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 protein folding process of heme proteins entails generation of not only a correct global polypeptide structure, but also a correct, functionally competent heme environment. We employed a variety of spectroscopic approaches to probe the structure and dynamics of the heme pocket of a recombinant sperm whale myoglobin. The conformational characteristics were examined by circular dichroism, time-resolved fluorescence spectroscopy, FTIR spectroscopy, and optical absorption spectroscopy in the temperature range 300-20 K. Each of these spectroscopic probes detected modifications confined exclusively to the heme pocket of the expressed myoglobin relative to the native protein. The functional properties were examined by measuring the kinetics of CO binding after flash-photolysis. The kinetics of the expressed myoglobin were more heterogeneous than those of the native protein. Mild acid exposure of the ferric derivative of the recombinant protein resulted in a protein with "nativelike" spectroscopic properties and homogeneous CO binding kinetics. The heme pocket modifications observed in this recombinant myoglobin do not derive from inverted heme. In contrast, when native apomyoglobin is reconstituted with the heme in vitro, the heme pocket disorder could be attributed exclusively to 180 degrees rotation of the bound heme [La Mar, G. N., Toi, H., and Krishnamoorthi, R. (1984) J. Am. Chem. Soc. 106, 6395-6401; Light, W. R., Rohlfs, R. J., Palmer, G., and Olson, J. S. (1987) J. Biol. Chem. 262, 46-52]. We conclude that exposure to low pH decreases the affinity of globin for the heme and allows an extended conformational sampling or "soft refolding" to a nativelike conformation.  相似文献   

2.
The exposure to non-thermal microwave electromagnetic field (MW-EMF) at 1.95 MHz, a frequency used in mobile communication, affects the refolding kinetics of eukaryotic proteins (Mancinelli et al., 2004). On these basis we have evaluated the in vivo effect of MW-EMF in human epidermoid cancer KB cells. We have found that MW-EMF induces time-dependent apoptosis (45% after 3 h) that is paralleled by an about 2.5-fold decrease of the expression of ras and Raf-1 and of the activity of ras and Erk-1/2. Although also the expression of Akt was reduced its activity was unchanged likely as a consequence of the increased expression of its upstream activator PI3K. In the same experimental conditions an about 2.5-fold increase of the ubiquitination of ras and Raf-1 was also found and the addition for 12 h of proteasome inhibitor lactacystin at 10 microM caused an accumulation of the ubiquitinated isoforms of ras and Raf-1 and counteracted the effects of MW-EMF on ras and Raf-1 expression suggesting an increased proteasome-dependent degradation induced by MW-EMF. The exposure of KB cells to MW-EMF induced a differential activation of stress-dependent pathway with an increase of JNK-1 activity and HSP70 and 27 expression and with a reduction of p38 kinase activity and HSP90 expression. The overexpression of HSP90 induced by transfection of KB cells with a plasmid encoding for the factor completely antagonized the apoptosis and the inactivation of the ras --> Erk-dependent survival signal induced by MW-EMF. Conversely, the inhibition of Erk activity induced by 12 h exposure to 10 mM Mek-1 inhibitor U0126 antagonized the effects induced by HSP90 transfection on apoptosis caused by MW-EMF. In conclusion, these results demonstrate for the first time that MW-EMF induces apoptosis through the inactivation of the ras --> Erk survival signaling due to enhanced degradation of ras and Raf-1 determined by decreased expression of HSP90 and the consequent increase of proteasome dependent degradation.  相似文献   

3.
Lactoperoxidase (LPO), a mammalian secretory heme peroxidase, catalyzes the oxidation of thiocyanate by hydrogen peroxide to produce hypothiocyanate, an antibacterial agent. Although LPO is known to be activated at acidic pH and in the presence of iodide, the structural basis of the activation is not well understood. We have examined the effects of pH and iodide concentration on the catalytic activity and the structure of LPO. Electrochemical and colorimetric assays have shown that the catalytic activity is maximized at pH 4.5. The heme Soret absorption band exhibits a small red‐shift at pH 5.0 upon acidification, which is ascribable to a structural transition from a neutral to an acidic form. Resonance Raman spectra suggest that the heme porphyrin core is slightly contracted and the Fe‐His bond is strengthened in the acidic form compared to the neutral form. The structural change of LPO upon activation at acidic pH is similar to that observed for myeloperoxidase, another mammalian heme peroxidase, upon activation at neutral pH. Binding of iodide enhances the catalytic activity of LPO without affecting either the optimum pH of activity or the heme structure, implying that the iodide binding occurs at a protein site away from the heme‐linked protonation site. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 113–120, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

4.
The last step of the folding reaction of myoglobin is the incorporation of a prosthetic group. In cells, myoglobin is soluble, while heme resides in the mitochondrial membrane. We report here an exhaustive study of the interactions of apomyoglobin with lipid vesicles. We show that apomyoglobin interacts with large unilamellar vesicles under acidic conditions, and that this requires the presence of negatively charged phospholipids. The pH dependence of apomyoglobin interactions with membranes is a two-step process, and involves a partially folded state stabilized at acidic pH. An evident role for the interaction of apomyoglobin with lipid bilayers would be to facilitate the uptake of heme from the outer mitochondrial membrane. However, heme binding to apomyoglobin is observed at neutral pH when the protein remains in solution, and slows down as the pH becomes more favorable to membrane interactions. The effective incorporation of soluble heme into apomyoglobin at neutral pH suggests that the interaction of apomyoglobin with membranes is not necessary for the heme uptake from the lipid bilayer. In vivo, however, the ability of apomyoglobin to interact with membrane may facilitate its localization in the vicinity of the mitochondrial membranes, and so may increase the yield of heme uptake. Moreover, the behavior of apomyoglobin in the presence of membranes shows striking similarities with that of other proteins with a globin fold. This suggests that the globin fold is well adapted for soluble proteins whose functions require interactions with membranes.  相似文献   

5.
M Sono 《Biochemistry》1990,29(6):1451-1460
The binding of a number of ligands to the heme protein indolamine 2,3-dioxygenase has been examined with UV-visible absorption and with natural and magnetic circular dichroism spectroscopy. Relatively large ligands (e.g., norharman) which do not readily form complexes with myoglobin and horseradish peroxidase (HRP) can bind to the dioxygenase. Except for only a few cases (e.g., 4-phenylimidazole) for the ferric dioxygenase, a direct competition for the enzyme rarely occurs between the substrate L-tryptophan (Trp) and the ligands examined. L-Trp and small heme ligands (CN-,N3-,F-) markedly enhance the affinity of each other for the ferric enzyme in a reciprocal manner, exhibiting positive cooperativity. For the ferrous enzyme, L-Trp exerts negative cooperativity with some ligands such as imidazoles, alkyl isocyanides, and CO binding to the enzyme. This likely reflects the proximity of the Trp binding site to the heme iron. Other indolamine substrates also exert similar but smaller cooperative effects on the binding of azide or ethyl isocyanide. The pH dependence of the ligand affinity of the dioxygenase is similar to that of myoglobin rather than that of HRP. These results suggest that indolamine 2,3-dioxygenase has the active-site heme pocket whose environmental structure is similar to, but whose size is considerably larger than, that of myoglobin, a typical O2-binding heme protein. Although the L-Trp affinity of the ferric cyanide and ferrous CO enzyme varies only slightly between pH 5.5 and 9.5, the unligated ferric and ferrous enzymes have considerably higher affinity for L-Trp at alkaline pH than at acidic pH. L-Trp binding to the ferrous dioxygenase is affected by an ionizable residue with a pKa value of 7.3.  相似文献   

6.
UV-visible absorption and magnetic circular dichroism (MCD) data are reported for the cavity mutants of sperm whale H93G myoglobin and human H25A heme oxygenase in their ferric states at 4 degreesC. Detailed spectral analyses of H93G myoglobin reveal that its heme coordination structure has a single water ligand at pH 5.0, a single hydroxide ligand at pH 10.0, and a mixture of species at pH 7.0 including five-coordinate hydroxide-bound, and six-coordinate structures. The five-coordinate aquo structure at pH 5 is supported by spectral similarity to acidic horseradish peroxidase (pH 3.1), whose MCD data are reported herein for the first time, and acidic myoglobin (pH 3.4), whose structures have been previously assigned by resonance Raman spectroscopy. The five-coordinate hydroxide structure at pH 10.0 is supported by MCD and resonance Raman data obtained here and by comparison with those of other known five-coordinate oxygen donor complexes. In particular, the MCD spectrum of alkaline ferric H93G myoglobin is strikingly similar to that of ferric tyrosinate-ligated human H93Y myoglobin, whose MCD data are reported herein for the first time, and that of the methoxide adduct of ferric protoporphyrin IX dimethyl ester (FeIIIPPIXDME). Analysis of the spectral data for ferric H25A heme oxygenase at neutral pH in the context of the spectra of other five-coordinate ferric heme complexes with proximal oxygen donor ligands, in particular the p-nitrophenolate and acetate adducts of FeIIIPPIXDME, is most consistent with ligation by a carboxylate group of a nearby glutamyl (or aspartic) acid residue.  相似文献   

7.
Changes in the circular dichroic and absorption spectra were studied on solutions of myoglobin whose histidine residues had been modified by carboxymethylation under denaturing conditions. Carboxymethylation resulted in a dramatic decrease in the molar extinction coefficient in the Soret region indicative of a major change in the heme environment. This was accompanied by a remarkable change in the secondary structure of the protein involving helix-to-random coil transition, indicating that extensive histidine modification prevented unfolded myoglobin from refolding to its native conformation.  相似文献   

8.
9.
The molecular properties of the salt-induced partly folded acidic state of apomyoglobin as well as myoglobin were investigated by fluorescence and circular dichroism of the extrinsic fluorophore 1,8-anilinonaphthalenesulfonate. The occurrence of a fluctuating tertiary structure ("molten globule") at acidic pH in the presence of salt was suggested by the disappearance of the dichroic activity of the fluorophore bound to the partly folded protein. Moreover, the structure of the intermediate is not influenced by the presence of heme, thus suggesting that heme is not crucial in the early stage of myoglobin folding.  相似文献   

10.
Equilibrium and kinetic folding studies of horse cytochrome c in the reduced state have been carried out under strictly anaerobic conditions at neutral pH, 10 degrees C, in the entire range of aqueous solubility of guanidinium hydrochloride (GdnHCl). Equilibrium unfolding transitions observed by Soret heme absorbance, excitation energy transfer from the lone tryptophan residue to the ferrous heme, and far-UV circular dichroism (CD) are all biphasic and superimposable, implying no accumulation of structural intermediates. The thermodynamic parameters obtained by two-state analysis of these transitions yielded DeltaG(H2O)=18.8(+/-1.45) kcal mol(-1), and C(m)=5.1(+/-0.15) M GdnHCl, indicating unusual stability of reduced cytochrome c. These results have been used in conjunction with the redox potential of native cytochrome c and the known stability of oxidized cytochrome c to estimate a value of -164 mV as the redox potential of the unfolded protein. Stopped-flow kinetics of folding and unfolding have been recorded by Soret heme absorbance, and tryptophan fluorescence as observables. The refolding kinetics are monophasic in the transition region, but become biphasic as moderate to strongly native-like conditions are approached. There also is a burst folding reaction unobservable in the stopped-flow time window. Analyses of the two observable rates and their amplitudes indicate that the faster of the two rates corresponds to apparent two-state folding (U<-->N) of 80-90 % of unfolded molecules with a time constant in the range 190-550 micros estimated by linear extrapolation and model calculations. The remaining 10-20 % of the population folds to an off-pathway intermediate, I, which is required to unfold first to the initial unfolded state, U, in order to refold correctly to the native state, N (I<-->U<-->N). The slower of the two observable rates, which has a positive slope in the linear functional dependence on the denaturant concentration indicating that an unfolding process under native-like conditions indeed exists, originates from the unfolding of I to U, which rate-limits the overall folding of these 10-20 % of molecules. Both fast and slow rates are independent of protein concentration and pH of the refolding milieu, suggesting that the off-pathway intermediate is not a protein aggregate or trapped by heme misligation. The nature or type of unfolded-state heme ligation does not interfere with refolding. Equilibrium pH titration of the unfolded state yielded coupled ionization of the two non-native histidine ligands, H26 and H33, with a pK(a) value of 5.85. A substantial fraction of the unfolded population persists as the six-coordinate form even at low pH, suggesting ligation of the two methionine residues, M65 and M80. These results have been used along with the known ligand-binding properties of unfolded cytochrome c to propose a model for heme ligation dynamics. In contrast to refolding kinetics, the unfolding kinetics of reduced cytochrome c recorded by observation of Soret absorbance and tryptophan fluorescence are all slow, simple, and single-exponential. In the presence of 6.8 M GdnHCl, the unfolding time constant is approximately 300(+/-125) ms. There is no burst unfolding reaction. Simulations of the observed folding-unfolding kinetics by numerical solutions of the rate equations corresponding to the three-state I<-->U<-->N scheme have yielded the microscopic rate constants.  相似文献   

11.
Three heme proteins, myoglobin, hemoglobin, and cytochrome c, have been adsorbed onto chitosan-stabilized gold nanoparticles (Chit-Aus) modified Au electrode via a molecule bridge like cysteine. UV-vis spectra indicated that the proteins on Chit-Aus films retained near-native secondary structures. The fabricated procedures and electrochemical behaviors of proteins on such an interface were characterized with electrochemical impedance spectra and cyclic voltammetric techniques. It was demonstrated that Chit-Aus film could not only offer a friendly environment to immobilize protein molecules but also enhance the electron transfer ability between protein molecules and underlying electrode. The effects of scan rate and pH on the electrochemical behaviors of each heme protein are discussed in detail. The resultant electrode displayed an excellent electrocatalytic response to the reduction of H(2)O(2), long-term stability, and good reproducibility.  相似文献   

12.
The present work investigates the variations of electrostatic interactions within the myoglobin molecule associated with azide heme binding and pH variations. Far ultraviolet (223 nm) resonance Raman spectroscopy of the tryptophan and tyrosine residues, along with acid-base titration measurements, have been used to monitor variations in the protein matrix. With previously determined mode assignments, it is shown that the Trp and Tyr residues of the globin moiety are influenced by the charge spatial distribution. Upon ligand binding or under various pH conditions, the polar interactions inside the protein appear to be modulated by the electric field generated by the charge array. It is concluded that the binding site properties of myoglobin can be modulated by the charge spatial distribution within the protein, even in the absence of measurable conformational changes of the bulk.  相似文献   

13.
Abstract

The interaction between myoglobin and negatively-charged liposomes composed of phosphatidylcholine/phosphatidylglycerol (1:1) was studied at low ionic strength under acidic conditions. Changes in the absorbance and the fluorescence spectra of myoglobin were recorded upon addition of liposomes to partially unfolded (pH 3.5) and native (pH 4.5 and pH 6.5) myoglobin. Association of myoglobin with liposomes was a relatively fast process at pH 3.5 and pH 4.5. Although at pH 3.5 myoglobin was unfolded partially before the addition of the liposomes while at pH 4.5 before the addition of liposomes myoglobin retained its native form, similar interaction patterns of myoglobin with liposomes were observed. The fluorescence and absorption spectra in the Soret region of myoglobin clearly indicated that at these pH values myoglobin was associated with the liposomes in a (partially) unfolded state. At pH 6.5 the kinetics of myoglobin association with liposomes was much slower than at pH 3.5 and 4.5. The spectroscopic measurements also indicated that the interaction of myoglobin with liposomes at pH 6.5 followed a different pattern and resulted in different protein structures in comparison with pH 3.5/4.5.  相似文献   

14.
The photosynthetic purple sulfur bacterium Chromatium vinosum has been shown to possess two previously undetected heme c-containing, soluble proteins. One is an acidic, c-type cytochrome with a molecular weight of 12 300 and an oxidation-reduction midpoint potential (at pH 8.0) of ?82 mV. The other protein is a basic protein with a molecular weight of 11 900 and an oxidation-reduction midpoint potential (at pH 8.0) of ?110 mV. The basic protein, in both oxidized and reduced forms, has optical spectra similar to those of myoglobin and the oxidized C. vinosum protein exhibits a high-spin heme EPR spectrum similar to that of metmyoglobin. Furthermore, the basic C. vinosum protein binds CO and O2. The spectra of the CO and O2 complexes show significant similarities with the respective myoglobin complexes. Possible functions for an O2-binding protein in C. vinosum are discussed.  相似文献   

15.
Cytochrome c(553) (cyt c(553)) from Desulfovibrio vulgaris is a small helical heme protein that displays apparent two-state equilibrium-unfolding behavior. The covalently attached heme is low-spin, ligated by Met and His residues, in the native state but becomes high-spin upon unfolding at pH 7. Here, we show that in contrast to other c-type heme proteins, where misligations in the unfolded states are prominent, cyt c(553) refolding kinetics at pH 7 proceeds rapidly without detectable intermediates. The extrapolated folding rate constant in water for oxidized cyt c(553) matches exactly that predicted from the cyt c(553) native-state topology: 5300 s(-1 )(experimental) versus 5020 s(-1) (predicted). We therefore conclude that the presence of the oxidized cofactor does not affect the intrinsic formation speed of the cyt c(553 )structural motif.  相似文献   

16.
The folding and stabilization of α-helical transmembrane proteins are still not well understood. Following cofactor binding to a membrane protein provides a convenient method to monitor the formation of appropriate native structures. We have analyzed the assembly and stability of the transmembrane cytochrome b559′, which can be efficiently assembled in vitro from a heme-binding PsbF homo-dimer by combining free heme with the apo-cytochrome b559′. Unfolding of the protein dissolved in the mild detergent dodecyl maltoside may be induced by addition of SDS, which at high concentrations leads to dimer dissociation. Surprisingly, absorption spectroscopy reveals that heme binding and cytochrome formation at pH 8.0 are optimal at intermediate SDS concentrations. Stopped-flow kinetics revealed that genuine conformational changes are involved in heme binding at these SDS concentrations. GPS (Global Protein folding State mapping) NMR measurements showed that optimal heme binding is intimately related to a change in the degree of histidine protonation. In the absence of SDS, the pH curve for heme binding is bell-shaped with an optimum at around pH 6-7. At alkaline pH values, the negative electrostatic potential of SDS lowers the local pH sufficiently to restore efficient heme binding, provided the amount of SDS needed for this does not denature the protein. Accordingly, the higher the pH value above 6-7, the more SDS is needed to improve heme binding, and this competes with the inherent tendency of SDS to dissociate cytochrome b559′. Our work highlights that, in addition to its denaturing properties, SDS can affect protein functions by lowering the local pH.  相似文献   

17.
The rates of reaction of myoglobin with carbon monoxide at low pH are reported. The pH versus rate profile of these kinetics resembles that found for heme model compounds, revealing an increase in combination rate at low pH. These facts suggest that CO binding by myoglobin changes from a mechanism of "direct ligant association" at pH 5 to a mechanism, similar to that proposed for heme model compounds, which assumes a tetracoordinated intermediate as a result of the protonation of the proximal imidazole.  相似文献   

18.
The effects of aqueous ethanol or 2,2,2-trifluoroethanol media on the structure of sperm whale myoglobin have been investigated by absorption, CD, and NMR spectra. The structural properties of myoglobin such as heme environments, helix contents, protein folding, and interactions between heme and the protein moiety have been sharply manifested in these spectra. The characterization demonstrated that alcohol-induced conformational change of myoglobin depends on the nature of alcohol and its concentration. It was shown for the first time that, upon the alcohol-induced denaturation of myoglobin, heme is released from partially denatured protein of which helix contents is altered by only about 20% relative to that of native state. Myoglobin has shown to unfold and refold reversibly by controlling the alcohol concentration. Novel methods for the preparation of apomyoglobin and in situ reconstitution of apomyoglobin with heme, based on the alcohol-induced denaturation of the protein, were presented.  相似文献   

19.
Denitrifying NO reductases are evolutionarily related to the superfamily of heme--copper terminal oxidases. These transmembrane protein complexes utilize a heme-nonheme diiron center to reduce two NO molecules to N(2)O. To understand this reaction, the diiron site has been modeled using sperm whale myoglobin as a scaffold and mutating distal residues Leu-29 and Phe-43 to histidines and Val-68 to a glutamic acid to create a nonheme Fe(B) site. The impact of incorporation of metal ions at this engineered site on the reaction of the ferrous heme with one NO was examined by UV-vis absorption, EPR, resonance Raman, and FTIR spectroscopies. UV--vis absorption and resonance Raman spectra demonstrate that the first NO molecule binds to the ferrous heme, but while the apoproteins and Cu(I)- or Zn(II)-loaded proteins show characteristic EPR signatures of S = 1/2 six-coordinate heme {FeNO}(7) species that can be observed at liquid nitrogen temperature, the Fe(II)-loaded proteins are EPR silent at ≥30 K. Vibrational modes from the heme [Fe-N-O] unit are identified in the RR and FTIR spectra using (15)NO and (15)N(18)O. The apo and Cu(I)-bound proteins exhibit ν(FeNO) and ν(NO) that are only marginally distinct from those reported for native myoglobin. However, binding of Fe(II) at the Fe(B) site shifts the heme ν(FeNO) by 17 cm(-1) and the ν(NO) by -50 cm(-1) to 1549 cm(-1). This low ν(NO) is without precedent for a six-coordinate heme {FeNO}(7) species and suggests that the NO group adopts a strong nitroxyl character stabilized by electrostatic interaction with the nearby nonheme Fe(II). Detection of a similarly low ν(NO) in the Zn(II)-loaded protein supports this interpretation.  相似文献   

20.
The refolding kinetics of ribonuclease S have been measured by tyrosine absorbance, by tyrosine fluorescence emission, and by rapid binding of the specific inhibitor 2′CMP 2 to folded RNAase S. The S-protein is first unfolded at pH 1.7 and then either mixed with S-peptide as refolding is initiated by a stopped-flow pH jump to pH 6.8, or the same results are obtained if S-protein and S-peptide are present together before refolding is initiated. The refolding kinetics of RNAase S have been measured as a function of temperature (10 to 40 °C) and of protein concentration (10 to 120 μm). The results are compared to the folding kinetics of S-protein alone and to earlier studies of RNAase A. A thermal folding transition of S-protein has been found below 30 °C at pH 1.7; its effects on the refolding kinetics are described in the following paper (Labhardt &; Baldwin, 1979).In this paper we characterize the refolding kinetics of unfolded S-protein, as it is found above 30 °C at pH 1.7, together with the kinetics of combination between S-peptide and S-protein during folding at pH 6.8. Two classes of unfolded S-protein molecules are found, fast-folding and slow-folding molecules, in a 20: 80 ratio. This is the same result as that found earlier for RNAase A; it is expected if the slow-folding molecules are produced by the slow cis-trans isomerization of proline residues after unfolding, since S-protein contains all four proline residues of RNAase A.The refolding kinetics of the fast-folding molecules show clearly that combination between S-peptide and S-protein occurs before folding of S-protein is complete. If combination occurred only after complete folding, then the kinetics of formation of RNAase S should be rather slow (5 s and 100 s at 30 °C) and nearly independent of protein concentration, as shown by separate measurements of the folding kinetics of S-protein, and of the combination between S-peptide and folded S-protein. The observed folding kinetics are faster than predicted by this model and also the folding rate increases strongly with protein concentration (apparent 1.6 order kinetics). The fact that RNAase S is formed more rapidly than S-protein alone is sufficient by itself to show that combination with S-peptide precedes complete folding of S-protein. Computer simulation of a simple, parallel-pathway scheme is able to reproduce the folding kinetics of the fast-folding molecules. All three probes give the same folding kinetics.These results exclude the model for protein folding in which the rate-limiting step is an initial diffusion of the polypeptide chain into a restricted range of three-dimensional configurations (“nueleation”) followed by rapid folding (“propagation”). If this model were valid, one would expect comparable rates of folding for RNAase A and for S-protein and one would also expect to find no populated folding intermediates, so that combination between S-peptide and S-protein should occur after folding is complete. Instead, RNAase A folds 60 times more rapidly than S-protein and also combination with S-peptide occurs before folding of S-protein is complete. The results demonstrate that the folding rate of S-protein increases after the formation, or stabilization, of an intermediate which results from combination with S-peptide. They support a sequential model for protein folding in which the rates of successive steps in folding depend on the stabilities of preceding intermediates.The refolding kinetics of the slow-folding molecules are complex. Two results demonstrate the presence of folding intermediates: (1) the three probes show different kinetic progress curves, and (2) the folding kinetics are concentration-dependent, in contrast to the results expected if complete folding of S-protein precedes combination with S-peptide. A faster phase of the slow-refolding reaction is detected both by tyrosine absorbance and fluorescence emission but not by 2′CMP binding, indicating that native RNAase S is not formed in this phase. Comparison of the kinetic progress curves measured by different probes is made with the use of the kinetic ratio test, which is defined here.  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