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.
The energetic contributions of the protein to the redox potential in an iron-sulfur protein are studied via energy minimization, comparing homologous rubredoxins from Clostridium pasteurianum, Desulfovibrio gigas, Desulfovibrio vulgaris, and Pyrococcus furiosus. The reduction reaction was divided into 1) the change in the redox site charge without allowing the protein to respond and 2) the relaxation of the protein in response to the new charge state, focusing on the latter. The energy minimizations predict structural relaxation near the redox site that agrees well with that in crystal structures of oxidized and reduced P. furiosus rubredoxin, but underpredicts it far from the redox site. However, the relaxation energies from the energy-minimized structures agree well with those from the crystal structures, because the polar groups near the redox site are the main determinants and the charged groups are all located at the surface and thus are screened dielectrically. Relaxation energies are necessary for good agreement with experimentally observed differences in reduction energies between C. pasteurianum and the other three rubredoxins. Overall, the relaxation energy is large (over 500 mV) from both the energy-minimized and the crystal structures. In addition, the range in the relaxation energy for the different rubredoxins is large (300 mV), because even though the structural perturbations of the polar groups are small, they are very near the redox site. Thus the relaxation energy is an important factor to consider in reduction energetics.  相似文献   

2.
The structures of the oxidized and reduced forms of the rubredoxin from the archaebacterium, Pyrococcus furiosus, an organism that grows optimally at 100 degrees C, have been determined by X-ray crystallography to a resolution of 1.8 A. Crystals of this rubredoxin grow in space group P2(1)2(1)2(1) with room temperature cell dimensions a = 34.6 A, b = 35.5 A, and c = 44.4 A. Initial phases were determined by the method of molecular replacement using the oxidized form of the rubredoxin from the mesophilic eubacterium, Clostridium pasteurianum, as a starting model. The oxidized and reduced models of P. furiosus rubredoxin each contain 414 nonhydrogen protein atoms comprising 53 residues. The model of the oxidized form contains 61 solvent H2O oxygen atoms and has been refined with X-PLOR and TNT to a final R = 0.178 with root mean square (rms) deviations from ideality in bond distances and bond angles of 0.014 A and 2.06 degrees, respectively. The model of the reduced form contains 37 solvent H2O oxygen atoms and has been refined to R = 0.193 with rms deviations from ideality in bond lengths of 0.012 A and in bond angles of 1.95 degrees. The overall structure of P. furiosus rubredoxin is similar to the structures of mesophilic rubredoxins, with the exception of a more extensive hydrogen-bonding network in the beta-sheet region and multiple electrostatic interactions (salt bridge, hydrogen bonds) of the Glu 14 side chain with groups on three other residues (the amino-terminal nitrogen of Ala 1; the indole nitrogen of Trp 3; and the amide nitrogen group of Phe 29). The influence of these and other features upon the thermostability of the P. furiosus protein is discussed.  相似文献   

3.
G Hernández  D M LeMaster 《Biochemistry》2001,40(48):14384-14391
Spatially localized differences in the conformational dynamics of the rubredoxins from the hyperthermophile Pyrococcus furiosus (Pf) and the mesophile Clostridium pasteurianum (Cp) are monitored via amide exchange measurements. As shown previously for the hyperthermophile protein, nearly all backbone amides of the Cp rubredoxin exhibit EX(2) hydrogen exchange kinetics with conformational opening rates of >1 s(-)(1). Significantly slower amide exchange is observed for Pf rubredoxin in the region surrounding the metal site and the proximal end of the three-stranded beta-sheet, while for the rest of the structure, the exchange rates at 23 degrees C are similar for both proteins. For the multiple-turn region comprising residues 14-32 in both rubredoxins, the uniformity of both the exchange rate constants and the values of the activation energy at the slowly exchanging sites is consistent with a model of solvent exposure via a subglobal cooperative conformational opening. In contrast to the common expectation of increased rigidity in the hyperthermophile proteins, below room temperature Pf rubredoxin exhibits a larger apparent flexibility in this multiple-turn region. The smaller enthalpy for the conformational opening process of this region in Pf rubredoxin reflects the much weaker temperature dependence of the underlying conformational equilibrium in the hyperthermophile protein.  相似文献   

4.
The purification, amino acid sequence, and two-dimensional 1H NMR results are reported for the rubredoxin (Rd) from the hyperthermophilic archaebacterium Pyrococcus furiosus, an organism that grows optimally at 100 degrees C. The molecular mass (5397 Da), iron content (1.2 +/- 0.2 g-atom of Fe/mol), UV-vis spectrophotometric properties, and amino acid sequence (60% sequence identity with Clostridium pasteurianum Rd) are found to be typical of this class of redox protein. However, P. furiosus Rd is remarkably thermostable, being unaffected after incubation for 24 h at 95 degrees C. One- and two-dimensional 1H nuclear magnetic resonance spectra of the oxidized [Fe(III)Rd] and reduced [Fe(II)Rd] forms of P. furiosus Rd exhibited substantial paramagnetic line broadening, and this precluded detailed 3D structural studies. The apoprotein was not readily amenable to NMR studies due to apparent protein oxidation involving the free cysteine sulfhydryls. However, high-quality NMR spectra were obtained for the Zn-substituted protein, Zn(Rd), enabling detailed NMR signal assignment for all backbone amide and alpha and most side-chain protons. Secondary structural elements were determined from qualitative analysis of 2D Overhauser effect spectra. Residues A1-K6, Y10-E14, and F48-E51 form a three-strand antiparallel beta-sheet, which comprises ca. 30% of the primary sequence. Residues C5-Y10 and C38-A43 form types I and II amide-sulfur tight turns common to iron-sulfur proteins. These structural elements are similar to those observed by X-ray crystallography for native Rd from the mesophile C. pasteurianum. However, the beta-sheet domain in P. furiosus Rd is larger than that in C. pasteurianum Rd and appears to begin at the N-terminal residue. From analysis of the secondary structure, potentially stabilizing electrostatic interactions involving the charged groups of residues Ala(1), Glu(14), and Glu(52) are proposed. These interactions, which are not present in rubredoxins from mesophilic organisms, may prevent the beta-sheet from "unzipping" at elevated temperatures.  相似文献   

5.
The branched sidechain residues 24 and 33 in the hydrophobic core of rubredoxin differ between the Clostridium pasteurianum (Cp) and Pyrococcus furiosus (Pf) sequences. Their X-ray structures indicate that these two sidechains are in van der Waals contact with each other, while neither appears to significantly interact with the other nonconserved residues. The simultaneous interchange of residues 24 and 33 between the Cp and Pf rubredoxin sequences yield a complementary pair of hybrid proteins for which the sum of their thermodynamic stabilities equals that of the parental rubredoxins. The 1.2 kcal/mol change arising from this two residues interchange accounts for 21% of the differential thermodynamic stability between the mesophile and hyperthermophile proteins. The additional interchange of the sole nonconserved aromatic residue in the hydrophobic core yields a 0.78 kcal/mol deviation from thermodynamic additivity.  相似文献   

6.
Tan ML  Kang C  Ichiye T 《Proteins》2006,62(3):708-714
Rubredoxins may be separated into high and low reduction potential classes, with reduction potentials differing by approximately 50 mV. Our previous work showed that a local shift in the polar backbone due to an A(44) versus V(44) side-chain size causes this reduction potential difference. However, this work also indicated that in the low potential Clostridium pasteurianum (Cp) rubredoxin, a V(44) --> A(44) mutation causes larger local backbone flexibility, because the V(44) side-chain present in the wild-type (wt) is no longer present to interlock with neighboring residues to stabilize the subsequent G(45). Since Pyrococcus furiosus (Pf) and other high potential rubredoxins generally have a P(45), it was presumed that a G(45) --> P(45) mutation might stabilize a V(44) --> A(44) mutation in Cp rubredoxin. Here crystal structure analysis, energy minimization, and molecular dynamics (MD) were performed for wt V(44)G(45), single mutant A(44)G(45) and double mutant A(44)P(45) Cp, and for wt A(44)P(45) Pf rubredoxins. The local structural, dynamical, and electrostatic properties of Cp gradually approach wt Pf in the order wt Cp to single to double mutant because of greater sequence similarity, as expected. The double mutant A(44)P(45) Cp exhibits increased backbone stability near residue 44 and thus enhances the probability that the backbone dipoles point toward the redox site, which favors an increase in the electrostatic contribution to the reduction potential. It appears that the electrostatic potential of residue 44 and the solvent accessibility to the redox are both determinants for the reduction potentials of homologous rubredoxins. Overall, these results indicate that an A(44) in a rubredoxin may require a P(45) for backbone stability whereas a V(44) can accommodate a G(45), since the valine side-chain can interlock with its neighbors.  相似文献   

7.
S Aono  F O Bryant    M W Adams 《Journal of bacteriology》1989,171(6):3433-3439
The archaebacterium Pyrococcus furiosus is a strict anaerobe that grows optimally at 100 degrees C by a fermentative-type metabolism in which H2 and CO2 are the only detectable products. A ferredoxin, which functions as the electron donor to the hydrogenase of this organism was purified under anaerobic reducing conditions. It had a molecular weight of approximately 12,000 and contained 8 iron atoms and 8 cysteine residues/mol but lacked histidine or arginine residues. Reduction and oxidation of the ferredoxin each required 2 electrons/mol, which is consistent with the presence of two [4Fe-4S] clusters. The reduced protein gave rise to a broad rhombic electronic paramagnetic resonance spectrum, with gz = 2.10, gy = 1.86, gx = 1.80, and a midpoint potential of -345 mV (at pH 8). However, this spectrum represented a minor species, since it quantitated to only approximately 0.3 spins/mol. P. furiosus ferredoxin is therefore distinct from other ferredoxins in that the bulk of its iron is not present as iron-sulfur clusters with an S = 1/2 ground state. The apoferredoxin was reconstituted with iron and sulfide to give a protein that was indistinguishable from the native ferredoxin by its iron content and electron paramagnetic resonance properties, which showed that the novel iron-sulfur clusters were not artifacts of purification. The reduced ferredoxin also functioned as an electron donor for H2 evolution catalyzed by the hydrogenase of the mesophilic eubacterium Clostridium pasteurianum. P. furiosus ferredoxin was resistant to denaturation by sodium dodecyl sulfate (20%, wt/vol) and was remarkably thermostable. Its UV-visible absorption spectrum and electron carrier activity to P. furiosus hydrogenase were unaffected by a 12-h incubation of 95 degrees C.  相似文献   

8.
The role of solvent composition and temperature on equilibrium electron transfer in seven rubredoxin variants [ Clostridium pasteurianum ( Cp), V8D, V8R, V8A, V44A Cp, Pyrococcus furiosus ( Pf), and A44V Pf] were investigated to examine the role of both monopolar and dipolar interactions. The reduction potentials of all variants decreased as the polarity of the solvent decreased. The enthalpy and entropy associated with electron transfer were determined from temperature-controlled voltammetric studies. The entropic contribution [delta( Tdelta S degrees )] to the change in the reduction potential was larger for charged variants (V8D and V8R), while the enthalpic contribution [delta(-delta H degrees )] was larger for the other mutants. The large entropy change observed for monopolar variants is likely due to solvent reorganization that occurs between oxidation states. Entropic-enthalpic compensation phenomena, an observation that most proteins have an entropic term [delta( Tdelta S degrees )] and enthalpic term [delta(-delta H degrees )] with opposite signs, was observed. A correlation of the size of the amino acid side chain with delta E degrees ', delta(-delta H degrees ), and delta( Tdelta S degrees ) is also discussed.  相似文献   

9.
Kleeb AC  Kast P  Hilvert D 《Biochemistry》2006,45(47):14101-14110
Prephenate dehydratase (PDT) is an important but poorly characterized enzyme that is involved in the production of L-phenylalanine. Multiple-sequence alignments and a phylogenetic tree suggest that the PDT family has a common structural fold. On the basis of its sequence, the PDT from the extreme thermophile Methanocaldococcus jannaschii (MjPDT) was chosen as a promising representative of this family for pursuing structural and functional studies. The corresponding pheA gene was cloned and expressed in Escherichia coli. It encodes a monofunctional and thermostable enzyme with an N-terminal catalytic domain and a C-terminal regulatory ACT domain. Biophysical characterization suggests a dimeric (62 kDa) protein with mixed alpha/beta secondary structure elements. MjPDT unfolds in a two-state manner (Tm = 94 degrees C), and its free energy of unfolding [DeltaGU(H2O)] is 32.0 kcal/mol. The purified enzyme catalyzes the conversion of prephenate to phenylpyruvate according to Michaelis-Menten kinetics (kcat = 12.3 s-1 and Km = 22 microM at 30 degrees C), and its activity is pH-independent over the range of pH 5-10. It is feedback-inhibited by L-phenylalanine (Ki = 0.5 microM), but not by L-tyrosine or L-tryptophan. Comparison of its activation parameters (DeltaH(++)= 15 kcal/mol and DeltaS(++)= -3 cal mol-1 K-1) with those for the spontaneous reaction (DeltaH(++) = 17 kcal/mol and DeltaS(++)= -28 cal mol-1 K-1) suggests that MjPDT functions largely as an entropy trap. By providing a highly preorganized microenvironment for the dehydration-decarboxylation sequence, the enzyme may avoid the extensive solvent reorganization that accompanies formation of the carbocationic intermediate in the uncatalyzed reaction.  相似文献   

10.
At 1.0 atm pressure and in 150 mM sodium phosphate (pH = 7.0), bovine carbonic anhydrase undergoes a nondenaturational conformational change at 30.3 degrees C and an unfolding transition from the physiological conformer to the molten globule state at 67.4 degrees C. The pressure dependences of the temperatures of these transitions have been studied under reversible conditions for the purpose of understanding DeltaH degrees , DeltaS degrees , and DeltaV for each conformational change. Temperatures for the low-temperature to physiological-temperature conformational change T(L-->P) are obtained from physiologically relevant conditions using slow-scan-rate differential scanning calorimetry. Temperatures for the physiological-temperature conformation to molten globule state conversion T(P-->MG) are obtained from differential scanning calorimetry measurements of the apparent transition temperature in the presence of guanidine hydrochloride extrapolated to zero molar denaturant. The use of slow-scan-rate differential scanning calorimetry permits the calculation of the activation volume for the conversion of the low-temperature conformer to the physiological-temperature conformer DeltaV(double dagger)(L-->P). At 1.0 atm pressure, the transition from the low-temperature conformer to the physiological-temperature conformer involves a volume change DeltaV(L-->P) = 15 +/- 2 L/mole, which contrasts with the partial unfolding of the physiological-temperature conformer to the molten globule state (DeltaV(P-->MG) = 26 +/- 9 L/mole). The activation volume for this process DeltaV(double dagger)(L-->P) = 51 +/- 9 L/mole and is consistent with a prior thermodynamic analysis that suggests the conformational transition from the low-temperature conformation to the physiological-temperature conformation possesses a substantial unfolding quality. These results provide further evidence the structure of the enzyme obtained from crystals grown below 30 degrees C should not be regarded as the physiological structure (the normal bovine body temperature is 38.3 degrees C). These results should therefore have implications in any area that seeks to correlate the crystal structure of bovine carbonic anhydrase to physiological function.  相似文献   

11.
A scheme for the detoxification of superoxide in Pyrococcus furiosus has been previously proposed in which superoxide reductase (SOR) reduces (rather than dismutates) superoxide to hydrogen peroxide by using electrons from reduced rubredoxin (Rd). Rd is reduced with electrons from NAD(P)H by the enzyme NAD(P)H:rubredoxin oxidoreductase (NROR). The goal of the present work was to reconstitute this pathway in vitro using recombinant enzymes. While recombinant forms of SOR and Rd are available, the gene encoding P. furiosus NROR (PF1197) was found to be exceedingly toxic to Escherichia coli, and an active recombinant form (rNROR) was obtained via a fusion protein expression system, which produced an inactive form of NROR until cleavage. This allowed the complete pathway from NAD(P)H to the reduction of SOR via NROR and Rd to be reconstituted in vitro using recombinant proteins. rNROR is a 39.9-kDa protein whose sequence contains both flavin adenine dinucleotide (FAD)- and NAD(P)H-binding motifs, and it shares significant similarity with known and putative Rd-dependent oxidoreductases from several anaerobic bacteria, both mesophilic and hyperthermophilic. FAD was shown to be essential for activity in reconstitution assays and could not be replaced by flavin mononucleotide (FMN). The bound FAD has a midpoint potential of -173 mV at 23 degrees C (-193 mV at 80 degrees C). Like native NROR, the recombinant enzyme catalyzed the NADPH-dependent reduction of rubredoxin both at high (80 degrees C) and low (23 degrees C) temperatures, consistent with its proposed role in the superoxide reduction pathway. This is the first demonstration of in vitro superoxide reduction to hydrogen peroxide using NAD(P)H as the electron donor in an SOR-mediated pathway.  相似文献   

12.
A new non-heme iron protein from the periplasmic fraction of Desulfovibrio vulgaris (Hildenbourough NCIB 8303) has been purified to homogeneity, and its amino acid composition, molecular weight, redox potential, iron content, and optical, EPR, and M?ssbauer spectroscopic properties have been determined. This new protein is composed of two identical subunits with subunit molecular weight of 21,900 and contains four iron atoms per molecule. The as-purified oxidized protein exhibits an optical spectrum with absorption maxima at 492, 365, and 280 nm, and its EPR spectrum shows resonances at g = 4.3 and 9.4, characteristic of oxidized rubredoxin. The M?ssbauer data indicate the presence of approximately equal amounts of two types of iron; we named them the Rd-like and the Hr-like iron due to their similarity to the iron centers of rubredoxins (Rds) and hemerythrins (Hrs), respectively. For the Rd-like iron, the measured fine and hyperfine parameters (D = 1.5 cm-1, E/D = 0.26, delta EQ = -0.55 mm/s, delta = 0.27 mm/s, Axx/gn beta n = -16.5 T, Ayy/gn beta n = -15.6 T, and Azz/gn beta n = -17.0 T) are almost identical with those obtained for the rubredoxin from Clostridium pasteurianum. Redox-titration studies monitored by EPR, however, showed that these Rd-like centers have a midpoint redox potential of +230 +/- 10 mV, approximately 250 mV more positive than those reported for rubredoxins. Another unusual feature of this protein is the presence of the Hr-like iron atoms.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The kinetic and thermodynamic properties of ascorbate oxidase (AO) activity and stability of a Cucurbita maxima extract were investigated. Activity tests performed at 25 degrees C using initial ascorbic acid concentration in the range 50-750 M allowed estimating the Michaelis constant for this substrate (Km = 126 microM) and the maximum initial rate of ascorbic acid oxidation (A0,max = 1.57 mM min-1). The main thermodynamic parameters of the enzyme reaction (DeltaH* = 10.3 kJ mol-1; DeltaG* = 87.2 kJ mol-1; DeltaS* = -258 J mol-1 K-1) were estimated through activity tests performed at 25-48 C. Within such a temperature range, no decrease in the initial reaction rate was detected. The long-term thermostability of the raw extract was then investigated by means of residual activity tests carried out at 10-70 degrees C, which allowed estimating the thermodynamic parameters of the irreversible enzyme inactivation as well (DeltaH*D = 51.7 kJ mol-1; DeltaG*D = 103 kJ mol-1; S*D = -160 J mol-1 K-1). Taking into account the specific rate of AO inactivation determined at different temperatures, we also estimated the enzyme half-life (1047 min at 10 degrees C and 21.2 min at 70 degrees C) and predicted the integral activity of a continuous system using this enzyme preparation. This work should be considered as a preliminary attempt to characterize the AO activity of a C. maxima extract before its concentration by liquid-liquid extraction techniques.  相似文献   

14.
Manganese(II) activation of the methionyl aminopeptidases from Escherichia coli (EcMetAP-I) and the hyperthermophilic archaeon Pyrococcus furiosus (PfMetAP-II) was investigated. Maximum catalytic activity for both enzymes was obtained with 1 equiv of Mn(II), and the dissociation constants (K(d)) for the first metal binding site were found to be 6 +/- 0.5 and 1 +/- 0.5 microM for EcMetAP-I and PfMetAP-II, respectively. These K(d) values were verified by isothermal titration calorimetry (ITC) and found to be 3.0 +/- 0.2 and 1.4 +/- 0.2 microM for EcMetAP-I and PfMetAP-II, respectively. The hydrolysis of MGMM was measured in triplicate between 25 and 85 degrees C at eight substrate concentrations ranging from 2 to 20 mM for PfMetAP-II. Both specific activity and K(m) values increased with increasing temperature. An Arrhenius plot was constructed from the kcat values and was found to be linear over the temperature range 25-85 degrees C. The activation energy for the Mn(II)-loaded PfMetAP-II hydrolysis of MGMM was found to be 25.7 kJ/mol while the remaining thermodynamic parameters calculated at 25 degrees C are DeltaG+ = 50.1 kJ/mol, DeltaH+ = 23.2 kJ/mol, and DeltaS++ = -90.2 J x mol(-1) x K(-1).  相似文献   

15.
Kinetics for the breakdown of the trinuclear chromium acetate cluster, [Cr(3)O(OAc)(6)](+), with a series of monoprotic and diprotic ligands in weakly acidic aqueous media (pH approximately 4 or approximately 5) have been investigated spectrophotometrically at 40-60 degrees C. The results point to an ion-pair equilibrium as the first step followed by associative interchange mechanism forming the mononuclear product of the reaction. Pseudo-first-order rates were determined from absorbance data and associated activation parameters were calculated using the Eyring equation. Enthalpy and entropy terms of the reactions (e.g., histidine, DeltaH(double dagger) = 75 +/- 15 kJ mol(-1), DeltaS(double dagger) = -130 +/- 25 J K(-1) mol(-1); lactic acid, DeltaH(double dagger) = 66 +/- 13 kJ mol(-1), DeltaS(double dagger) = -155 +/- 30 J K(-1) mol(-1); glycine, DeltaH(double dagger) = 31 +/- 6 kJ mol(-1), DeltaS(double dagger) = -225 +/- 45 J K(-1) mol(-1)) are consistent with an associative interchange (I(a)) mechanism, and produce a linear isokinetic plot (slope = 50 degrees C). Rates and activation parameters are comparable to those of substitution reactions of the chromium(III) hexaaqua cation. Other ligands studied included malonic acid and the amino acid, aspartic acid. Observed rates are faster than water exchange rates, but typically slower than anion substitution rates, and indicate that trinuclear chromium(III) clusters are expected to be kinetically stable in neutral to slightly acidic conditions.  相似文献   

16.
Predicting the effects of mutation on the reduction potential of proteins is crucial in understanding how reduction potentials are modulated by the protein environment. Previously, we proposed that an alanine vs. a valine at residue 44 leads to a 50-mV difference in reduction potential found in homologous rubredoxins because of a shift in the polar backbone relative to the iron site due to the different side-chain sizes. Here, the aim is to determine the effects of mutations to glycine, isoleucine, and leucine at residue 44 on the structure and reduction potential of rubredoxin, and if the effects are proportional to side-chain size. Crystal structure analysis, molecular mechanics simulations, and experimental reduction potentials of wild-type and mutant Clostridium pasteurianum rubredoxin, along with sequence analysis of homologous rubredoxins, indicate that the backbone position relative to the redox site as well as solvent penetration near the redox site are both structural determinants of the reduction potential, although not proportionally to side-chain size. Thus, protein interactions are too complex to be predicted by simple relationships, indicating the utility of molecular mechanics methods in understanding them.  相似文献   

17.
C Saez  N M Thielens  E S Bjes  A F Esser 《Biochemistry》1999,38(21):6807-6816
The association of terminal complement proteins was investigated by analytical ultracentrifugation and multi-angle laser light scattering. Native C8 and C9 formed a heterodimer in solution of physiological ionic strength with a free-energy change DeltaG degrees of -8.3 kcal/mol and a dissociation constant Kd of 0.6 microM (at 20 degrees C) that was ionic strength- and temperature-dependent. A van't Hoff plot of the change in Kd was linear between 10 and 37 degrees C and yielded values of DeltaH degrees = -12.9 kcal/mol and DeltaS degrees = -15.9 cal mol-1 deg-1, suggesting that electrostatic forces play a prominent role in the interaction of C8 with C9. Native C8 also formed a heterodimer with C5, and low concentrations of polyionic ligands such as protamine and suramin inhibited the interaction. Suramin induced high-affinity trimerization of C8 (Kd = 0.10 microM at 20 degrees C) and dimerization of C9 (Kd = 0.86 microM at 20 degrees C). Suramin-induced C8 oligomerization may be the primary reason for the drug's ability to prevent complement-mediated hemolysis. Analysis of sedimentation equilibria and also of the fluorescence enhancement of suramin when bound to protein provided evidence for two suramin-binding sites on each C9 and three on each C8 in the oligomers. Oligomerization could be reversed by high suramin concentrations, but 8-aminonaphthalene-1,3,6- trisulfonate (ANTS2- ), which mimics half a suramin molecule, could not compete with suramin binding and oligomerization suggesting that the drug also binds nonionically to the proteins.  相似文献   

18.
Binding constant data K degrees (T) are commonly subjected to van't Hoff analysis to extract estimates of DeltaH degrees, DeltaS degrees, and DeltaCP degrees for the process in question. When such analyses employ unweighted least-squares fitting of lnK degrees to an appropriate function of the temperature T, they are tacitly assuming constant relative error in K degrees. When this assumption is correct, the statistical errors in DeltaG degrees, DeltaH degrees, DeltaS degrees, DeltaCP degrees, and the T-derivative of DeltaCP degrees (if determined) are all independent of the actual values of K degrees and can be computed from knowledge of just the T values at which K degrees is known and the percent error in K degrees. All of these statistical errors except that for the highest-order constant are functions of T, so they must normally be calculated using a form of the error propagation equation that is not widely known. However, this computation can be bypassed by defining DeltaH degrees as a polynomial in (T-T0), the coefficients of which thus become DeltaH degrees, DeltaCP degrees, and 1/2 dDeltaCP degrees/dT at T=T0. The errors in the key quantities can then be computed by just repeating the fit for different T0. Procedures for doing this are described for a representative data analysis program. Results of such calculations show that expanding the T range from 10-40 to 5-45 degrees C gives significant improvement in the precision of all quantities. DeltaG degrees is typically determined with standard error a factor of approximately 30 smaller than that for DeltaH degrees. Accordingly, the error in TDeltaS degrees is nearly identical to that in DeltaH degrees. For 4% error in K degrees, the T-derivative in DeltaCP degrees cannot be determined unless it is approximately 10 cal mol-1 K-2 or greater; and DeltaCP degrees must be approximately 50 cal mol-1 K-1. Since all errors scale with the data error and inversely with the square root of the number of data points, the present results for 4% error cover any other relative error and number of points, for the same approximate T structure of the data.  相似文献   

19.
LeMaster DM  Tang J  Hernández G 《Proteins》2004,57(1):118-127
The striking kinetic stability of many proteins derived from hyperthermophilic organisms has led to the proposal that such stability may result from a heightened activation barrier for unfolding independent of a corresponding increase in the thermodynamic stability. This in turn implies a corresponding retardation of the folding reaction. A commonly cited model for kinetic thermal stabilization is the rubredoxin from Pyrococcus furiosus (Pf), which exhibits an irreversible denaturation lifetime at 100 degrees C of nearly a week. Utilizing protein resonances shifted well outside of the random coil chemical shift envelope, nuclear magnetic resonance (NMR) chemical exchange measurements on Pf rubredoxin as well as on the mesophile Clostridium pasteurianum (Cp) rubredoxin demonstrate reversible thermal transition temperatures of 144 degrees C (137 degrees C for the N-terminal modified A2K variant) and 104 degrees C, respectively, with similar (un)folding rates of approximately 25,000 s(-1), only modestly slower than the diffusion controlled rate. The absence of a substantial activation barrier to rubredoxin folding as well as the similar folding kinetics of the mesophile protein indicate that kinetic stabilization has not been utilized by the hyperthermophile rubredoxin in achieving its extreme thermal stability. The two-state folding kinetics observed for Pf rubredoxin contradict a previous assertion of multiphasic folding based on hydrogen exchange data extrapolated to an estimated midpoint of transition temperature (T(m)) of nearly 200 degrees C. This discrepancy is resolved by the observation that the base-catalyzed hydrogen exchange of the model dipeptide (N-acetyl-L-cysteine-N-methylamide)4-Cd2+ is 23-fold slower than that of the free cysteine model dipeptide used to normalize the Pf rubredoxin hydrogen exchange data.  相似文献   

20.
NAD(P)H:rubredoxin oxidoreductase (NROR) has been purified from the hyperthermophilic archaeon Pyrococcus furiosus. The enzyme is exceedingly active in catalyzing the NADPH-dependent reduction of rubredoxin, a small (5.3-kDa) iron-containing redox protein that had previously been purified from this organism. The apparent Vmax at 80 degrees C is 20,000 micromol/min/mg, which corresponds to a kcat/Km value of 300,000 mM(-1) s(-1). The apparent Km values measured at 80 degrees C and pH 8.0 for rubredoxin, NADPH, and NADH were 50, 5, and 34 microM, respectively. The enzyme did not reduce P. furiosus ferredoxin. NROR is a monomer with a molecular mass of 45 kDa and contains one flavin adenine dinucleotide molecule per mole but lacks metals and inorganic sulfide. The possible physiological role of this hyperactive enzyme is discussed.  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