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.
TspMI, a thermostable isoschizomer of XmaI from a Thermus sp., has been characterized. The enzyme was purified to homogeneity using Cibacron-Blue 3GA agarose, Heparin agarose, SP sephadex C50, and Mono-Q fast protein liquid chromatography and was found to be a homodimer of 40 kDa. Restriction mapping and run-off sequencing of TspMI-cleaved DNA ends depicted that it cleaved at 5′C/CCGGG3′ to generate a four-base, 5′-CCGG overhang. The enzyme was sensitive to methylation of second and third cytosines in its recognition sequence. TspMI worked optimally at 60°C with 6 mM Mg2+, no Na+/K+, and showed no star activity in the presence of 25% glycerol. The enzyme could efficiently digest the DNA labeled with a higher concentration of YOYO-I (one dye molecule to one nucleotide), making it a useful candidate for real-time imaging experiments. Single molecule interaction between TspMI and λ DNA was studied using total internal reflection fluorescence microscopy. The enzyme survived 30 polymerase chain reaction (PCR) cycles in the presence of 10% glycerol and 0.5 M trehalose without any activity loss and, hence, is suitable for incorporation in restriction-endonuclease-mediated selective-PCR for various applications.Electronic Supplementary Material Supplementary material for this article is available at  相似文献   

2.
3.
The structural requirements of the NADP+ molecule as a coenzyme in the oxidative decarboxylation reaction catalysed by pigeon liver malic enzyme were studied by kinetic and fluorimetric analyses with various NADP+ analogues and fragments. The substrate L-malate had little effect on the nucleotide binding. Etheno-NADP+, 3-acetylpyridine-adenine dinucleotide phosphate, and nicotinamide-hypoxanthine dinucleotide phosphate act as alternative coenzymes for the enzyme. Their kinetic parameters were similar to that of NADP+. Thionicotinamide-adenine dinucleotide phosphate, 3-aminopyridine-adenine dinucleotide phosphate, 5'-adenylyl imidodiphosphate, nicotinamide-adenine dinucleotide 3'-phosphate and NAD+ act as inhibitors for the enzyme. The first two were competitive with respect to NADP+ and non-competitive with respect to L-malate; the other inhibitors were non-competitive with NADP+. All NADP+ fragments were inhibitory to the enzyme, with a wide range of affinity, depending on the presence or absence of a 2'-phosphate group. Compounds with this group bind to the enzyme 2-3 orders of magnitude more tightly than those without this group. Only compounds with this group were competitive inhibitors with respect to NADP+. We conclude that the 2'-phosphate group is crucial for the nucleotide binding of this enzyme, whereas the carboxyamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity. There is a strong synergistic effect between the binding of the nicotinamide and adenosine moieties of the nucleotide molecule.  相似文献   

4.
Time profiles for degradation of DNA via reaction of H2O2 with the DNA-Cu+ complex were analyzed over a wide range of concentrations of the components. The yield of DNA damage per H2O2 molecule is 10 times lower than that obtained with gamma-radiolytically generated .OH radicals. The observations can be explained by a model in which H2O2 reacts, slowly on the one hand with DNA-Cu+ by formation of toxic .OH radicals immediately at the DNA and faster on the other hand with Cu+ in the bulk solution by formation of less toxic Cu(III) intermediates.  相似文献   

5.
Employing "phosphorylating" submitochondrial particles as the source of pyridine nucleotide transhydrogenase, the occurrence of an energy-linked NADH----NADP+ transhydrogenation in the adult cestode Hymenolepis diminuta was demonstrated. The isolated particles displayed rotenone-sensitive NADH utilization and the reversible transhydrogenase, with the NADPH----NAD+ transhydrogenation being more prominent. Although not inhibiting the NADPH----NAD+ reaction, rotenone, but not oligomycin, inhibited the catalysis of NADH----NADP+ transhydrogenation. In the presence of rotenone, Mg2+ plus ATP stimulated by more than 3-fold NADH----NADP+ transhydrogenation. This stimulation was ATP specific and was abolished by EDTA or oligomycin. Succinate was essentially without effect on the NADH----NADP+ reaction. These data demonstrate the occurrence of an energy-linked transhydrogenation between NADH and NADP+ with energization resulting from either electron transport-dependent NADH oxidation or ATP utilization via the phosphorylating mechanism in accord with the preparation of "phosphorylating" particles. This is the first demonstration of an energy-linked transhydrogenation in the parasitic helminths and apparently in the invertebrates generally.  相似文献   

6.
Free fatty acid (FFA), monoacylglycerol (MG) and diacylglycerol (DG) in high-FFA rice bran oil were continuously converted with glycerol (G) to form triacylglycerol (TG), using lipase fromRhizomucor miehei immobilized on anion-exchange resin. The reaction was continued for more than 1 month by a reactor with two circulation loops, each being connected to a fixed-bed reactor and a dehydrator. The reaction of 2 FFA + G DG + 2H2O appeared to occur until the glycerol was exhausted; the reaction of FFA + DG TG + H2O then followed. The consecutive esterificaion continued in the presence of 2–8 ppm water and the TG content reached 74%–88%. The industrial feasibility of this process was assessed from the standpoints of enzyme cost and value added by esterification.  相似文献   

7.
Dictyostelium alpha-actinin is a Ca(2+)-regulated F-actin cross-linking protein. To test the inhibitory function of the two EF hands, point mutations were introduced into either one or both Ca(2+)-binding sites. After mutations, the two EF hands were distinguishable with respect to their regulatory activities. Inactivation of EF hand I abolished completely the F-actin cross-linking activity of Dictyostelium discoideum alpha-actinin but Ca2+ binding by EF hand II was still observed in a 45Ca2+ overlay assay. In contrast, after mutation of EF hand II the molecule was still active and inhibited by Ca2+; however, approximately 500-fold more Ca2+ was necessary for inhibition and 45Ca2+ binding could not be detected in the overlay assay. These data indicate that EF hand I has a low affinity for Ca2+ and EF hand II a high affinity, implying a regulatory function of EF hand I in the inhibition of F-actin cross-linking activity. Biochemical data is presented which allows us to distinguish two functions of the EF hand domains in D. discoideum alpha-actinin: (a) at the level of the EF- hands, the Ca(2+)-binding affinity of EF hand I was increased by EF hand II in a cooperative manner, and (b) at the level of the two subunits, the EF hands acted as an on/off switch for actin-binding in the neighboring subunit. To corroborate in vitro observations in an in vivo system we tried to rescue the abnormal phenotype of a mutant (Witke, W., M. Schleicher, A. A. Noegel. 1992. Cell. 68:53-62) by introducing the mutated alpha-actinin cDNAs. In agreement with the biochemical data, only the molecule modified in EF hand II could rescue the abnormal phenotype. Considering the fact that the active construct is "always on" because it requires nonphysiological, high Ca2+ concentrations for inactivation, it is interesting to note that an unregulated alpha-actinin was able to rescue the mutant phenotype.  相似文献   

8.
To sustain efficient translation, eukaryotic elongation factor B alpha (eEF1B alpha) functions as the guanine nucleotide exchange factor for eEF1A. Stopped-flow kinetics using 2'-(or 3')-O-N-methylanthraniloyl (mant)-GDP showed spontaneous release of nucleotide from eEF1A is extremely slow and accelerated 700-fold by eEF1B alpha. The eEF1B alpha-stimulated reaction was inhibited by Mg2+ with a K(1/2) of 3.8 mM. Previous structural studies predicted the Lys-205 residue of eEF1B alpha plays an important role in promoting nucleotide exchange by disrupting the Mg2+ binding site. Co-crystal structures of the lethal K205A mutant in the catalytic C terminus of eEF1B alpha with eEF1A and eEF1A.GDP established that the lethality was not due to a structural defect. Instead, the K205A mutant drastically reduced the nucleotide exchange activity even at very low concentrations of Mg2+. A K205R eEF1B alpha mutant on the other hand was functional in vivo and showed nearly wild-type nucleotide dissociation rates but almost no sensitivity to Mg2+. These results indicate the significant role of Mg2+ in the nucleotide exchange reaction by eEF1B alpha and establish the catalytic function of Lys-205 in displacing Mg2+ from its binding site.  相似文献   

9.
The reaction of mitochondrial F1-ATPase with immobilized substrate was studied by using columns of agarose-hexane-ATP. Mg2+ was required for binding of the enzyme to the column matrix. The column-bound enzyme could be eluted fully by ATP and other nucleoside triphosphates. Nucleoside di- and mono-phosphates were less effective. At a fixed concentration of nucleotide the effectiveness of elution was proportional to the charge on the eluting molecule. The ATP of the column matrix was hydrolysed by the bound F1-ATPase to release phosphate, probably by a uni-site reaction mechanism. Thus the F1-ATPase was bound to the immobilized ATP by a catalytic site. Treatment of the bound F1-ATPase with 4-chloro-7-nitrobenzofurazan prevented complete release of the enzyme by ATP. Only one-third of the bound enzyme was now eluted by the nucleotide. The inhibition of release could be due either to the inhibitor blocking co-operative interactions between sites or to its increasing the tightness of binding of immobilized ADP at the catalytic site.  相似文献   

10.
The 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical-scavenging mechanism of 2-O-α-D-glucopyranosyl-L-ascorbic acid (AA-2G) was studied. We found two undefined products, named X and Y, in the reaction mixture of AA-2G and the DPPH radical under acidic conditions by HPLC analysis. The reaction mixture was further subjected to LC–MS analysis. X was found to be a covalent adduct of AA-2G and the DPPH radical. On the other hand, Y could not be identified, probably because it was a mixture. A time-course study of the radical-scavenging reaction revealed that one molecule of AA-2G scavenged one molecule of DPPH radical to generate an AA-2G radical, which readily reacted with another molecule of the DPPH radical to form a covalent adduct (X). Subsequently, this adduct slowly quenched a third molecule of the DPPH radical, resulting in reaction products (Y). Therefore, one molecule of AA-2G has only one oxidizable –OH group, but can scavenge three molecules of the DPPH radical. The radical-scavenging mechanism of AA-2G elucidated in this study should be useful in understanding the biological roles of AA-2G per se in the food and cosmetic fields.  相似文献   

11.
Actin-gelsolin interactions. Evidence for two actin-binding sites   总被引:26,自引:0,他引:26  
We have used a fluorescence enhancement of actin labeled with 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-actin) to study the interactions between rabbit skeletal muscle G-actin and either purified platelet gelsolin or a 130-kDa binary complex of platelet actin and gelsolin that is stable in EGTA and can be purified from human platelets. We have delineated four binding reactions. The exchange of Mg2+ for Ca2+ on the divalent cation-binding site of NBD-actin gives a small fluorescence increase. Binding of monomeric NBD-actin to the binary complex results in a 2.5-fold increase in the emission at 530 nm in the presence of Ca2+ and a 2-fold increase in the presence of EGTA. Titration experiments show that, under nonpolymerizing conditions, one additional actin is bound to the 130-kDa species to form a ternary complex. This binding is Ca2+-sensitive. Purified gelsolin does not appear to bind to NBD-actin in the presence of EGTA, as determined by fluorescence enhancement, gel filtration, or sedimentation measurements, but the addition of Ca2+ promotes rapid binding with a 1.6-1.7-fold enhancement of the emission intensity. A comparison of the relative fluorescence yields/NBD-actin molecule for a binary complex of gelsolin and one NBD-actin, a ternary complex of gelsolin and two NBD-actin molecules, and a ternary complex with an unlabeled actin in the EGTA-stable site and an NBD-actin in the second site indicates that the first NBD-actin, in the EGTA-stable site, does not give a fluorescence increase on binding but the second one does. Finally, we have demonstrated that one molecule of 45Ca2+ is "trapped" when the binary complex is formed and cannot be removed by EGTA. A summary model for these reactions is presented that indicates the interaction between actin and gelsolin is not a freely reversible Ca2+-controlled reaction.  相似文献   

12.
The physiological correlation between NDP-kinase and the enzyme-associated guanine nucleotide binding proteins (G1 and G2) has been studied in vitro. It was found that incubation of the phosphoenzyme (enzyme-bound high-energy phosphate intermediate) of NDP-kinases with one of the nucleoside 5'-diphosphates (NDPs) in the presence of divalent cations (Mg2+ and Ca2+) results in the formation of nucleoside 5'-triphosphates (NTPs) within 40 sec even at low temperatures (below 4 degrees C) without strict base-specificity; and high-energy phosphates on the phosphoenzyme can transfer preferentially to GDP on the guanine nucleotide binding proteins (G1, G2 and r-p21 protein) in the presence of 0.25 mM Ca2+ or 1 mM Mg2+ even if any other NDPs are present in the reaction mixtures. These observations suggest that NDP-kinase may be responsible for the phosphate-transfer between GDP on the guanine nucleotide binding proteins and its phosphoenzyme.  相似文献   

13.
In stopped-flow experiments in which oxidized cytochrome c oxidase was mixed with ferrocytochrome c in the presence of a range of oxygen concentrations and in the absence and presence of cyanide, a fast phase, reflecting a rapid approach to an equilibrium, was observed. Within this phase, one or two molecules of ferrocytochrome were oxidized per haem group of cytochrome a, depending on the concentration of ferrocytochrome c used. The reasons for this are discussed in terms of a mechanism in which all electrons enter through cytochrome a, which, in turn, is in rapid equilibrium with a second site, identified with 'visible' copper (830 nm-absorbing) Cud (Beinert et al., 1971). The value of the bimolecular rate constant for the reaction between cytochromes c2+ and a3+ was between 10(6) and 10(7) M(-1)-S(-1); some variability from preparation to preparation was observed. At high ferrocytochrome c concentrations, the initial reaction of cytochrome c2+ with cytochrome a3+ could be isolated from the reaction involving the 'visible' copper and the stoicheiometry was found to approach one molecule of cytochrome c2+ oxidized for each molecule of cytochrome a3+ reduced. At low ferrocytochrome c concentrations, however, both sites (i.e. cytochrome a and Cud) were reduced simultaneously and the stoicheiometry of the initial reaction was closer to two molecules of cytochrome c2+ oxidized per molecule of cytochrome a reduced. The bleaching of the 830 nm band lagged behind or was simultaneous with the formation of the 605 nm band and does not depend on the cytochrome c concentration, whereas the extinction at the steady-state does. The time-course of the return of the 830 nm-absorbing species is much faster than the bleaching of the 605 nm-absorbing component, and parallels that of the turnover phase of cytochrome c2+ oxidation. Additions of cyanide to the oxidase preparations had no effect on the observed stoicheiometry or kinetics of the reduction of cytochrome a and 'visible' copper, but inhibited electron transfer to the other two sites, cytochrome a3 and the undetectable copper, Cuu.  相似文献   

14.
1. Pulse-radiolysis experiments were performed in the presence of methyl viologen and cytochrome c3. After the pulse, methyl viologen radicals are formed and the kinetics of these radicals with cytochrome c3 are studied, The reaction between cytochrome c3 and methyl viologen radicals (MV+) is diffusion controlled. The ionic strength dependence and the pH-dependence of this reaction were studied. From the ionic strength dependence (at pH 7.8) we found that the net charge of the fully oxidized cytochrome c3 molecule was Z = + 4.7 +/- 0.7. 2. After the pulse an equilibrium is reached for the reaction of MV+ with cytochrome c3. From this equilibrium an apparent midpoint potential can be obtained. The apparent midpoint potential of this multihaem molecule was found to depend on the degree of reduction, alpha. With the help of the Nernst equation an empirical equation is obtained to describe this dependence of the midpoint potential: E0 = - 0.250 - 0.088 alpha (in V). 3. An estimation is made of the energy of interaction between the haems due to electrostatic interactions (delta epsilon less than 32 mV) and due to ionic strength effects (- 12 mV less than delta epsilon less than 26 mV). The results suggest that the redox properties of the individual haems in the cytochrome c3 molecule are dependent on the degree of reduction of the other haems in the molecule. 4. The reaction of cytochrome c3 with MV+ or with ethanol radicals (EtOH) has been compared with the reactions of horse-heart cytochrome c and of metmyoglobin with the same radicals. The reaction of MV+ or EtOH with horse-heart cytochrome c is found to be diffusion controlled; the reactions with metmyoglobin on the other hand are most probably controlled by an activation energy.  相似文献   

15.
The association process of FAD and apo-electron-transferring flavoprotein (apoETF) from hog kidney was investigated. The reaction schemes which involve the association-dissociation of the protein species could be excluded by the light scattering data, which indicated that the molecular weights of apoETF and holoETF are identical. The binding reaction between FAD and a large excess of apoETF was monophasic and obeyed pseudo-first order kinetics. On the other hand, the reaction between apoETF and a large excess of FAD was biphasic: the fast phase obeyed a pseudo-first order reaction, and the rate of the slow phase was almost independent of FAD concentration. These results suggest the existence of two different forms of apoETF, as represented in the following reaction scheme: [formula: see text] where "F" is FAD, "H" is holoETF, and "A" and "A" are the different forms of apoETF. The kinetic parameters were determined as k-1 = 3.9 x 10(4) M-1.s-1, k-1 approximately 10(-5) s-1, k+2 = 1.0 x 10(-3) s-1, and k-2 = 3.1 x 10(-3) s-1, in 50 mM potassium phosphate buffer, pH 7.6, containing 0.3 mM EDTA, and 5% v/v glycerol, at 7 degrees C. The elution patterns of apoETF on molecular sieve chromatography were very different from that of holoETF although the true molecular weights were identical. This result suggests that the structure of apoETF differs greatly from that of holoETF.  相似文献   

16.
Many types of microorganisms, including both prokaryotes and eukaryotes, have developed mechanisms to adapt to severe osmotic stress. In this study, we isolated multicopy suppressor genes for a Schizosaccharomyces pombe mutant, which exhibited the clear phenotype of being osmosensitive for growth (Osms) on agar plates containing high concentrations of either non-ionic or ionic osmotic solutes. Two genes were thus identified, and each was suggested to encode an NADH-dependent glycerol-3-phosphate dehydrogenase (GPD), which is required for glycerol synthesis. The nucleotide sequences, determined for these genes (named gpd1 + and gpd2 +, respectively), revealed that S. pombe has two distinct GPD isozymes. They are only 60% identical to each other in their amino acid sequences. One such isozyme, GPD1, was shown to be directly involved in osmoregulation, based on the following observations. (i) Expression of gpd1 + was regulated at the mRNA level in response to osmotic upshift, (ii) It was demonstrated that wild-type cells markedly accumulated internal glycerol under high-osmolarity growth conditions. (iii) Δ gpd1 mutants, however, failed to do so even in a high-osmolarity medium, and thus exhibited an Osms phenotype. On the other hand, the gpd2 + gene was constitutively expressed at a particular low level, regardless of the osmolarity of the medium.  相似文献   

17.
Periodate-oxidized guanine nucleotides (GTPox and GDPox) were shown to bind stoichiometrically to rat liver elongation factor 2 (EF-2). This binding was quantitatively inhibited in the presence of GTP. After binding, oxidized nucleotides remained on EF-2 despite extensive dialysis. They exchanged, however, with free quanine nucleotides in the course of prolonged (greater than 1 h) incubations. The prior reduction EF-2.GTPox with NaBH4 abolished, to a large extent, this slow exchange. Thus, a Schiff's base was implicated to be formed between EF-2 and oxidized guanine nucleotides. Mg2+ increased the GTPox concentration necessary for a stoichiometric binding to EF-2. EF-2-oxidized nucleotide conjugates bound in the presence of ribosomes a second molecule of GTP (or GTPox). GTPox bound to EF-2 in the presence of ribosomes appeared to exchange readily with free GTP. Moreover, GTPox proved to be active as substrate in EF-2 and ribosome-dependent GTPase reaction: Km values found for GTPox and GTP were 7.7 and 3.4 microM, respectively. The binding of GTPox to EF-2 inhibited only partially the subsequent ribosome-dependent GTP binding, and GTPase reaction or polyphenylalanine (polyPhe) synthesis. On the other hand, the binding of GuoPP[CH2]Pox to EF-2 inhibited all of these reactions strongly. The nature of the binding site involved in the direct interactions of EF-2 with guanine nucleotides is discussed in the light of these results.  相似文献   

18.
Sullivan SM  Holyoak T 《Biochemistry》2007,46(35):10078-10088
The structures of the rat cytosolic isoform of phosphoenolpyruvate carboxykinase (PEPCK) reported in the PEPCK-Mn2+, -Mn2+-oxaloacetic acid (OAA), -Mn2+-OAA-Mn2+-guanosine-5'-diphosphate (GDP), and -Mn2+-Mn2+-guanosine-5'-tri-phosphate (GTP) complexes provide insight into the mechanism of phosphoryl transfer and decarboxylation mediated by this enzyme. OAA is observed to bind in a number of different orientations coordinating directly to the active site metal. The Mn2+-OAA and Mn2+-OAA-Mn2+GDP structures illustrate inner-sphere coordination of OAA to the manganese ion through the displacement of two of the three water molecules coordinated to the metal in the holo-enzyme by the C3 and C4 carbonyl oxygens. In the PEPCK-Mn2+-OAA complex, an alternate bound conformation of OAA is present. In this conformation, in addition to the previous interactions, the C1 carboxylate is directly coordinated to the active site Mn2+, displacing all of the waters coordinated to the metal in the holo-enzyme. In the PEPCK-Mn2+-GTP structure, the same water molecule displaced by the C1 carboxylate of OAA is displaced by one of the gamma-phosphate oxygens of the triphosphate nucleotide. The structures are consistent with a mechanism of direct in-line phosphoryl transfer, supported by the observed stereochemistry of the reaction. In the catalytically competent binding mode, the C1 carboxylate of OAA is sandwiched between R87 and R405 in an environment that would serve to facilitate decarboxylation. In the reverse reaction, these two arginines would form the CO2 binding site. Comparison of the Mn2+-OAA-Mn2+GDP and Mn2+-Mn2+GTP structures illustrates a marked difference in the bound conformations of the nucleotide substrates in which the GTP nucleotide is bound in a high-energy state resulting from the eclipsing of all three of the phosphoryl groups along the triphosphate chain. This contrasts a previously determined structure of PEPCK in complex with a triphosphate nucleotide analogue in which the analogue mirrors the conformation of GDP as opposed to GTP. Last, the structures illustrate a correlation between conformational changes in the P-loop, the nucleotide binding site, and the active site lid that are important for catalysis.  相似文献   

19.
Phosphoenolpyruvate carboxykinase from bullfrog liver mitochondria has been purified to electrophoretical and immunological homogeneity by an improved method using hydrophobic chromatography on Sepharose-hexane-GMP and affinity chromatography on phosphocellulose. The molecular weight was determined to be 70,000 by SDS-gel electrophoresis, 65,000 by Sephadex G-100 gel filtration and 72,000 by glycerol gradient centrifugation. The isoelectric point was determined to be 6.2, differing from that of the cytosol enzyme. The rabbit IgG fraction against the mitochondrial PEP carboxykinase precipitated not only the mitochondrial but also the cytosol enzyme. The dissociation constant of the nucleotide-enzyme complex was determined to be 3 microM for GTP, 8.5 microM for GDP, and 171 microM for GMP. The affinity of GTP for the enzyme was reduced in the presence of phosphoenolpyruvate or Mn2+, whereas that of GDP was not changed. GMP inhibited the enzyme competitively with GDP for the phosphoenolpyruvate carboxylation and competitively with GTP for the exchange reaction between [14C]HCO3- and oxaloacetate. The purified enzyme was found to have a cysteine residue which reacted with iodoacetamide to form inactive enzyme. Guanine nucleotides or IDP and Mn2+ at a lower concentration prevented the inactivation by iodoacetamide of the enzyme in a competitive manner. Binding of guanine nucleotide to the enzyme and the relation of the sulfhydryl group to the nucleotide binding are discussed.  相似文献   

20.
Experiments were carried out to measure the effect of concentrations of glycerol on H-exchange (HX) rates by using myoglobin as a test protein. Concentrated glycerol has only a small slowing effect on the HX kinetics of freely exposed amides, studied in a small molecule model (acetamide). Larger effects occur in structured proteins. The effect of solvent glycerol on different parts of the HX curve of myoglobin was studied by use of a selective "kinetic labeling" approach. Concentrated glycerol exerts an apparently reverse effect on protein H exchange; the faster exchanging "surface" protons are least affected, while the slower and slower amide NH is further slowed by larger and larger factors. These results seem inconsistent with solvent penetration models which generally visualize slower and slower protons as being placed, and undergoing exchange, farther and farther from the solvent-protein interface. On the other hand, the results are as expected for the local unfolding model for protein H exchange since concentrated glycerol is known to stabilize proteins against unfolding. In the local unfolding model, slower exchanging protons are released by way of higher energy and therefore generally larger, unfolding reactions. Larger unfoldings must be more inhibited by the glycerol effect.  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