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.
Guanidine hydrochloride-induced denaturation and thermal denaturation of three kinds of tryptophan synthase alpha subunit have been compared by circular dichroism measurements. The three alpha subunits are from Escherichia coli, Salmonella typhimurium, and an interspecies hybrid in which the C-terminal domain comes from E. coli (alpha-2 domain) and the N-terminal domain comes from S. typhimurium (alpha-1 domain). Analysis of denaturation by guanidine hydrochloride at 25 degrees C showed that the alpha-2 domain of S. typhimurium was more stable than the alpha-2 domain of E. coli, but the alpha-1 domain of S. typhimurium was less stable than the alpha-1 domain of the E. coli protein; overall, the hybrid protein was slightly less stable than the two original proteins. It is concluded that the stability to guanidine hydrochloride denaturation of each of the domains of the interspecies hybrid is similar to the stability of the domain of the species from which it originated. The E. coli protein was more stable to thermal denaturation than the other proteins near the denaturation temperature, but the order of their thermal stability was reversed at 25 degrees C and coincided with that obtained from guanidine hydrochloride-induced denaturation.  相似文献   

2.
Hemoglobins A and S were crosslinked between Lys 82 beta 1 and Lys 82 beta 2 using bis (3,5-dibromosalicyl) fumarate (J. A. Walder et al. (1979) Biochemistry 18, 4265). Thermal denaturation experiments were used to compare the stabilities of the met, cyanomet, and carbonmonoxy forms of these crosslinked hemoglobins to the corresponding uncrosslinked proteins. Uncrosslinked carbonmonoxy- and cyanomethemoglobins had transition temperatures about 11 degrees C higher than the corresponding met samples. The increase in denaturation temperature (Tm) due to crosslinking was 15 degrees C for the methemoglobins, 10 degrees C for the cyanomethemoglobins, and 4 degrees C for the carbonmonoxy ones. There was no significant difference in stability between the met and carbonmonoxy crosslinked proteins. In order of increasing stability the samples were: met Hb S less than met Hb A less than CO Hb S less than CO Hb A = CN-met Hb A less than met XL-Hb S = CO XL-Hb S less than met XL-Hb A = CO XL-Hb A less than CN-met XL-Hb A. The slight decrease in the stability of Hb S (beta 6 Glu----Val) compared to Hb A can be explained by the replacement of an external ionic group by a hydrophobic residue in Hb S. In mixtures of crosslinked and normal Hb A, the Tm of the uncrosslinked material was slightly increased by the presence of the more stable crosslinked hemoglobin. The effects of both crosslinking and cyanide or carbon monoxide binding can be explained by Le Chatelier's principle since both would favor the native form of the protein.  相似文献   

3.
Guanidine hydrochloride-induced denaturation and thermal denaturation of three kinds of tryptophan synthase α subunit have been compared by circular dichroism measurements. The three α subunits are from Escherichia coli, Salmonella typhimurium, and an interspecies hybrid in which the C-terminal domain comes from E. coli (α-2 domain) and the N-terminal domain comes from S. typhimurium (α-1 domain). Analysis of denaturation by guanidine hydrochloride at 25 °C showed that the α-2 domain of S. typhimurium was more stable than the α-2 domain of E. coli, but the α-1 domain of S. typhimurium was less stable than the α-1 domain of the E. coli protein; overall, the hybrid protein was slightly less stable than the two original proteins. It is concluded that the stability to guanidine hydrochloride denaturation of each of the domains of the interspecies hybrid is similar to the stability of the domain of the species from which it originated. The E. coli protein was more stable to thermal denaturation than the other proteins near the denaturation temperature, but the order of their thermal stability was reversed at 25 °C and coincided with that obtained from guanidine hydrochloride-induced denaturation.  相似文献   

4.
M Burke  S Zaager  J Bliss 《Biochemistry》1987,26(5):1492-1496
The stability of myosin subfragment 1 (S1) to thermal denaturation has been followed by limited tryptic proteolysis. Digestions done during the thermal denaturation show that at temperatures at and above 37 degrees C there is a marked increase in the susceptibility of S1 to tryptic degradation, as evidenced by the loss of all bands corresponding to the normally trypsin-resistant fragments of 50, 27, and 21 kDa of the heavy chain and to the light chain. The enhanced digestion of S1 appears to be due to a general unfolding of all segments of S1, although the 50-kDa segment appears to unfold at a lower temperature than the remainder of the S1 structure. Digestions done after 30-min exposure to higher temperatures or after subsequent cooling to 25 degrees C show marked differences in the susceptibility of the S1 to trypsin. This suggests that, on cooling, a substantial portion of the S1, but not the 50-kDa segment, is capable of refolding to a state corresponding closely to that in the native S1. These data indicate that in terms of thermal denaturation the S1 behaves as though it is comprised of two domains--an unstable 50-kDa domain and a more stable domain comprised of the 27- and 21-kDa segments of the heavy chain interacting with the light chain, as proposed recently by Setton and Muhlrad [Setton, A., & Muhlrad, A. (1984) Arch. Biochem. Biophys. 235, 411-417]. The rates of thermal inactivation of the ATPase of S1 are found to correspond closely to the decay rates for the 50-kDa fragment, suggesting that this segment in S1 is closely associated with the ATPase function of the protein.  相似文献   

5.
Azurin has a beta-barrel fold comprising eight beta-strands and one alpha helix. A disulfide bond between residues 3 and 26 connects the N-termini of beta strands beta1 and beta3. Three mutant proteins lacking the disulfide bond were constructed, C3A/C26A, C3A/C26I and a putative salt bridge (SB) in the C3A/S25R/C26A/K27R mutant. All three mutants exhibit spectroscopic properties similar to the wild-type protein. Furthermore, the crystal structure of the C3A/C26A mutant was determined at 2.0 A resolution and, in comparison to the wild-type protein, the only differences are found in the immediate proximity of the mutation. The mutants lose the 628 nm charge-transfer band at a temperature 10-22 degrees C lower than the wild-type protein. The folding of the zinc loaded C3A/C26A mutant was studied by guanidine hydrochloride (GdnHCl) induced denaturation monitored both by fluorescence and CD spectroscopy. The midpoint in the folding equilibrium, at 1.3 M GdnHCl, was observed using both CD and fluorescence spectroscopy. The free energy of folding determined from CD is -24.9 kJ.mol-1, a destabilization of approximately 20 kJ.mol-1 compared to the wild-type Zn2+-protein carrying an intact disulfide bond, indicating that the disulfide bond is important for giving azurin its stable structure. The C3A/C26I mutant is more stable and the SB mutant is less stable than C3A/C26A, both in terms of folding energy and thermal denaturation. The folding intermediate of the wild-type Zn2+-azurin is not observed for the disulfide-deficient C3A/C26A mutant. The rate of unfolding for the C3A/C26A mutant is similar to that of the wild-type protein, suggesting that the site of the mutation is not involved in an early unfolding reaction.  相似文献   

6.
The contribution of His64 to the function and stability of tissue plasminogen activator (t-PA) kringle-2 domain (His244 in t-PA numbering) has been studied by using microcalorimetric methods to compare the ligand binding and thermal denaturation behavior of wild-type kringle-2 and mutants having His64 replaced with Tyr or Phe. This site was examined because modeling studies suggested that the His64 side chain could play an important role in ligand binding by forming an ion-pair with the carboxylate of the ligand, L-lysine. Kringle-2 domains were expressed by secretion of the 174-263 portion of t-PA in E. coli and purified as previously described for the wild-type domain. Both mutant proteins retain affinity for L-lysine, although reduced three- to four-fold relative to wild-type, demonstrating that His64 does not interact with the ligand carboxylate through an ion-pair interaction or by hydrogen bonding. The H64Y substitution does result in an altered specificity of the lysine binding site with the mutant domain having greatest affinity for a ligand of 6.8 A chain length, whereas the wild-type domain prefers an 8.8 A long ligand. For both wild-type and mutant, the binding of the optimal chain length ligand is dominated by enthalpic effects (delta H = -6,000 to -7,000 cal/mol) and T delta S accounts for less than 15% of delta G. In addition, the H64Y mutant differs from wild-type in the effect of ligand alpha-amino group modification on binding affinity. Based on examination of the x-ray structure recently determined for wild-type kringle-2, the specificity changes accompanying the H64Y substitution probably result from changes in side chain interactions in the lysine binding site. Thermal denaturation experiments show that the H64Y mutant is also more stable than the wild-type protein with the difference in stabilization free energy (delta delta G) equal to 2.7 kcal/mol at 25 degrees C and pH 3. The increased stability of the mutant appears to be related to the difference in hydrophobicity between His and Tyr.  相似文献   

7.
The residue-specific urea-induced unfolding patterns of recombinant prion proteins from different species (bovine, rabbit, mouse, and Syrian hamster) were monitored using high-resolution (1)H nuclear magnetic resonance (NMR) spectroscopy. Protein constructs of different lengths, and with and without a His tag attached at the N-terminus, were studied. The various species showed different overall sensitivities toward urea denaturation with stabilities in the following order: hamster ≤ mouse < rabbit < bovine protein. This order is in agreement with recent circular dichroism (CD) spectroscopic measurements for several species [Khan, M. Q. (2010) Proc. Natl. Acad. Sci. U.S.A.107, 19808-19813] and for the bovine protein presented herein. The [urea](1/2) values determined by CD spectroscopy parallel those of the most stable residues observed by NMR spectroscopy. Neither the longer constructs containing an additional hydrophobic region nor the His tag influenced the stability of the structured domain of the constructs studied. The effect of the S174N mutation in rabbit PrP(C) was also investigated. The rank order of the regional stabilities within each protein remained the same for all species. In particular, the residues in the β-sheet region in all four species were more sensitive to urea-induced unfolding than residues in the α2 and α3 helical regions. These observations indicate that the regional specific unfolding pattern is the same for the four mammalian prion proteins studied but militate against the idea that PrP(Sc) formation is linked with the global stability of PrP(C).  相似文献   

8.
S C Larson  G W Fisher  N T Ho  T J Shen  C Ho 《Biochemistry》1999,38(29):9549-9555
Three recombinant mutants of human fetal hemoglobin (Hb F) have been constructed to determine what effects specific amino acid residues in the gamma chain have on the biophysical and biochemical properties of the native protein molecule. Target residues in these recombinant fetal hemoglobins were replaced with the corresponding amino acids in the beta chain of human normal adult hemoglobin (Hb A). The recombinant mutants of Hb F included rHb F (gamma 112Thr --> Cys), rHb F (gamma 130Trp --> Tyr), and rHb F (gamma 112Thr --> Cys/gamma 130Trp --> Tyr). Specifically, the importance of gamma 112Thr and gamma 130Trp to the stability of Hb F against alkaline denaturation and in the interaction with sickle cell hemoglobin (Hb S) was investigated. Contrary to expectations, these rHbs were found to be as stable against alkaline denaturation as Hb F, suggesting that the amino acid residues mentioned above are not responsible for the stability of Hb F against the alkaline denaturation as compared to that of Hb A. Sub-zero isoelectric focusing (IEF) was employed to investigate the extent of hybrid formation in equilibrium mixtures of Hb S with these hemoglobins and with several other hemoglobins in the carbon monoxy form. Equimolar mixtures of Hb A and Hb S and of Hb A(2) and Hb S indicate that 48-49% of the Hb exists as the hybrid tetramer, which is in agreement with the expected binomial distribution. Similar mixtures of Hb F and Hb S contain only 44% hybrid tetramer. The results for two of our recombinant mutants of Hb F were identical to the results for mixtures of Hb F and Hb S, while the other mutant, rHb F (gamma 130Trp --> Tyr), produced 42% hybrid tetramer. The sub-zero IEF technique discussed here is more convenient than room-temperature IEF techniques, which require Hb mixtures in the deoxy state. These recombinant mutants of Hb F were further characterized by equilibrium oxygen binding studies, which indicated no significant differences from Hb F. While these mutants of Hb F did not have tetramer-dimer dissociation properties significantly altered from those of Hb F, future mutants of Hb F may yet prove useful to the development of a gene therapy for the treatment of patients with sickle cell anemia.  相似文献   

9.
《FEBS letters》1986,208(2):373-378
Incubation of 60 S ribosomal subunits with the ricin A chain reduced their stability during heat treatment. The toxin shifted the thermal denaturation curve of the subunits towards lower temperatures, in a similar way to that produced by the decrease in Mg2+ concentration. A brief heating (3 min at 57°C), which did not affect control subunit activity, enhanced protein synthesis inhibition of the toxin-treated subunits that released more 5 S RNA, in the form of nucleoprotein complex(es) with protein L5 and phosphoproteins P1P2 (RNPH), than did heated control subunits [(1984) Eur. J. Biochem, 143, 303-307]. No nuclease activity tested on 60 S subunits and purified 5 S and 5.8 S RNA was found associated with the toxin. The results suggest that the toxin induced a limited conformational change of the 60 S subunit, which destabilized the interaction between RNPH and the rest of the subunit.  相似文献   

10.
Introduction of Pro residues into helix interiors results in protein destabilization. It is currently unclear if the converse substitution (i.e., replacement of Pro residues that naturally occur in helix interiors would be stabilizing). Maltose-binding protein is a large 370-amino acid protein that contains 21 Pro residues. Of these, three nonconserved residues (P48, P133, and P159) occur at helix interiors. Each of the residues was replaced with Ala and Ser. Stabilities were characterized by differential scanning calorimetry (DSC) as a function of pH and by isothermal urea denaturation studies as a function of temperature. The P48S and P48A mutants were found to be marginally more stable than the wild-type protein. In the pH range of 5-9, there is an average increase in T(m) values of P48A and P48S of 0.4 degrees C and 0.2 degrees C, respectively, relative to the wild-type protein. The other mutants are less stable than the wild type. Analysis of the effects of such Pro substitutions in MBP and in three other proteins studied to date suggests that substitutions are more likely to be stabilizing if the carbonyl group i-3 or i-4 to the mutation site is not hydrogen bonded in the wild-type protein.  相似文献   

11.
The conformation of porcine serum ferric transferrin (Tf) and its stability against denaturation were studied by circular dichroism. Tf was estimated to have 19-24% alpha-helix and 50-55% beta-sheet based on the methods of Chang et al. (Chang, C.T., Wu, C.-S.C., & Yang, J.T., 1978, Anal. Biochem. 91, 13-31) and Provencher and Glöckner (Provencher, S.W. & Glöckner, J., 1981, Biochemistry 20, 33-37). Removal of the bound ferric ions (apo-Tf) did not alter the overall conformation, but there were subtle changes in local conformation based on its near-UV CD spectrum. The Tfs were stable between pH 3.5 and 11. Denaturation by guanidine hydrochloride (Gu-HCl) showed two transitions at 1.6 and 3.4 M denaturant. The process of denaturation by acid and base was reversible, whereas that by Gu-HCl was partially reversible. The irreversible thermal unfolding of Tfs began at temperatures above 60 degrees C and was not complete even at 80 degrees C. The bound irons (based on absorbance at 460 nm) were completely released at pH < 4 or in Gu-HCl solution above 1.7 M, when the protein began to unfold, but they remained intact in neutral solution even at 85 degrees C. The NH2- and COOH-terminal halves of the Tf molecule obtained by limited trypsin digestion had CD spectra similar to the spectrum of native Tf, and the COOH-terminal fragment had more stable secondary structure than the NH2-terminal fragment.  相似文献   

12.
Meinhold D  Beach M  Shao Y  Osuna R  Colón W 《Biochemistry》2006,45(32):9767-9777
Two crossed-linked variants of the homodimeric DNA binding protein factor for inversion stimulation (FIS) were created via engineering of single intermolecular disulfide bonds. The conservative S30C and the nonconservative V58C FIS independent mutations resulted in FIS crossed-linked at the A helix (C30-C30) and at the middle of the B helix (C58-C58). This study sought to investigate how the location of an intermolecular disulfide bond may determine the effect on stability and its propagation through the structure to preserve or alter the denaturation cooperativity of FIS. The oxidized and reduced S30C and V58C FIS exhibited a far-UV CD spectrum and DNA binding affinities that were similar to WT FIS, indicating no significant changes in secondary and tertiary structure. However, the reduced and oxidized forms of the mutants revealed significant differences in the stability and equilibrium denaturation mechanism between the two mutants. In the reduced state, S30C FIS had very little effect on FIS stability, whereas V58C FIS was 2-3 kcal/mol less stable than WT FIS. Interestingly, while both disulfide bonds significantly increased the resistance to urea- and guanidine hydrochloride (GuHCl)-induced denaturation, oxidized V58C FIS exhibited a three-state GuHCl-induced transition. In contrast, oxidized S30C FIS displayed a highly cooperative WT-like transition with both denaturants. The three-state denaturation mechanism of oxidized V58C FIS induced by the GuHCl salt was reproduced by urea denaturation at pH 4, suggesting that disruption of a C-terminus salt-bridge network is responsible for the loss of denaturation cooperativity of V58C FIS in GuHCl or urea, pH 4. A second mutation on V58C FIS created to place a single tryptophan probe (Y95W) at the C-terminus further implies that the denaturation intermediate observed in disulfide crossed-linked V58C FIS results from a decoupling of the stabilities of the C-terminus and the rest of the protein. These results show that, unlike the C30-C30 intermolecular disulfide bond, the C58-C58 disulfide bond did not evenly stabilize the FIS structure, thereby highlighting the importance of the location of an engineered disulfide bond on the propagation of stability and the denaturation cooperativity of a protein.  相似文献   

13.
We have carried out guanidinium chloride (GdmCl) and urea denaturations of bovine beta-lactoglobulin A (beta-lgA) at pH 2.0 and 25 degrees C, using far-UV and near-UV circular dichroism, near-UV absorption and tryptophan fluorescence spectroscopies. The stable intermediate state that occurs during GdmCl denaturation has been characterized by the far- and near-UV circular dichroism, tryptophan difference absorption, tryptophan fluorescence and 8-anilino-1-naphthalene sulphonic acid binding measurements. Following conclusions have been reached. (a) Urea-induced denaturation is not a two-state process. (b) GdmCl-induced denaturation is composed of two distinct two-state processes. (c) alpha-Helical content, burial of tryptophan residues and burial of hydrophobic surface area are more in the GdmCl-induced stable intermediate than those originally present in the native protein.  相似文献   

14.
The acid-induced denaturation of wild-type Staphylococcal nuclease (WT) and its eight mutant forms, L25A, V66L, G79S, A90S, G88V, H124L, V66L/G88V, and V66L/G79S/G88V, was investigated using Trp140 fluorescence as a probe at 30 degrees C. The values of pH(1/2), at which the denaturation is half completed, and n, the apparent number of protons which trigger the denaturation and are taken up by the proteins upon denaturation at pH(1/2), were evaluated from the pH dependence of the fluorescence intensity. The values of pH(1/2) and n for WT were 3.8 and 1.8 respectively. The amino acid replacements changed the pH(1/2) values to a range between 3.0 (H124L) and 4.4 (G79S) and also changed the n values to a range between 1.0 (A90S) and 3.0 (G88V). There was a negative correlation between the values of pH(1/2) and n. It was suggested that the amino acid replacements may change the energy levels of the native state and/or the denatured state mainly in the neutral (stable) pH region, not in the acidic (unstable) region, resulting in the correlative changes in pH(1/2) and n.  相似文献   

15.
The effect of sodium dodecyl sulfate (SDS) and Triton X-100 on the hydrolytic activity of lipases A and B from Candida rugosa has been studied. Lipase B is significantly more affected than lipase A by the presence of both surfactants; Triton X-100 produces a more deleterious effect than SDS with both isoenzymes. In addition, the stability of lipases A and B in the presence of different concentrations of SDS was investigated; lipase A was more stable than isoform B. Both isoenzymes were chemically modified by reaction of their amino groups with octanoyl chloride or activated polyethylene glycol (PEG, mol. wt. 5000). In all cases the modification produced a protective effect against denaturation by SDS. In particular, PEG5000-liPases A and B were significantly more stable (stabilization factor: 3-4) than the native enzymes at the surfactant concentrations tested.  相似文献   

16.
Hyperchromicity measurements and quasi-elastic laser light scattering (QELS) have been used to assess the solution structure of the metabolically stable E. coli 4.5S RNA. Results from thermal denaturation measurements revealed the 4.5S species to be markedly more stable than most other RNAs characterized thus far. Optical Tm's range from 79 degrees to 88 degrees with transitions approximately 25 degrees C wide. The Tm values show little dependence on ionic strength, but stability is enhanced considerably by Mg+2. In the QELS experiments the diffusion coefficient does not decrease until T greater than 70 degrees C. Neither the diffusive melting nor the diffusion coefficient at infinite dilution (D0(20,w)) show dependencies on ionic strength but both are influenced by Mg+2. The diffusion behavior is in agreement with that predicted for a rigid cylindrical molecule 125 to 160 A long and 37 to 26 A in diameter. Taken together these results are consistent with the more stable hypothetical secondary structures that can be formed, in which 70-75% of the 114 bases are paired to form a single extended hairpin helix.  相似文献   

17.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

18.
The solution conformation of alpha 1-antitrypsin from human blood plasma was studied by the circular dichroism (CD) probe. The CD spectra revealed in this glycoprotein approximately 16-20% of alpha-helix, the rest of the main polypeptide chain possessing the pleated sheet (beta) and the aperiodic structures. The conformation was stable between pH 4.7 and 8.8. Reversible change in conformation was observed at pH 10.3, and more dratic denaturation occurred at pH 11.6. The environment of the side chain chromophores was strongly affected by acid at pH 2.5, whereas the main chain conformation was changed slightly. A drastic change in the CD spectra, indicating denaturation, was observed in 3.5 M guanidine hydrochloride. Sodium dodecyl sulfate was effective in disorganizing the tertiary structure and in enhancing the helix content. The phenylalanine band fine structure was observed in the native protein and also after denaturation with acid, guanidine hydrochloride and sodium dodecyl sulfate.  相似文献   

19.
We have used alanine-scanning site-directed mutagenesis of the dimer contact region of starch phosphorylase from Corynebacterium callunae to explore the relationship between a protein conformational change induced by phosphate binding and the up to 500-fold kinetic stabilization of the functional quarternary structure of this enzyme when phosphate is present. Purified mutants (at positions Ser-224, Arg-226, Arg-234, and Arg-242) were characterized by Fourier transform-infrared (FT-IR) spectroscopy and enzyme activity measurements at room temperature and under conditions of thermal denaturation. Difference FT-IR spectra of wild type and mutants in (2)H(2)O solvent revealed small changes in residual amide II band intensities at approximately 1,550 cm(-1), indicating that (1)H/(2)H exchange in the wild type is clearly perturbed by the mutations. Decreased (1)H/(2)H exchange in comparison to wild type suggests formation of a more compact protein structure in S224A, R234A, and R242A mutants and correlates with rates of irreversible thermal denaturation at 45 degrees C that are up to 10-fold smaller for the three mutants than the wild type. By contrast, the mutant R226A inactivates 2.5-fold faster at 45 degrees C and shows a higher (1)H/(2)H exchange than the wild type. Phosphate (20 mM) causes a greater change in FT-IR spectra of the wild type than in those of S224A and 234A mutants and leads to a 5-fold higher stabilization of the wild type than the two mutants. Therefore, structural effects of phosphate binding leading to kinetic stability of wild-type starch phosphorylase are partially complemented in the S224A and R234A mutants. Infrared spectroscopic measurements were used to compare thermal denaturations of the mutants and the wild type in the absence and presence of stabilizing oxyanion. The broad denaturation transition of unliganded wild type in the range 40-50 degrees C is reduced in the S224A and R234A mutants, and this reflects mainly a shift of the onset of denaturation to a 4-5 degrees C higher value.  相似文献   

20.
The reaction of oxyhaemoglobin and acetylphenylhydrazine, which results in haemoglobin denaturation and precipitation, was found to be influenced by H202 and superoxide (O2-.) generated during the reaction. By analysing the different haemoglobin oxidation products, it was found that by influencing the rate at which oxyhaemoglobin was oxidized, H2O2 accelerated the overall haemoglobin breakdown, and O2-. inhibited it. By adding GSH (reduced glutathione) or ascorbate, it was possible to slow down the rates of both oxyhaemoglobin oxidation and O2-. production, and the overall rate of haemoglobin breakdown. These results are compatible with a mechanism involving production of the acetylphenylhydrazyl free radical, and with GSH, ascorbate and O2-. acting as radical scavengers and preventing its further reactions. The reaction produced choleglobin, as well as acetylphenyldiazine and methaemoglobin, which combined to form a haemichrome. The haemichrome was less stable and precipitated first. It was also less stable than the haemichrome formed by direct reaction of acetylphenyldiazine with methaemoglobin, and it is proposed that this is because the methaemoglobin produced from oxyhaemoglobin and acetylphenylhydrazine was modified by the free radicals and H2O2 produced in the reaction.  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