphorylated species is always entropically driven, but the affinity and the enthalpy vary widely.  相似文献   

5.
The structural and thermodynamic impact of phosphorylation on the interaction of the N-terminal domain of enzyme I (EIN) and the histidine phosphocarrier protein (HPr), the two common components of all branches of the bacterial phosphotransferase system, have been examined using NMR spectroscopy and isothermal titration calorimetry. His-189 is located at the interface of the alpha and alphabeta domains of EIN, resulting in rather widespread chemical shift perturbation upon phosphorylation, in contrast to the highly localized perturbations seen for HPr, where His-15 is fully exposed to solvent. Residual dipolar coupling measurements, however, demonstrate unambiguously that no significant changes in backbone conformation of either protein occur upon phosphorylation: for EIN, the relative orientation of the alpha and alphabeta domains remains unchanged; for HPr, the backbone /Psi torsion angles of the active site residues are unperturbed within experimental error. His --> Glu/Asp mutations of the active site histidines designed to mimic the phosphorylated states reveal binding equilibria that favor phosphoryl transfer from EIN to HPr. Although binding of phospho-EIN to phospho-HPr is reduced by a factor of approximately 21 relative to the unphosphorylated complex, residual dipolar coupling measurements reveal that the structures of the unphosphorylated and biphosphorylated complexes are the same. Hence, the phosphorylation states of EIN and HPr shift the binding equilibria predominantly by modulating intermolecular electrostatic interactions without altering either the backbone scaffold or binding interface. This facilitates highly efficient phosphoryl transfer between EIN and HPr, which is estimated to occur at a rate of approximately 850 s(-1) from exchange spectroscopy.  相似文献   

6.
Yu TK  Yun YJ  Lee KO  Ahn KJ  Suh JY 《FEBS letters》2012,586(10):1439-1444
The N-terminal domain of Enzyme I (EIN) and phosphocarrier HPr can form a biphosphorylated complex when they are both phosphorylated by excess cellular phosphoenolpyruvate. Here we show that the electrostatic repulsion between the phosphoryl groups in the biphosphorylated complex results in characteristic dynamics at the active site in a millisecond time scale. The dynamics is localized to phospho-His15 and the stabilizing backbone amide groups of HPr, and does not impact on the phospho-His189 of EIN. The dynamics occurs with the k(ex) of ~500 s(-1) which compares to the phosphoryl transfer rate of ~850 s(-1) between EIN and HPr. The conformational dynamics in HPr may be important for its phosphotransfer reactions with multiple partner proteins.  相似文献   

7.
Pujols-Ayala I  Barry BA 《Biochemistry》2002,41(38):11456-11465
In photosynthesis, photosystem II (PSII) conducts the light-driven oxidation of water to oxygen. Tyrosine Z is Tyr 161 of the D1 polypeptide; Z acts as an intermediary electron carrier in water oxidation. In this report, EPR spectroscopy was used to study the effect of His 190 and Glu 189 on Z* yield and reduction kinetics. Neither mutation has a significant impact on the EPR line shape of Z*. At room temperature and pH 7.5, the E189Q-D1 mutation has a single turnover Z* yield that is 84% compared to wild-type. The H190Q-D1 mutation decreases the Z* yield at room temperature by a factor of 2.6 but has a more modest effect (factor of 1.6) at -10 degrees C. The temperature dependence is shown to be primarily reversible. Neither mutation has a dramatic effect on Z* decay kinetics. The Z* minus Z FT-IR spectrum, recorded at pH 7.5 on H190Q, reveals perturbations, including an increased spectral contribution from a PSII chlorophyll. The Z* minus Z FT-IR spectrum, recorded at pH 7.5 on E189Q, shows perturbations, including a decreased contribution from the carboxylate side chain of a glutamate or aspartate. Temperature-dependent changes in H190Q-D1 and E189Q-D1 Z. yield are attributed to a reversible conformational change, which alters the electron-transfer rate from Z to P(680)(+). On the basis of these results, we conclude that H190 and E189 play a role in the structural stabilization of PSII. We postulate that some or all of the phenotypic changes observed in H190Q and E189Q mutants may be caused by structural alterations in PSII.  相似文献   

8.
Staphylococcal nuclease mutants, E57G and E75G, were generated. A comparison of the kinetic parameters both for mutants and wild-type protein shows that the Michaelis constants (Km) were almost identical for the wild-type protein and E57G mutant. An approximately 30-fold decrease in Km compared with the wild-type protein was observed for the E75G mutant. The turnover numbers for the enzyme (kcat) were higher with both the wild-type protein and the E57G mutant (3.88 +/- 0.21 x 103 s-1 and 3.71 +/- 0.28 x 103 s-1) than with the E75G mutant (3.04 +/- 0.02 x 102 s-1). The results of thermal denaturation with differential scanning microcalorimetry indicate that the excess calorimetric enthalpy of denaturations, DeltaHcal, was almost identical for the wild-type protein and E57G mutant (84.1 +/- 6.2 kcal.mol-1 and 79.3 +/- 7.1 kcal.mol-1, respectively). An approximately twofold decrease in DeltaHcal compared with the wild-type protein was observed for the E75G mutant (42.7 +/- 5.5 kcal.mol-1). These outcomes imply that Glu at position 75 plays a significant role in maintaining enzyme activity and protein stability. Further study of the unfolding of the wild-type protein and E75G mutant was conducted by using time-resolved fluorescence with a picosecond laser pulse. Two fluorescent lifetimes were found in the subnanosecond time range. The faster lifetime (tau2) did not generally vary with either pH or the concentration of guanidinium hydrochloride (GdmHCl) in the wild-type protein and the E75G mutant. The slow lifetime (tau1), however, did vary with these parameters and was faster as the protein is unfolded by either pH or GdmHCl denaturation. The midpoints of the transition for tau1 are pH 3.5 and 5.8 for the wild-type protein and E75G mutant, respectively, and the GdmHCl concentrations are 1.1 m and 0.6 m for the wild-type protein and E75G mutant, respectively. Parallel steady-state fluorescence measurements have also been carried out and the results are in general agreement with the time-resolved fluorescence experiments, indicating that Glu at position 75 plays an important role in protein unfolding.  相似文献   

9.
A mutant version of the N-terminal domain of Escherichia coli DnaB helicase was used as a model system to assess the stabilization against unfolding gained by covalent cyclization. Cyclization was achieved in vivo by formation of an amide bond between the N and C termini with the help of a split mini-intein. Linear and circular proteins were constructed to be identical in amino acid sequence. Mutagenesis of Phe102 to Glu rendered the protein monomeric even at high concentration. A difference in free energy of unfolding, DeltaDeltaG, between circular and linear protein of 2.3(+/-0.5) kcal mol(-1) was measured at 10 degrees C by circular dichroism. A theoretical estimate of the difference in conformational entropy of linear and circular random chains in a three-dimensional cubic lattice model predicted DeltaDeltaG=2.3 kcal mol(-1), suggesting that stabilization by protein cyclization is driven by the reduced conformational entropy of the unfolded state. Amide-proton exchange rates measured by NMR spectroscopy and mass spectrometry showed a uniform, approximately tenfold decrease of the exchange rates of the most slowly exchanging amide protons, demonstrating that cyclization globally decreases the unfolding rate of the protein. The amide proton exchange was found to follow EX1 kinetics at near-neutral pH, in agreement with an unusually slow refolding rate of less than 4 min(-1) measured by stopped-flow circular dichroism. The linear and circular proteins differed more in their unfolding than in their folding rates. Global unfolding of the N-terminal domain of E.coli DnaB is thus promoted strongly by spatial separation of the N and C termini, whereas their proximity is much less important for folding.  相似文献   

10.
Thermoanaerobacter tengcongensis is a thermophilic eubacterium that has a phosphoenolpyruvate (PEP) sugar phosphotransferase system (PTS) of 22 proteins. The general PTS proteins, enzyme I and HPr, and the transporters for N-acetylglucosamine (EIICB(GlcNAc)) and fructose (EIIBC(Fru)) have thermal unfolding transitions at ~90 °C and a temperature optimum for in vitro sugar phosphotransferase activity of 65 °C. The phosphocysteine of a EIICB(GlcNAc) mutant is unusually stable at room temperature with a t(1/2) of 60 h. The PEP binding C-terminal domain of enzyme I (EIC) forms a metastable covalent adduct with PEP at 65 °C. Crystallization of this adduct afforded the 1.68 ? resolution structure of EIC with a molecule of pyruvate in the active site. We also report the 1.83 ? crystal structure of the EIC-PEP complex. The comparison of the two structures with the apo form and with full-length EI shows differences between the active site side chain conformations of the PEP and pyruvate states but not between the pyruvate and apo states. In the presence of PEP, Arg465 forms a salt bridge with the phosphate moiety while Glu504 forms salt bridges with Arg186 and Arg195 of the N-terminal domain of enzyme I (EIN), which stabilizes a conformation appropriate for the in-line transfer of the phosphoryl moiety from PEP to His191. After transfer, Arg465 swings 4.8 ? away to form an alternative salt bridge with the carboxylate of Glu504. Glu504 loses the grip of Arg186 and Arg195, and the EIN domain can swing away to hand on the phosphoryl group to the phosphoryl carrier protein HPr.  相似文献   

11.
The conformational stability and reversibility of unfolding of the human dimeric enzyme Cu Zn superoxide dismutase (HSOD) and the three mutant enzymes constructed by replacement of Cys6 by Ala and Cys111 by Ser, singly and in combination, were determined by differential scanning calorimetry. The differential scanning calorimetry profile of wild-type HSOD consists of two components, which probably represent the unfolding of the oxidized and reduced forms of the enzyme, with denaturation temperatures (Tm) of 74.9 and 83.6 degrees C, approximately 7 degrees lower than those for bovine superoxide dismutase (BSOD). The conformational stabilities of the two components of the mutant HSOD's differ only slightly from those of the wild type (delta delta Gs of -0.2 to +0.8 kcal/mol of dimer), while replacement of the BSOD Cys6 by Ala is somewhat destabilizing (delta delta G of -0.7 to -1.3 kcal/mol of dimer). These small alterations in conformational stability do not correlate with the large increases in resistance to thermal inactivation following substitution of free Cys in both HSOD and BSOD (McRee, D.E., Redford, S.M., Getzoff, E.D., Lepock, J.R., Hallewell, R.A., and Tainer, J.A. (1990) J. Biol. Chem. 265, 14234-14241 and Hallewell, R.A., Imlay, K.C., Laria, I., Gallegos, C., Fong, N., Irvine, B., Getzoff, E.D., Tainer, J.A., Cubelli, D.E., Bielski, B.H.J., Olson, P., Mallenbach, G.T., and Cousens, L.S. (1991) Proteins Struct. Funct. Genet., submitted for publication). The reversibility of unfolding was determined by scanning part way through the profile, cooling, rescanning, and calculating the amount of protein irreversibly unfolded by the first scan. The order of reversibility at a constant level of unfolding is the same as the order of resistance to inactivation for both the HSOD and BSOD wild-type and mutant enzymes. Thus, the greater resistance to thermal inactivation of the superoxide dismutase enzymes with free Cys replaced by Ala or Ser is dominated by a greater resistance to irreversible unfolding and relatively unaffected by changes in conformational stability.  相似文献   

12.
HPr is a protein of the bacterial phosphoenolpyruvate:sugar phosphotransferase transport system (PTS). In Gram-positive bacteria, HPr can be phosphorylated on Ser(46) by HPr(Ser) kinase/phosphorylase (HPrK/P) and on His(15) by enzyme I (EI) of the PTS. In vitro studies have shown that phosphorylation on one residue greatly inhibits the second phosphorylation. However, streptococci contain significant amounts of HPr(Ser-P)(His approximately P) during exponential growth, and recent studies suggest that phosphorylation of HPr(Ser-P) by EI is involved in the recycling of HPr(Ser-P)(His approximately P). We report in this paper a study on the phosphorylation of Streptococcus salivarius HPr, HPr(Ser-P), and HPr(S46D) by EI. Our results indicate that (i) the specificity constant (k(cat)/K(m)) of EI for HPr(Ser-P) at pH 7.9 was approximately 5000-fold smaller than that observed for HPr, (ii) no metabolic intermediates were able to stimulate HPr(Ser-P) phosphorylation, (iii) the rate of HPr phosphorylation decreased at pHs below 6.5, while that of HPr(Ser-P) increased and was almost 10-fold higher at pH 6.1 than at pH 7.9, (iv) HPr(S46D), a mutated HPr alleged to mimic HPr(Ser-P), was also phosphorylated more efficiently under acidic conditions, and, lastly, (v) phosphorylation of Bacillus subtilis HPr(Ser-P) by B. subtilis EI was also stimulated at acidic pH. Our results suggest that the high levels of HPr(Ser-P)(His approximately P) in streptococci result from the combination of two factors, a high physiological concentration of HPr(Ser-P) and stimulation of HPr(Ser-P) phosphorylation by EI at acidic pH, an intracellular condition that occurs in response to the acidification of the external medium during growth of the culture.  相似文献   

13.
M Zolkiewski  A Ginsburg 《Biochemistry》1992,31(48):11991-12000
Dodecameric glutamine synthetase (GS) from Escherichia coli undergoes reversible, thermally induced partial unfolding without subunit dissociation. A single endotherm for Mn.GS (+/- active-site ligands) in the presence of 1 mM free Mn2+ and 100 mM KCl at pH 7 is observed by differential scanning calorimetry (DSC). Previous deconvolutions of DSC data for Mn.GS showed only two two-state transitions (with similar tm values; 51.6 +/- 2 degrees C), and indicated that cooperative interactions link partial unfolding reactions of all subunits within the Mn.enzyme dodecamer [Ginsburg, A., & Zolkiewski, M. (1991) Biochemistry 30, 9421]. A net uptake of 8.0 equiv of H+ by Mn.GS occurs during partial unfolding, as determined in the present DSC experiments conducted with four buffers having different heats of protonation at 50 degrees C. These data gave a value of 176 +/- 12 kcal (mol of dodecamer)-1 for delta Hcal corrected for buffer protonation. L-Glutamine and L-Met-(SR)-sulfoximine stabilize the Mn.GS dodecamer through the free energies of ligand binding, and these were shown to be partially and totally released, respectively, from the 12 active sites at high temperature. Ligand effects on Tm values from DSC were similar to those from spectral measurements of Trp and Tyr exposures in two subunit domains. Effects of varying [ADP] on DSC profiles of Mn.GS were complex; Tm is increased by low [ADP] and decreased by > 100 microM free ADP. This is due to the exposure of an additional low-affinity ADP binding site per GS subunit at high temperature with log K1' = 4.3 and log K2' = 3.6 at 60 degrees C relative to log K' = 5.5 for ADP binding at 30 degrees C, as determined by isothermal calorimetric and fluorescence titrations. Moreover, delta Hcal at > 27% saturation with ADP (corrected for ADP binding/dissociation) is approximately 80-100 kcal/mol more than in the absence of ligands. Changes in domain interactions could result from ADP bridging subunit contacts in the dodecamer. Each of the active-site ligands investigated here produces different effects on DSC profiles without uncoupling the extremely cooperative, partial unfolding reactions in the Mn.GS dodecamer.  相似文献   

14.
Ketosteroid isomerase (KSI) from Pseudomonas putida biotype B is a homodimeric enzyme catalyzing an allylic isomerization of Delta(5)-3-ketosteroids at a rate of the diffusion-controlled limit. The dimeric interactions mediated by Arg72, Glu118, and Asn120, which are conserved in the homologous KSIs, have been characterized in an effort to investigate the roles of the conserved interface residues in stability, function and structure of the enzyme. The interface residues were replaced with alanine to generate the interface mutants R72A, E118A, N120A and E118A/N120A. Equilibrium unfolding analysis revealed that the DeltaG(U)(H(2)O) values for the R72A, E118A, N120A, and E118A/N120A mutants were decreased by about 3.8, 3.9, 7.8, and 9.5 kcal/mol, respectively, relative to that of the wild-type enzyme. The interface mutations not only decreased the k(cat)/K(M) value by about 8- to 96-fold, but also increased the K(D) value for d-equilenin, a reaction intermediate analogue, by about 7- to 17.5-fold. The crystal structure of R72A determined at 2.5 A resolution and the fluorescence spectra of all the mutants indicated that the interface mutations altered the active-site geometry and resulted in the decreases of the conformational stability as well as the catalytic activity of KSI. Taken together, our results strongly suggest that the conserved interface residues contribute to stabilization and structural integrity of the active site in the dimeric KSI.  相似文献   

15.
Point mutations at the dimer interface of the homodimeric enzyme ascorbate peroxidase (APx) were constructed to assess the role of quaternary interactions in the stability and activity of APx. Analysis of the APx crystal structure shows that Glu112 forms a salt bridge with Lys20 and Arg24 of the opposing subunit near the axis of dyad symmetry between the subunits. Two point mutants, E112A and E112K, were made to determine the effects of a neutral (alanine) and repulsive (lysine) mutation on dimerization, stability, and activity. Gel filtration analysis indicated that the ratio of the monomer to dimer increased as the dimer interface interactions went from attractive to repulsive. Differential scanning calorimetry (DSC) data exhibited a decrease in both the transition temperature (Tm) and enthalpy of unfolding (deltaHc) with Tm = 58.3 +/- 0.5 degrees C, 56.0 +/- 0.8 degrees C, and 53.0 +/- 0.9 degrees C and deltaHc = 245 +/- 29 kcal/mol, 199 +/- 38 kcal/mol, and 170 +/- 25 kcal/mol for wild-type APx, E112A, and E112K, respectively. Similar changes were observed based on thermal melting curves obtained by absorption spectroscopy. No change in enzyme activity was found for the E112A mutant, and only a 25% drop in activity was observed for the E112K mutant which demonstrates that the non-Michaelis Menten kinetics of APx is not due to the APx oligomeric structure. The cryogenic crystal structures of the wild-type and mutant proteins show that mutation induced changes are limited to the dimer interface including an alteration in solvent structure.  相似文献   

16.
M E Holtzer  A Holtzer 《Biopolymers》1990,30(9-10):985-993
Circular dichroism (CD) experiments in the backbone (200-240 nm) region are reported for four isolated, excised two-chain, coiled-coil segments whose chains comprise, respectively, residues 11-127, 142-281, 1-189, and 190-284 of the rabbit alpha alpha-tropomyosin (Tm) sequence. The uv and CD spectra for the noncross-linked segments are very similar to those for parent Tm. At 3 degrees C, all have a helix content of 90% or more; moreover, all thermal denaturation curves depend on concentration, as required by mass action, and are completely reversible. At comparable concentrations, solutions show values of T1/2 (the temperature at which the helix content is 50%) following the order of 11Tm127 approximately 1Tm189 greater than 142Tm281 greater than 190Tm284. The thermal unfolding data for 11Tm127, 190Tm284, and 142Tm281 fall on apparently monophasic curves (single inflection point). However, curves for 1Tm189 show a heretofore unknown low temperature transition in which the helix content drops from approximately 90% at 2 degrees C to approximately 73% at 20 degrees C, indicating that this segment has one or more weak sections totaling approximately 50 residues per chain. Since thermal denaturation curves for noncross-linked 11Tm127, 142Tm281, and Tm have no such low temperature transition, i.e., the helix content is not additive, the weak region probably comprises the bulk of the residues between 127 and 189 in 1Tm189, but is somehow stabilized in 142Tm281 and in parent Tm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Conformational stability and mechanism of folding of ribonuclease T1   总被引:5,自引:0,他引:5  
Urea and thermal unfolding curves for ribonuclease T1 (RNase T1) were determined by measuring several different physical properties. In all cases, steep, single-step unfolding curves were observed. When these results were analyzed by assuming a two-state folding mechanism, the plots of fraction unfolded protein versus denaturant were coincident. The dependence of the free energy of unfolding, delta G (in kcal/mol), on urea concentration is given by delta G = 5.6 - 1.21 (urea). The parameters characterizing the thermodynamics of unfolding are: midpoint of the thermal unfolding curve, Tm = 48.1 degrees C, enthalpy change at Tm, delta Hm = 97 kcal/mol, and heat capacity change, delta Cp = 1650 cal/mol deg. A single kinetic phase was observed for both the folding and unfolding of RNase T1 in the transition and post-transition regions. However, two slow kinetic phases were observed during folding in the pre-transition region. These two slow phases account for about 90% of the observed amplitude, indicating that a faster kinetic phase is also present. The slow phases probably result from cis-trans isomerization at the 2 proline residues that have a cis configuration in folded RNase T1. These results suggest that RNase T1 folds by a highly cooperative mechanism with no structural intermediates once the proline residues have assumed their correct isomeric configuration. At 25 degrees C, the folded conformation is more stable than the unfolded conformations by 5.6 kcal/mol at pH 7 and by 8.9 kcal/mol at pH 5, which is the pH of maximum stability. At pH 7, the thermodynamic data indicate that the maximum conformational stability of 8.3 kcal/mol will occur at -6 degrees C.  相似文献   

18.
The conformational stability of the histidine-containing phosphocarrier protein (HPr) from Bacillus subtilis has been determined using a combination of thermal unfolding and solvent denaturation experiments. The urea-induced denaturation of HPr was monitored spectroscopically at fixed temperatures and thermal unfolding was performed in the presence of fixed concentrations of urea. These data were analyzed in several different ways to afford a measure of the cardinal parameters (delta Hg, Tg, delta Sg, and delta Cp) that describe the thermodynamics of folding for HPr. The method of Pace and Laurents (Pace CN, Laurents DV, 1989, Biochemistry 28:2520-2525) was used to estimate delta Cp as was a global analysis of the thermal- and urea-induced unfolding data. Each method used to analyze the data gives a similar value for delta Cp (1,170 +/- 50 cal mol-1K-1). Despite the high melting temperature for HPr (Tg = 73.5 degrees C), the maximum stability of the protein, which occurs at 26 degrees C, is quite modest (delta Gs = 4.2 kcal mol-1). In the presence of moderate concentrations of urea, HPr exhibits cold denaturation, and thus a complete stability curve for HPr, including a measure of delta Cp, can be achieved using the method of Chen and Schellman (Chen B, Schellman JA, 1989, Biochemistry 28:685-691). A comparison of the different methods for the analysis of solvent denaturation curves is provided and the effects of urea on the thermal stability of this small globular protein are discussed. The methods presented will be of general utility in the characterization of the stability curve for many small proteins.  相似文献   

19.
Enzyme I (EI), the first component of the phosphoenolpyruvate (PEP):sugar phosphotransferase system (PTS), consists of an N-terminal protein-binding domain (EIN) and a C-terminal PEP-binding domain (EIC). EI transfers phosphate from PEP by double displacement via a histidine residue on EIN to the general phosphoryl carrier protein HPr. Here, we report the 1.82A crystal structure of the homodimeric EIC domain from Thermoanaerobacter tengcongensis, a saccharolytic eubacterium that grows optimally at 75 degrees C. EIC folds into a (betaalpha)(8) barrel with three large helical insertions between beta2/alpha2, beta3/alpha3 and beta6/alpha6. The large amphipathic dimer interface buries 3750A(2) of accessible surface area per monomer. A comparison with pyruvate phosphate dikinase (PPDK) reveals that the active-site residues in the empty PEP-binding site of EIC and in the liganded PEP-binding site of PPDK have almost identical conformations, pointing to a rigid structure of the active site. In silico models of EIC in complex with the Z and E-isomers of chloro-PEP provide a rational explanation for their difference as substrates and inhibitors of EI. The EIC domain exhibits 54% amino acid sequence identity with Escherichia coli and 60% with Bacillus subtilis EIC, has the same amino acid composition but contains additional salt-bridges and a more complex salt-bridge network than the homology model of E.coli EIC. The easy crystallization of EIC suggests that T.tengcongensis can serve as source for stable homologs of mesophilic proteins that are too labile for crystallization.  相似文献   

20.
The bacterial phosphoenolpyruvate (PEP) sugar phosphotransferase system mediates sugar uptake and controls the carbon metabolism in response to carbohydrate availability. Enzyme I (EI), the first component of the phosphotransferase system, consists of an N-terminal protein binding domain (EIN) and a C-terminal PEP binding domain (EIC). EI transfers phosphate from PEP by double displacement via a histidine residue on EIN to the general phosphoryl carrier protein HPr. Here we report the 2.4 Å crystal structure of the homodimeric EI from Staphylococcus aureus. EIN consists of the helical hairpin HPr binding subdomain and the phosphorylatable βα phospho-histidine (P-His) domain. EIC folds into an (βα)8 barrel. The dimer interface of EIC buries 1833 Å2 of accessible surface per monomer and contains two Ca2+ binding sites per dimer. The structures of the S. aureus and Escherichia coli EI domains (Teplyakov, A., Lim, K., Zhu, P. P., Kapadia, G., Chen, C. C., Schwartz, J., Howard, A., Reddy, P. T., Peterkofsky, A., and Herzberg, O. (2006) Proc. Natl. Acad. Sci. U.S.A. 103, 16218–16223) are very similar. The orientation of the domains relative to each other, however, is different. In the present structure the P-His domain is docked to the HPr binding domain in an orientation appropriate for in-line transfer of the phosphate to the active site histidine of the acceptor HPr. In the E. coli structure the phospho-His of the P-His domain projects into the PEP binding site of EIC. In the S. aureus structure the crystallographic temperature factors are lower for the HPr binding domain in contact with the P-His domain and higher for EIC. In the E. coli structure it is the reverse.  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