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.
At pH 1.7 S-peptide dissociates from S-protein but S-protein remains partly folded below 30 °C. A folded form of S-protein, labeled I3, is detected and measured by its ability to combine rapidly with S-peptide at pH 6.8 and then to form native ribonuclease S. The second-order combination reaction (k = 0.7 × 106m?1s?1 at 20 °C) can be monitored either by tyrosine absorbance or fluorescence emission; the subsequent first-order folding reaction (half-time, 68 ms; 20 °C) is monitored by 2′CMP 2 binding. Combination with S-peptide and folding to form native RNAase S is considerably slower for both classes of unfolded S-protein (see preceding paper).I3 shows a thermal folding transition at pH 1.7: it is completely unfolded above 32 °C and reaches a limiting low-temperature value of 65% below 10 °C. The 35% S-protein remaining at 10 °C is unfolded as judged by its refolding behavior in forming native RNAase S at pH 6.8. The folding transition of S-protein at pH 1.7 is a broad, multi-state transition. This is shown both by the large fraction of unfolded S-protein remaining at low temperatures and by the large differences between the folding transition curves monitored by I3 and by tyrosine absorbance.The fact that S-protein remains partly folded after dissociation of S-peptide at pH 1.7 but not at pH 6.8 may be explained by two earlier observations. (1) Native RNAase A is stable in the temperature range of the S-protein folding transition at pH 1.7, and (2) the binding constant of S-protein for S-peptide falls steadily as the pH is lowered, by more than four orders of magnitude between pH 8.3 and pH 2.7, at 0 °C. The following explanation is suggested for why folding intermediates are observed easily in the transition of S-protein but not of RNAase A. The S-protein transition is shifted to lower temperatures, where folding intermediates should be more stable: consequently, intermediates in the folding of RNAase A which do not involve the S-peptide moiety and which are populated to almost detectable levels can be observed at the lower temperatures of the S-protein transition.  相似文献   

2.
A medium resolution hydrogen exchange method (Rosa & Richards, 1979) has been used to measure the average rates of amide hydrogen exchange for known segments of the S-protein portion of ribonuclease-S. The analytical procedure permitted exchange rates to be monitored for seven S-protein fragments distributed throughout the structure, including regions of α-helix and β-sheet. Kinetics were measured as a function of pH, temperature and S-peptide binding.The pH dependence of exchange from isolated S-protein between pH 2·8 and pH 7·0 was found to deviate significantly from a first-order dependence on hydroxide ion concentration. The protection against exchange with increasing pH appeared to be closely related to the electrostatic stabilization of S-protein. It is suggested that such favorable electrostatic interactions result in increased energy barriers to the conformational fluctuations that provide solvent access to the time-average crystallographic structure. This explanation of the observed correlation between stability and exchange kinetics is also consistent with the calculated apparent activation energies for exchange from S-protein between 5·5 and 20 °C.S-peptide binding dramatically slows exchange from many S-protein sites, even those distant from the area of S-peptide contact. Interestingly, the effects of complex formation are not evenly propagated throughout S-protein. The most significantly perturbed sites (≥103-fold reduction in exchange rate constants) lie within fragments derived from regions of secondary structure. Exchange from several other fragments is not significantly affected. The S-peptide—S-protein dissociation constant at neutral pH is so small that the measured exchange must have occurred from the complex and not from the dissociated parts.  相似文献   

3.
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   

4.
The kinetics of regain of 2′-CMP binding are monitored during renaturation of RNAase S. Experiments were performed by mixing equimolar amounts of S-peptide with S-protein. The S-protein fragment was incubated initially (i.e. before mixing with S-peptide) at pH 6.2 or 1.7 and various guanidine hydrochloride (GuHCl) concentrations. Three well-resolved phases are observed. The fastest phase is second-order. The reciprocal half-time increases linearly with fragment concentration and is independent of initial conditions for the S-protein fragment. An apparent on rate of kon = 2 × 105m?1s?1 is measured in 0.5 m-GuHCl (pH 6.2) and 20 ° C. Identical association kinetics are observed by changes in tyrosine absorbance. The fraction of native RNAase S formed in this second-order reaction strictly equals the fraction of S-protein molecules with intact β-sheet in initial conditions. The relation holds for different pH values, GuHCl concentrations and temperatures. The fraction of apparent helical content of S-protein in initial conditions may also vary but this is not reflected by the association reaction. We interpret this to mean that the β-sheet but not the α-helices must be preformed in initial conditions in order to generate the high-affinity peptide binding site of S-protein. Furthermore, it is concluded that the S-protein moiety β-sheet forms or unfolds in a single one-step reaction. 2′-CMP binding reports, additionally, two slower phases of renaturation. These are produced by S-protein molecules that have their β-sheet unfolded in initial conditions. It is observed that a unique dependence of these two folding rates exists for RNAase A, RNAase S and S-protein as function of tm, the temperature of half-completion of thermal denaturation as measured by unfolding of the β-sheet in the respective compound in final conditions. The tm value varies with changing pH, with GuHCl concentration and (for RNAase S) with changing fragment concentration. The findings are interpreted to argue in favor of a sequential mechanism of folding, where the stability of a structural precursor determines the rate of folding.  相似文献   

5.
We make use of the known exchange rates of individual amide proton in the S-peptide moiety of ribonuclease S (RNAase S) to determine when during folding the alpha-helix formed by residues 3 to 13 becomes stable. The method is based on pulse-labeling with [3H]H2O during the folding followed by an exchange-out step after folding that removes 3H from all amide protons of the S-peptide except from residues 7 to 14, after which S-peptide is separated rapidly from S-protein by high performance liquid chromatography. The slow-folding species of unfolded RNAase S are studied. Folding takes place in strongly native conditions (pH 6.0, 10 degrees C). The seven H-bonded amide protons of the 3-13 helix become stable to exchange at a late stage in folding at the same time as the tertiary structure of RNAase S is formed, as monitored by tyrosine absorbance. At this stage in folding, the isomerization reaction that creates the major slow-folding species has not yet been reversed. Our result for the 3-13 helix is consistent with the finding of Labhardt (1984), who has studied the kinetics of folding of RNAase S at 32 degrees C by fast circular dichroism. He finds the dichroic change expected for formation of the 3-13 helix occurring when the tertiary structure is formed. Protected amide protons are found in the S-protein moiety earlier in folding. Formation or stabilization of this folding intermediate depends upon S-peptide: the intermediate is not observed when S-protein folds alone, and folding of S-protein is twice as slow in the absence of S-peptide. Although S-peptide combines with S-protein early in folding and is needed to stabilize an S-protein folding intermediate, the S-peptide helix does not itself become stable until the tertiary structure of RNAase S is formed.  相似文献   

6.
7.
A method is described which extends the structural resolution of the usual hydrogen-exchange experiment by quantifying the exchange kinetics from known regions of a protein. In the usual out-exchange experiment with tritium-labeled protein, all exchange events are simultaneously monitored by measuring total protein-bound radioactivity at various specified times. In the present procedure, the protein is adjusted from the out-exchange conditions to pH 2.8 at 8 °C, immediately digested with an acid protease and the digest run on a high pressure column at the same temperature and pH. The specific activities of the individual peptide peaks are then determined. The entire analytical process requires 20 to 30 minutes depending on the position of the peptide in the chromatogram. Since the peptides are fully exposed to solvent during the analysis, this time corresponds to several half-lives of exchange. However, with sufficient isotope in the starting material large amounts of radioactivity remain associated with each peptide fragment allowing accurate analyses. With care, the digestion and separation can be made very reproducible.The procedure was tested on the ribonuclease S system using labeled S-peptide (providing an extension of the observations of Schreier &; Baldwin, 1976). At pH 2.8 and pH 4.2 free S-peptide exchanges at rates which agree quite well with the values predicted by the data of Molday et al. (1972). In complex with S-protein, the S-peptide protons are not all protected to the same extent. For residues 7 through 13, 7 and 8 are more highly protected than 13, while 10 and 11 are essentially unaffected by complex formation. The model based on the X-ray structure determination indicates that all of these residues are part of an α-helical segment in the chain.  相似文献   

8.
The major unfolded form of ribonuclease A is known to show well-populated structural intermediates transiently during folding at 0°–10°C. We describe here how the exchange reaction between D2O and peptide NH protons can be used to trap folding intermediates. The protons protected from exchange during folding can be characterized by 1H-nmr after folding is complete. The feasibility of using 1H-nmr to resolve a set of protected peptide protons is demonstrated by using a specially prepared sample of ribonuclease S in D2O in which only the peptide protons of residues 7–14 are in the 1H-form. All eight of these protected peptide protons are H-bonded. Resonance assignments made on isolated peptides containing these residues have been used to identify the protected protons. Other sets of protected protons trapped in the 1H-form can also be isolated by differential exchange, using either ribonuclease A or S. Earlier model compound studies have indicated that H-bonded folding intermediates should be unstable in water unless stabilized by additional interactions. Nevertheless, peptides derived from ribonuclease A that contain residues 3–13 do show partial helix formation in water at low temperatures. We discuss the possibility that specific interactions between side chains can stabilize short α-helixes by nucleating the helix, and that specific interactions may also define the helix boundaries at early stages in folding.  相似文献   

9.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

10.
A. M. Labhardt 《Biopolymers》1981,20(7):1459-1480
The thermal-denaturation transition of ribonuclease S (RNAase S) is measured by circular dichroism at 225 nm. Only conformational transitions involving the S-peptide–S-protein complex are detected at this wavelength. Different pathways of thermal unfolding at high and low concentrations are apparent: at low concentrations the temperature of half-completion of denaturation (Tm) varies with concentration. Above a total enzyme concentration of 50 μM, Tm remains constant. The observed data can be explained on the basis of a model where the association–dissociation step occurs between S-peptide and thermally (at least partly) unfolded S-protein. The complex as a whole undergoes a major folding–unfolding transition in the course of which the S-peptide μ-helix appears to be formed. The unfolded complex is well populated in the unfolding transition region for enzyme concentrations of 100 μM or more. The model succeeds in deducing thermodynamic parameters from the thermal denaturation curves in various different ways. The values thus obtained are fully self-consistent and, moreover, consistent with the values for the apparent association constant and apparent association enthalpy as measured in enzyme-dilution experiments and by batch calorimetry.  相似文献   

11.
Isotopic exchange kinetics at equilibrium for E. coli native aspartate transcarbamylase at pH 7.8, 30 °C, are consistent with an ordered BiBi substrate binding mechanism. Carbamyl phosphate binds before l-Asp, and carbamyl-aspartate is released before inorganic phosphate. The rate of [14C]Asp C-Asp exchange is much faster than [32P]carbamyl phosphate Pi exchange. Phosphate, and perhaps carbamyl phosphate, appears to bind at a separate modifier site and prevent dissociation of active-site bound Pi or carbamyl phosphate. Initial velocity studies in the range of 0–40 °C reveal a biphasic Arrhenius plot for native enzyme: Ea (>15 °C) = 6.3 kcal/ mole and Ea (<15 °C) = 22.1 kcal/mole. Catalytic subunits show a monophasic plot with Ea ? 20.2 kcal/mole. This, with other data, suggests that with native enzyme a conformational change accompanying aspartate association contributes significantly to rate limitation at t > 15 °C, but that catalytic steps become definitively slower below 15 °C. Model kinetics are derived to show that this change in mechanism at low temperature can force an ordered substrate binding system to produce exchange-rate patterns consistent with a random binding system with all exchange rates equal. The nonlinear Arrhenius plot also has important consequences for current theories of catalytic and regulatory mechanisms for this enzyme.  相似文献   

12.
Unspecific binding of a protamine, namely fluorescein-labelled clupeine Z, to double-stranded calf thymus DNA was studied using fluorescence titration methods and chemical relaxation techniques. Both equilibrium and kinetic data have been analysed using general theoretical approaches discussed in the accompanying paper. The results agree well with the predictions made on the basis of a standard co-operative binding model.Basic parameters evaluated are the co-operative binding constant (K), the coefficient measuring co-operative interaction between nearest neighbours (q), the number of nucleotides occupied by one protamine molecule (n) and the rate constant of dissociation at the ends of bound ligand sequences (KD). Values obtained at 20 °C, pH 7.5 and 0.4 m-NaCl were K = 5.8 × 107m?1, q = 1700, n = 20 and KD = 0.29 s?1. They have been found to be sensitive to the concentration of added salt (NaCl). This effect apparently reflects the essentially electrostatic nature of the binding process. The results can be satisfactorily described in terms of competitive binding of sodium ions.  相似文献   

13.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

14.
The cytosolic estrogen receptor (calf uterus) bound to estradiol (E2) at 0°C changes from a state with fast into a state with slow E2 dissociation rates when placed at 28°C. This temperature accelerated transition in receptor affinity for its ligand takes place within 10 min at 28°C. Similarly, receptor bound to estriol (E3) at 0°C changes, when heated, from a state with fast into a state with slow E3 dissociation. The main difference between RE2 and RE3 was that E3 dissociates from unheated 8S RE3 and heat-transformed 5S RE3 at a much faster rate than E2 from RE2;In the mature ovariectomized rat a slow dissociating 5S receptor estrogen complex is found in nuclei 1 h after injection of [3H]E2 or [3H]E3. In vitro dissociation of these 2 estrogens from this nuclear bound receptor formed in vivo takes place at rates similar to those from heat-transformed cytosolic RE2 or RE3 complexes.Addition of pyridoxal 5'-phosphate (PLP) to the slow-dissociating heat-transformed 5S estrogen receptor complexes causes rapid dissociation of E2 or E3; this effect is dose-dependent and is not due to disruption of 5S dimers, since after PLP addition RE2; and RE3 sediment unchanged as 5S dimers.The presence of a large excess of non-radioactive 4S RE3 does not interfere with the temperature induced rapid transition of 4S R[3H]E2 complexes from the state with fast into a state with slow E2 dissociation kinetics.A model is presented to explain the temperature induced biphasic estrogen dissociation from the receptor. It is proposed that the low affinity 4S RE2 monomer undergoes a temperature and estrogen dependent conformation change, such that the ligand is “locked” into the receptor's binding site. This conformational change results in the formation of a high affinity 4S monomer from which estrogen dissociates at a slower rate. This reaction is independent from subsequent 4S to 5S dimerization (transformation). The different rates of ligand dissociation from the low and high affinity 4S receptors reflect the different interactions (hydrophobic and hydrogen bonding) of E2 and E3 with the estrogen binding domain.  相似文献   

15.
The observed equilibrium constants (Kobs) for the reactions of d-2-phosphoglycerate phosphatase, d-2-Phosphoglycerate3? + H2O → d-glycerate? + HPO42?; d-glycerate dehydrogenase (EC 1.1.1.29), d-Glycerate? + NAD+ → NADH + hydroxypyruvate? + H+; and l-serine:pyruvate aminotransferase (EC 2.6.1.51), Hydroxypyruvate? + l-H · alanine± → pyruvate? + l-H · serine±; have been determined, directly and indirectly, at 38 °C and under conditions of physiological ionic strength (0.25 m) and physiological ranges of pH and magnesium concentrations. From these observed constants and the acid dissociation and metal-binding constants of the substrates, an ionic equilibrium constant (K) also has been calculated for each reaction. The value of K for the d-2-phosphoglycerate phosphatase reaction is 4.00 × 103m [ΔG0 = ?21.4 kJ/mol (?5.12 kcal/mol)]([H20] = 1). Values of Kobs for this reaction at 38 °C, [K+] = 0.2 m, I = 0.25 M, and pH 7.0 include 3.39 × 103m (free [Mg2+] = 0), 3.23 × 103m (free [Mg2+] = 10?3m), and 2.32 × 103m (free [Mg2+] = 10?2m). The value of K for the d-glycerate dehydrogenase reaction has been determined to be 4.36 ± 0.13 × 10?13m (38 °C, I = 0.25 M) [ΔG0 = 73.6 kJ/mol (17.6 kcal/mol)]. This constant is relatively insensitive to free magnesium concentrations but is affected by changes in temperature [ΔH0 = 46.9 kJ/mol (11.2 kcal/mol)]. The value of K for the serine:pyruvate aminotransferase reaction is 5.41 ± 0.11 [ΔG0 = ?4.37 kJ/mol (?1.04 kcal/mol)] at 38 °C (I = 0.25 M) and shows a small temperature effect [ΔH0 = 16.3 kJ/ mol (3.9 kcal/mol)]. The constant showed no significant effect of ionic strength (0.06–1.0 m) and a response to the hydrogen ion concentration only above pH 8.5. The value of Kobs is 5.50 ± 0.11 at pH 7.0 (38 °C, [K+] = 0.2 m, [Mg2+] = 0, I = 0.25 M). The results have also allowed the value of K for the d-glycerate kinase reaction (EC 2.7.1.31), d-Glycerate? + ATP4? → d-2-phosphoglycerate3? + ADP3? + H+, to be calculated to be 32.5 m (38 °C, I = 0.25 M). Values for Kobs for this reaction under these conditions and at pH 7.0 include 236 (free [Mg2+] = 0) and 50.8 (free [Mg2+] = 10?3m).  相似文献   

16.
Bovine brain hexokinase enhances the effect of Mn(II) on the longitudinal relaxation rate of water protons. Direct interaction of Mn(II) with the enzyme has been studied using electron spin resonance and proton relaxation rate enhancement methods. The results indicate that brain hexokinase has 1.05 ± 0.13 tight binding sites and 7 ± 2 weak binding sites with a dissociation constant, KD = 25 ± 4 μM and KD = 1050 ± 290 μM, respectively, at pH 8.0, 23 °C. The characteristic enhancement ?b) for hexokinase-Mn(II) complex evaluated from proton relaxation rate enhancement studies, gave ?b = 3.5 ± 0.4 for tight binding sites and an average ?b = 2.3 ± 0.5 per site for weak binding sites at 9 MHZ. The dissociation constant of Mn(II) for tight binding sites on the enzyme exhibits strong temperature dependence. In the low-temperature region (5–12 °C) brain hexokinase probably undergoes a conformational change. Frequency dependence of the normalized relaxation rate for bound water at various temperatures has shown that the number of exchangeable water molecules left in the first coordination sphere of bound Mn(II) is about one at 30 °C and about two at 18 °C. Binding of glucose 6-phosphate to hexokinase results in large-line broadening of the resonances of anomeric protons of the sugar. However, no such effect was observed in the case of glucose binding. These results suggest different modes of interaction of these two sugars to hexokinase. Line broadening of the C-(1) hydrogen resonances of glucose caused by Mn(II) in the presence of hexokinase suggests the proximity of the Mn(II) binding site to that of glucose. A lower limit of 1330 ± 170 s?1 for the rate of dissociation of glucose from enzyme-Mn(II)-glucose complex has been obtained from these studies.  相似文献   

17.
Subtilisin modification of monodeamidated ribonuclease-A   总被引:1,自引:1,他引:0       下载免费PDF全文
Limited proteolysis of RNAase-Aa1 (monodeamidated ribonuclease-A) by subtilisin results in the formation of an active RNAase-S type of derivative, namely RNAase-Aa1S. RNAase-Aa1S was chromatographically distinct from RNAase-S, but exhibited very nearly the same enzymic activity, antigenic conformation and susceptibility to trypsin as did RNAase-S. Fractionation of RNAase-Aa1S by trichloroacetic acid yielded RNAase-Aa1S-protein and RNAase-Aa1S-peptide, both of which are inactive by themselves, but regenerate active RNAase-Aa1S′ when mixed together. RNAase-Aa1S-peptide was identical with RNAase-S-peptide, whereas the protein part was distinct from that of RNAase-S-protein. Titration of RNAase-Aa1S-protein with S-peptide exhibited slight but noticeably weaker binding of the peptide to the deamidated S-protein as compared with that of native protein. Unlike the subtilisin digestion of RNAase-A, which gives nearly 100% conversion into RNAase-S, the digestion of RNAase-Aa1 gives only a 50% conversion. The resistance of RNAase-Aa1 to further subtilisin modification after 50% conversion is apparently due to the interaction of RNAase-Aa1 with its subtilisin-modified product. RNAase-S was also found to undergo activity and structural changes in acidic solutions, similar to those of RNAase-A. The initial reaction product (RNAase-Sa1) isolated by chromatography was not homogeneous. Unlike the acid treatment of RNAase-A, which affected only the S-protein part, the acid treatment of RNAase-S affected both the S-protein and the S-peptide region of the molecule.  相似文献   

18.
In this work different aspects of the glucose-fructose enzymatic isomerization, using immobilized glucose isomerase, are studied and quantified. Reaction temperatures range from 40?°C to 60?°C. Intra-particle effective diffusivities (D e), determined after uptake experiments, are between 1.20?×?10?6?cm2/s, at 40?°C, and 2.52?×?10?6?cm2/s, at 60?°C. The estimated energy of activation for diffusion (E aD) is 7.71?kcal/mol. No significant adsorption of the sugars on the support gel matrix is observed. Crushed particles (φ = 150–350?μ) are used during kinetic experiments. For this range of particle diameters, inherent kinetics is approached. A reversible Michaelis–Menten rate equation is fitted to the data, providing the following parameters at pH = 7.0: k 0 = 2.15?×?10?6?g/IU/s; E a/R = 8998?K. Glucose (K G) and fructose (K F) affinity constants are essentially the same, ranging from 0.190?M, at 40?°C to 0.305?M, at 60?°C. The thermodynamic equilibrium constant is determined for the three temperatures, and the heat of reaction, estimated from a Van't Hoff plot, is ΔH = 1682?cal/mol. Independent experiments, where the reaction occurs in the presence of significant intra-particle mass transfer resistance, are used as validation tests.  相似文献   

19.
A reexamination of the equilibrium and the kinetics of 1,25-dihydroxy vitamin D3 binding with its receptor in chick intestinal cytosol was performed because of the recent availability in our laboratory of high specific activity 1,25-dihydroxy[3H-26,27]vitamin D3 (160 Ci/mmol). Under saturating conditions at 25 °C, Scatchard analysis revealed an equilibrium dissociation constant (Kd) of 7.1 × 10?11m which is several fold lower than previously reported for this binding reaction. Furthermore, an estimate of 1.8 × 103 receptor sites per cell was obtained from the intercept of the line with the abscissa of the Scatchard plot. From a kinetic analysis of 1,25-dihydroxy vitamin D3 binding with chick intestinal cytosol, association and dissociation rate constants were determined. Values that were obtained at 25 °C for these processes were 9.5 × 108m? min? and 7.1 × 10?3 min?, respectively. Although these studies, such as for other steroid hormones, were carried out using a crude native cytosol preparation, we have been able to demonstrate unequivocally through the use of high specific activity 1,25-dihydroxy[3H-26,27] vitamin D3 a truly high affinity binding site.  相似文献   

20.
Glyceraldehyde 3-P dehydrogenase was purified approximately 250-fold from pig liver and crystallized. The purification procedure consisted of treating liver homogenates with zinc chloride, followed by ammonium sulfate fractionation and ion exchange chromatography. The enzyme was monodisperse in the ultracentrifuge with a sedimentation coefficient of s20,w = 7.85 S. Sodium dodecyl sulfate polyacrylamide gel electrophoresis showed a single subunit band with an approximate molecular weight of 38,000. High-speed sedimentation equilibrium gave a molecular weight of 1.5 × 105. Incubation of the enzyme with ATP at 0 °C caused a loss of its dehydrogenase activity; some of the lost activity was regained upon warming to room temperature. Sucrose density gradient studies of the ATP-treated enzyme revealed a decrease in its sedimentation coefficient from 7.8 to 3.85 S. In the forward reaction direction, the Km for glyceraldehyde 3-P was 240 μm and the Km for NAD was 12 μm. In the backward reaction direction, the Km for NADH was 23 μm and the Ki for NAD was 850 μm. Pig liver glyceraldehyde-3-P dehydrogenase resembles the rabbit muscle enzyme in that it apparently contains 2 to 3 mol of tightly bound NAD. However, it differs strongly from that enzyme in its rate and extent of inactivation by ATP at 0 °C and by urea; the pig liver enzyme, like the yeast enzyme, dissociates much more slowly and much less completely than the rabbit muscle enzyme under comparable conditions.  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