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 286 毫秒
1.
This paper describes a calorimetric investigation of the effect of substitution of methyl and bromine groups at the 5-position of the cytosine ring upon the heats of reaction accompanying the formation of the different 1:1 complexes between these polyribocytidylic acid derivatives and poly(rI). The ΔH of reaction to form poly(rBrC) · poly(rI) at 25 and 37 °C was found to be − 7200 and − 7755 cal./mole of base pairs, respectively. This value is about 2 kcal./mole of base pairs greater than that found for poly(rC) · poly(rI); thus, the bromine substituent produces a large enthalpic stabilization. The ΔH's for poly r(C,BrC) copolymers reacting with poly(rI) were found to vary linearly with bromine content over the entire range of copolymer composition, indicating the absence of any interactions depending upon the sequence of bromine-containing residues. The ΔH of reaction to form poly(rMeC) · poly(rI) was found to be about − 6300 cal./mole of base pairs at 25 and 37 °C. This value represents an enthalpic stabilization relative to poly(rC) · poly(rI) of 700 cal./mole of base pairs arising from the presence of the methyl group. These results clearly show that different types of nucleic acid base pairs can have different enthalpies and entropies characterizing their interaction.  相似文献   

2.
The existence of an extracellular α-amylase, which is secreted by virus tumors from the roots of Rumex acetosa grown in vitro, has been demonstrated and the properties of the enzyme have been studied. As far as the authors are aware, this is the first enzyme demonstrated to be secreted from intact cells of higher plants.An assay for the quantitative determination of this enzyme is presented. This assay is based on the spectrophotometric determination of the decrease in color of the starch-iodine complex due to the activity of the amylase on the starch substrate.Among the factors discussed which affect the density of the starch-iodine complex are iodine concentration, temperature, and ionic strength.The optimum pH for activity of this enzyme is 4.6, both in acetate and citrate-phosphate buffers.At pH 4.6 in 0.02 M acetate buffer, the Q10 for the hydrolysis of starch by the enzyme is 1.6 in the range of 20–40 °C. The activation energy in this range calculated by the Arrhenius equation is 8000 cal./ mole.This α-amylase is protected by calcium and is sensitive to low pH values as are the cereal α-amylases.The energy of activation for the heat denaturation of the enzyme is 43,000 cal./mole, as calculated by the Arrhenius equation.  相似文献   

3.
We have measured the 31P n.m.r. spectra of NADP+ and NADPH in their binary complexes with Escherichia coli dihydrofolate reductase and in ternary complexes with the enzyme and folate or methotrexate. The 31P chemical shift of the 2′ phosphate group is the same in all complexes; its value indicates that it is binding in the dianionic state and its pH independence suggests that it is interacting strongly with cationic residue(s) on the enzyme. Similar behaviour has been noted previously for the complexes with the Lactobacillus casei enzyme although the 31P shift is somewhat different in this complex, possibly due to an interaction between the 2′ phosphate group and His 64 which is not conserved in the E. coli enzyme. For the coenzyme complexes with both enzymes 31POC21H2′ spin-spin interactions were detected (7.5–7.8 Hz) on the 2′ phosphate resonances, indicating a POC2H2′ dihedral angle of 30 or 330 : this is in good agreement with the value of 330° measured in crystallographic studies1 (Matthews et al., 1978) on the L. casei enzyme. NADPH-MTX complex. The pyrophosphate resonances are shifted to different extents in the various complexes and there is evidence that there is more OPO bond angle distortion in the E. coli enzyme complexes than in those with the L. casei enzyme. The effects of 31POC51H5′ spin coupling were detected on one pyrophosphate resonance and indicate that the POC5H5′ torsion angle has changed by at least ~30° on binding to the E. coli enzyme: this is considerably less than the distortion (~50°) observed previously in the L. casei enzyme complex.  相似文献   

4.
Casein kinase II and ornithine decarboxylase were purified from a virally-transformed macrophage-like cell line, RAW264. The addition of casein kinase II to a reaction mixture containing [tau-32P]GTP, Mg++, and ornithine decarboxylase led to the phosphorylation of a 55,000 dalton protein band in the purified preparation of ornithine decarboxylase. Stoichiometric estimates indicated that casein kinase II incorporated 0.15 mole of phosphate per mole of ornithine decarboxylase, which was increased to 0.3 mole/per mole in the presence of spermine. The apparent Km and Vmax values for the casein kinase II-mediated phosphorylation of ornithine decarboxylase were 0.36 microM and 62.5 nmol/min./mg kinase. The addition of spermine to the reaction did not alter the Km but increased the Vmax to 100 nmol/min./mg kinase. The phosphorylation of ornithine decarboxylase by casein kinase II affected neither the rate of maximal ornithine decarboxylase activity nor the affinity of the enzyme for ornithine.  相似文献   

5.
1. L-asparaginase from M. phlei was purified about 170-fold with an 11% yield. The purification procedure consisted of: fractionation with ammonium sulphate; adsorption of contaminating proteins on calcium phosphate gel; chromatography on Sephadex G-150 and DEAE-cellulose. The specific activity of the final preparation was 32.6 i.u./mg protein. 2. Molecular weight of the enzyme as determined by Sephadex G-100 filtration amounted to 126 000. Optimum pH was 8.8-9.2. The enzyme did not hydrolyse L-glutamine over the pH range 4-9, and was inhibited by D-asparagine. The apparent Michaelis constant for L-asparagine was 0.7 mM; energy of activation, 9800 cal/mole. 3. On polyacrylamide-gel electrophoresis the final preparation revealed two protein bands, one of which was coincident with the enzyme activity.  相似文献   

6.
1. delta-Aminolaevulate dehydratase (EC 4.2.1.24) was purified 80-fold from tobacco leaves and its properties were studied. 2. The enzyme had optimum pH7.4 in potassium phosphate buffer, K(m)6.25x10(-4)m at 37 degrees and pH7.4, optimum temperature 45 degrees and an activation energy of 11100 cal./mole. 3. The enzyme lost activity when prepared in the absence of cysteine, and this activity was only partly restored by the later addition of thiols. Reagents for thiol groups inactivated the enzyme. 4. Mg(2+) was essential for activity, and EDTA and Fe(2+) were inhibitory; Mn(2+) was an activator or an inhibitor depending on the concentration.  相似文献   

7.
An Mn2+-activated phosphoprotein phosphatase of Mr = 80,000 from rabbit muscle catalyzes the dephosphorylation of skeletal muscle proteins that are phosphorylated by either phosphorylase kinase or cAMP-dependent protein kinase. Phosphorylase or glycogen synthase labeled by phosphorylase kinase at seryl residues 14 or 7, respectively, are both dephosphorylated by the phosphatase. Phosphorylase a and glycogen synthase compete with one another for the phosphatase. The phosphatase discriminates between different sites labeled by the cAMP-dependent protein kinase: glycogen synthase phosphorylated either to 1.0 or 1.8 mol phosphate/mol, or phosphorylase kinase phosphorylated on its β-subunit serve as substrates for the phosphatase, but the phosphorylase kinase α-subunit, the phosphorylated phosphatase inhibitor 1, or casein do not. Histone fraction IIA, phosphorylated by the catalytic subunit, was a poor substrate even at a concentration of 100 μm. Phosphorylation of the α-subunit of phosphorylase kinase had no influence on the kinetics of dephosphorylation of the β-subunit. Thus, the Mr = 80,000 phosphatase meets the functional definition of a protein phosphatase 1 [Cohen, P. (1978) Curr. Top. Cell. Regul.14, 117–196]. Furthermore, from a comparison of the known phosphorylated sites of these proteins, it appears that the phosphatase discriminates between different sites present in the phosphoproteins tested on the basis of the Km values for the reactions. It displays a preferential activity toward proteins with a primary structure wherein basic residues are two positions amino-terminal from the phosphoserine, AgrLysX-YSer(P) or LysArgX-YSer(P), rather and one residue away, ArgArgX-Ser(P).  相似文献   

8.
Carter OG  Lathwell DJ 《Plant physiology》1967,42(10):1407-1412
The uptake of orthophosphate (32P) by excised corn roots, Zea mays L. was studied using roots grown on 0.2 mm CaSO4. Nine concentrations of KH2PO4 from 1 to 256 μm were used at temperatures of 20°, 30°, and 40°. Enzyme kinetic analysis was applied to the data obtained. Two apparent mechanisms (sites) of phosphate uptake were observed, 1 dominating at high P concentrations and 1 at low P concentrations. A Km of 1.36 × 10−4 and a Vmax of 177 × 10−9 moles per gram of roots per hour at 30° was calculated for the mechanism dominating at high P concentrations. Similar calculations gave a Km of 6.09 × 10−6 and a Vmax of 162 × 10−9 moles per gram of roots per hour at 30° for the mechanism dominating at low P concentrations. The Q10 for both mechanisms was approximately 2. Calculation of thermodynamic values from the data gave ΔF of − 5200 cal, ΔH of − 950 to − 1400 cal, and a enthalpy of activation (A) of 10,300 to 13,800 cal per mole for the mechanism dominating at high P concentrations. Similar calculations from the data for the mechanism dominating at low P concentrations gave a ΔF of − 7300 cal, ΔH of − 10,700 to − 8200 cal, and a A of 9300 to 18,900 cal per mole. If the dual mechanism interpretation of this kind of data adequately describes this system, then both mechanisms of P absorption by corn roots involve chemical reactions.  相似文献   

9.
400 MHz 1H-resonances of 1-phosphatidyl-myoinositol (PI), PI 4-phosphate (DPI) and PI 4,5-bisphosphate (TPI) in CD3OD were assigned. Proton resonances in the inositol moiety shift downfield with the increase in the number of the phosphate from PI to TPI. From the 1H-1H and 1H-31P vicinal coupling constants, rotamer populations around bonds in the polar headgroups were calculated. The H-C(5)O-P bond at position 5 of the inositol moiety tends to assume a gauche form. The H-C(1)O-P and the H-C(4)O-P bonds are not so strongly restricted to the gauche form as the H-C(5)O-P bond. The conformation of the glycerol moiety in PI, DPI and TPI is similar to that in phosphotidylcholine (PC) and phosphatidylethanolamine (PE). The CH-CH2O-P bond in the glycerol moiety assumes a trans form. The acyl chains prefer a gauche arrangement to each other around the CHCH2OCOR bond.  相似文献   

10.
1. The reaction between avidin and biotin was found to be exothermic, ΔH being −20·3kcal./mole of biotin bound. The corresponding value of ΔH for streptavidin was −23kcal./mole. 2. The heat evolved was independent of the pH (between 5 and 9), of the buffer (borate or ammonia) and of the fractional saturation of the avidin with biotin. 3. The entropy change for the reaction was zero, and it is suggested that the entropy increase to be expected from hydrophobic interactions was counterbalanced by a decrease in entropy accompanying the formation of buried hydrogen bonds. 4. Modification of the potential hydrogen-bonding sites of the imidazolidone ring led to a decreased heat output and a positive entropy of reaction.  相似文献   

11.
The four human Duchenne dystrophic isoenzymes (M-M, M-B, B-B, from the muscle and B-B from the brain) of ATP-creatine transphosphorylase (S. A. Kuby, H. J. Keutel, K. Okabe, H. K. Jacobs, F. Ziter, D. Gerber, and F. H. Tyler, 1977, J. Biol. Chem.252, 8382–8390) have now been compared physically and chemically with their normal human counterparts (viz., with the three isoenzymes, M-M, M-B, B-B, 2). All isoenzymes proved to be composed of two noncovalently linked polypeptide chains, by sedimentation equilibrium analyses in the presence and absence of disruptive agents. In the presence of 2-mercaptoethanol at 0.16(Γ/2), pH 7.8, the two native muscle types yielded identical values for s20,w, concentration dependencies, and molecular weight, and similarly for the brain types (from the brain). But the human brain type proved to be slightly heavier than the muscle type (viz. 88,400 vs 85,900). All of the isoenzymes showed similar electrophoretic behavior between their several counterparts between pH 5–8, except perhaps between pH 8–10, where small differences appeared. The three native normal human isoenzymes, as well as the dystrophic human isoenzymes (M-M from the muscle and B-B from the brain) all contain 2 reactive sulfhydryl groups per mole or 1 per polypeptide chain of these two-chain proteins, which may be titrated with 5,5′-dithiobis(2-nitrobenzoic acid) (Nbs2); and under acidic conditions, quantitative titrations with 4,4′-dithiodipyridine yield a total of 10 -SH groups per mole of each brain type and 8 -SH groups per mole of muscle type, in the case of man, dystrophic man, calf, and rabbit. The kinetics of reactions between Nbs2 and the sulfhydryl groups of all three normal human isoenzymes and two dystrophic human isoenzymes have been measured under several sets of denaturing conditions. A comparison of their reactive calculated second-order velocity constants reveal significant differences between these three normal human isoenzymes, but the ksecond order values for the reactions of the sulfhydryl groups of the dystrophic M-M and B-B with Nbs2, when compared with their normal counterparts, gave identical values in the presence of 7.3 m urea or 1.8% laurylsulfate, from which it may be inferred that very similar, if not identical, environments surround these two sets of sulfhydryl groups. A comparison of the amino acid compositions of the normal human muscle type and brain type with the human dystrophic M-M and B-B (from the brain) reveal essentially identical values for the muscle types but nearly identical values for the brain types, with a few differences. Their respective tryptic peptide maps have been compared of the S-carboxy-methylated proteins (alkylated with iodo[2-14C]acetic acid at the two exposed -SH groups per mole). Thus, the muscle types, normal and dystrophic, yield identical maps, but the brain types nearly identical maps, with a few significant differences. Isolation of the tryptic tridecapeptide from the S-carboxymethylated normal human and dystrophic human dimeric muscle-type ATP-creatine transphosphorylases, labeled at the single exposed SH group per polypeptide chain with iodo[2-14C]acetate, yielded the following sequence for both proteins: ValLeuThrCys(CH2COOH)ProSerAsnLeuGlyThr GlyLeuArg [where Cys(CH2COOH) is S-carboxymethyl cysteine]. This sequence showed remarkable homology with a few other equivalent peptides reported to be derived from the exposed SH group of other ATP-creatine transphosphorylases. In conclusion, there does not appear to be a mutation in the structural genes for the muscle-type creatine kinases detectable by the analyses presented here. However, the brain types warrant further investigation.  相似文献   

12.
The equivalence of the four dihydroxyacetone phosphate binding sites of aldolase was abolished by lowering the temperature. At pH 6.2 and ?13 2C, four binding sites were detected by gel filtration; two sites with a Kdiss ?0.1 μm, and a second set of sites with a Kdiss = 4 μm. The alteration of the binding was accompanied by the alteration of the catalytic activity. The low-affinity sites were incapable of catalyzing the cleavage of the (3S) CH bond of dihydroxyacetone phosphate, and form only the ketimine phosphate intermediate. The high-affinity sites were still able to cleave the (3S) CH bond of dihydroxyacetone phosphate; however, the eneamine phosphate intermediate formed was almost fully converted into the eneamine-aldehyde … phosphate intermediate, which was the prevailing species at the equilibrium. The mechanism of the half-of-the sites reactivity of aldolase at low temperature has been explained and the nonequivalence of sites in promoting catalysis has been utilized to dissect and characterize the individual partial reactions of the enzyme. In the course of these studies it has been shown that the rate of hydration-dehydration of dihydroxyacetone phosphate at ?24 °C was too slow to measure.  相似文献   

13.
《Inorganica chimica acta》1988,144(2):205-211
Two (1:2) silver monophosphine complexes have been studied by X-ray diffraction methods and in solution by P NMR spectroscopy. Both are monomeric and tricoordinated in the solid state but one of them, the perchlorate compound, is probably associated as a dimer species in solution from the lower 1J(107Ag31P) value when compared to the nitrate analogue. Previous structural correlations found in other silver-phosphine complexes have been confirmed for these new compounds. Thus, larger PAgP bond angles are associated with shorter AgP bond distances, longer Aganion bond distances and lower Lewis basicity of the anions. Selected structural data are: PAgP bond angle of 139.04(9)°, AgP bond lengths of 2.440(3) and 2.445(3) Å for the nitrate complex and 147.34(3)°, 2.429(1) Å and 2.432(1) Å, for the perchlorate one. J(107Ag31P) is 457 Hz and 447 Hz, respectively. The complexes are triclinic, Z = 2, with the parameters: a = 9.258(2), b = 9.828(2), c = 23.385(5) Å, α = 94.73(2)°, β = 96.35(2)°, γ = 116.42(1)° (nitrate) and a = 9.505(2), b = 9.790(2), c = 23.667(6) Å, α= 99.03(2) β = 95.44(2) γ = 115.97(1)° (perchlorate).  相似文献   

14.
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   

15.
An enzyme, apyrase, obtained from thermophile No. 2184 possesses heat stability at 65 °C., the temperature preferred by the organism for growth.Apyrase from the thermophile has an activation energy of 9600 cal./ mole. Its Michaelis constant is 0.00065 at 60 °C. The enzyme is inhibited by fluoride, citrate, and copper, but is activated by magnesium.Apyrase from a mesophilic bacterium, although possessing a fair amount of stability in the presence of substrate, does not have the resistance to heat possessed by the thermophile enzyme.The apyrase from potato is inactivated readily on heating in the absence of adenosine triphosphate.  相似文献   

16.
The denaturation of a macromolecule such as a polypeptide is considered in the case where the total number of noncovalent internal bonds broken in the binding process of the chemical agent is only a fraction, αs, of the total number of noncovalent internal bonds involved in its helical native conformation. Starting from a two-parameter (s, σ) model of the helix–random coil transition of polypeptides, the transition temperature is derived as a function of the fraction αs, and of the concentration and the binding constant of the chemical agent. The lower limit of the transition temperature and the corresponding slope of the transition curve are shown to depend on αs. As an illustration, existing data on the partial acid denaturation of the enzymes ribonuclease A and muromidase are analyzed. The resulting average enthalpies of a noncovalent bond in the native ribonuclease A and muromidase are found to be 1.81 kcal/mole and 1.54 kcal/mole, respectively; the corresponding entropies are 5.4 and 4.4 cal/deg./mole, respectively. As a further example, existing data on a partially methylated enzyme ribonuclease A are also considered.  相似文献   

17.
The heat effects accompanying the isothermalin vitro polymerization ofPr. mirabilis flagellin on short flagella fragments (seeds) have been measured in phosphate buffer pH 7, at various temperatures employing a batch microcalorimeter. Additionally, at 20 ?C, measurements have been performed in phosphate as well as Tris- HCl buffer at pH 7.5. The rate of both heat uptake and release during the process of polymerization was shown to be proportional to the rate of molar ellipticity changes observed by parallel circular dichroism experiments. No change in the state of protonation of flagellin occurs during the polymerization as indicated by the constancy of the enthalpy values determined in buffers with different heats of ionization. The apparent molar enthalpy of polymerization at 25 ?C, pH 7, is ?34.7±3 kcal per mole of flagellin, the relatively large error mainly resulting from uncertainties of the determination of the percentage of unpolymerized monomers after completion of the reaction. The most prominent feature of the results obtained in this study is the large temperature variation of the enthalpy, corresponding to a temperature independent heat capacity change ofδc p =?3039±100 cal per degree per mole of flagellin, the error limits referring to the standard deviation in a linear regression analysis.  相似文献   

18.
Sorption and desorption of copper by and from clay minerals   总被引:1,自引:0,他引:1  
Summary The sorption of Cu by different clay minerals from dilute CuSO4 solutions was studied as a function of pH. It was found that Cu sorption increased with increasing the equilibrium pH. The low sorption at low pH values was attributed to the competition effect of H+ ions and the release of octahedral Mg, Fe and Al from the 2: 1 minerals in acid solutions. The higher sorption of copper at higher pH values was related to the absence of H+ ions and to the increase in the pH-dependent negative charge in kaolinite and gibbsite.The desorption of copper using solutions of 1 N NaCl at different pH values indicated that kaolinite and montmorillonite released large amounts of their adsorbed Cu even at high pH values. Vermiculite, on the other hand, exhibited a strong retention of Cu against extraction with NaCl. Oven-drying had no significant effect on the desorption characteristics of the Cu-saturated clays.The Ca-Cu exchange isotherms on montmorillonite were studied at two different pH values. The isotherms indicated a preference of Ca at the lower pH (pH 3.5), with K = 0.931 and G = +41.0 cal./mole. At the higher pH (5.2) the isotherms indicated a preference of Cu with K = 1.282 and G = –141.0 cal./mole. The difference was attributed to the competition of H+ at low pH.  相似文献   

19.
Simultaneous oxidation-reduction potential versus absorbance titrations on intact mitochondria have allowed us to measure the individual spectral and thermodynamic properties of cytochromes a and a3 in the presence or absence of added ligands. In coupled pigeon heart mitochondria, saturated with CO, the reduced a3CO compound can be fully oxidized by ferricyanide under conditions of high phosphate potential, i.e., on addition of ATP. These ATP-induced effects are reversed by the presence of oligomycin or uncouplers with complete restoration of the reduced a3CO compound. This observation has permitted us to examine the spectral and thermodynamic properties of the a3CO compound in some detail. The data indicate that, in the presence of ATP, the midpoint potential of the a3CO compound varies as a function of CO concentration in a thermodynamically predictable fashion. The ATP-induced changes on the a3CO compound are directly analogous to the changes reported earlier for cytochrome a3 in the absence of added ligands.  相似文献   

20.
The consequences of incorporating non-complementary residues into the poly (I) · poly (C) helix have been investigated. Complexes of poly (I) and copolymers of C with different mole-ratios of I, A and U residues have been prepared and denatured in a variety of solvents. The results of both denaturation and analysis of the stoichiometry of the reactions suggest that in poly (I)· poly (C, Ix) complexes, the I residues are excluded from the helix matrix, whereas in the poly (I) · poly (C, Ux) and poly (I) · poly (C, Ax) systems the minor component bases are retained. Preliminaries to a quantitative analysis of the transition data are presented, permitting rough estimates of the difference in stability between poly (I) · poly (C) and poly (I) · poly (U) or poly (I) · poly (A) pairs in these complexes—the results being 1.7 kcal./mole and 1.3 kcal./mole, respectively. The differences in behavior of poly (I) · poly (C, I) complexes are found to be most evident in the presence of 8 m-urea.  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