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 highest productivity (20 IU l(-1) h(-1)) of beta-glucosidase by a mutant of Cellulomonas biazotea was 2.5-fold more than that of the parent organism. The enzyme had a lower activation energy (57 kJ mol(-1)) than the native enzyme (68 kJ mol(-1)). The enzyme from the mutant had enthalpy and entropy values for irreversible intactivation of 95.6 kJ mol(-1) and 60 J.mol(-1) K(-1) compared with 108 kJ mol(-1) and 86 J mol(-1) K(-1) for the native enzyme suggesting that the mutation had stabilized the enzyme.  相似文献   

2.
1. The partial purification of adenosine deaminase, types 1, 2 and 2-1, from human erythrocytes is described. 2. The isoenzyme components characteristic of the three forms of the enzyme were partially resolved by chromatography on DEAE-Sephadex. 3. Gel chromatography of the various forms of the enzyme gave estimates of the molecular weights in the range 30000-35000. 4. Electrophoresis in starch gels containing increasing percentages of starch did not reveal any differences in molecular weight between the genetic variants or their isoenzyme components. 5. Analytical isoelectric-focusing experiments in polyacrylamide gels gave the following pI values for the four isoenzyme components present in type 2-1 erythrocytes: 4.70, 4.83, 4.94 and 5.06. 6. All forms of the enzyme gave K(m) values for adenosine of about 30mum and K(i) values of about 8mum for the competitive inhibitor purine riboside. 7. Reaction rates of the type 1 and 2 enzymes were measured at different temperatures. Both enzymes gave values for the energy of activation for hydrolysis of adenosine of about 33.4kJ/mol (8kcal/mol). 8. Heat inactivation of all forms of the enzyme was markedly dependent on ionic strength, the rate of inactivation increasing with increasing ionic strength. The type 1 and type 2 forms of the enzyme differed significantly in their susceptibility to heat inactivation. From the variation of rates of inactivation with temperature, values were obtained for the energies of activation for the heat inactivation of both enzymes as follows: type 1 enzyme 275.5kJ/mol (65.9kcal/mol) and type 2 enzyme 241.6kJ/mol (57.8kcal/mol.).  相似文献   

3.
The binding of tRNAIIe to isoleucyl-tRNA synthetase in the presence of isoleucine or ATP was investigated using the equilibrium partition method. Isoleucine decreased the affinity of tRNAIIe for the enzyme by a factor of about 5. For the free standard energy of interaction a value of about 1 kcal/mol (4.2 kJ/mol) was calculated. ATP exhibits qualitatively the same effect as isoleucine. A binding of two molecules isoleucine per molecule of enzyme could not be demonstrated even in the presence of ATP and pyrophosphatase.  相似文献   

4.
In Escherichia coli, the phosphorylation and dephosphorylation of isocitrate dehydrogenase (IDH) are catalyzed by a bifunctional protein kinase/phosphatase. We have determined the nucleotide sequence of aceK, the gene encoding IDH kinase/phosphatase. This gene consists of a single open reading frame of 1,734 base pairs preceded by a Shine-Dalgarno ribosome-binding site. Examination of the deduced amino acid sequence of IDH kinase/phosphatase revealed sequences which are similar to the consensus sequence for ATP-binding sites. This protein did not, however, exhibit the extensive sequence homologies which are typical of other protein kinases. Multiple copies of the REP family of repetitive extragenic elements were found within the intergenic region between aceA (encoding isocitrate lyase) and aceK. These elements have the potential for combining to form an exceptionally stable stem-loop structure (delta G = -54 kcal/mol [ca. -226 kJ/mol]) in the mRNA. This structure, which masks the ribosome-binding site and start codon for aceK, may contribute to the downshift in expression observed between aceA and aceK. Another potential stem-loop structure (delta G = -29 kcal/mol [ca. 121 kJ/mol]), unrelated to the REP sequences, was found within aceK.  相似文献   

5.
The interaction of Escherichia coli RNA polymerase with poly[d(A-T)] and poly[d-(I-C)] was studied by difference absorption spectroscopy at temperatures, from 5 to 45 degrees C in the absence and presence of Mg2+. The effect of KCl concentration, at a fixed temperature, was studied from 12.5 to 400 mM. Difference absorption experiments permitted calculation of the extent of DNA opening induced by RNA polymerase and estimation of the equilibrium constant associated with the isomerization from a closed to an open RNA polymerase-DNA complex. delta H0 and delta S0 for the closed-to-open transition with poly[d(A-T)] or poly[d(I-C)] complexed with RNA polymerase are significantly lower than the values associated with the helix-to-coil transition for the free polynucleotides. For the RNA polymerase complexes with poly[d(A-T)] and poly[d(I-C)] in 50 mM KCl, delta H0 approximately 15-16 kcal/mol (63-67 kJ/mol) and delta S0 approximately 50-57 cal/K per mol (209-239 J/K per mol). The presence of Mg2+ does not change these parameters appreciably for the RNA polymerase-poly[d(A-T)] complex, but for the RNA polymerase-poly[d(I-C)] complex in the presence of Mg2+, the delta H0 and delta S0 values are larger and temperature-dependent, with delta H0 approximately 22 kcal/mol (92 kJ/mol) and delta S0 approximately 72 cal/K per mol (approx. 300 J/K per mol) at 25 degrees C, and delta Cp0 approximately 2 kcal/K per mol (approx. 8.3 kJ/K per mol). The circular dichroism (CD) changes observed for helix opening induced by RNA polymerase are qualitatively consistent with the thermally induced changes observed for the free polynucleotides, supporting the difference absorption method. The salt-dependent studies indicate that two monovalent cations are released upon helix opening. For poly[d(A-T)], the temperature-dependence of enzyme activity correlates well with the helix opening, implying this step to be the rate-determining step. In the case of poly[d(I-C)], the same is not true, and so the rate-determining step must be a process subsequent to helix opening.  相似文献   

6.
Cellobiase from Aspergillus niger was glycosylated by covalent coupling to cyanogen bromide activated dextran. The conjugated enzyme retained 62% of the original specific activity exhibited by the native cellobiase. The optimum pH as well as the pH stability of the conjugated form remain almost the same as for the native enzyme. Compared to the native enzyme, the conjugated form exhibited a higher optimal reaction temperature and energy of activation, a higher K(m) (Michaelis constant) and lower Vmax (maximal reaction rate), and improved thermal stability. The thermal deactivation of the native and conjugated cellobiase obeyed the first-order kinetics. The calculated half-life values of heat inactivation at 60, 70 and 80 degrees C was 10.7, 6.25, and 4.05 h, respectively, whereas at these temperatures the native enzyme was less stable (half-life of 3.5, 1.69, and 0.83 h, respectively). The deactivation rate constant at 80 degrees C for the conjugated cellobiase is about 7.9 x 10(-2) h-1, which is lower than that of the native enzyme (36.0 x 10(-2) h-1). The activation energy for denaturation of the native enzyme is about 10.58 kcal/mol, which is 7.25 kcal/mol lower than that of the conjugated enzyme. The effect of different surfactants and some metal ions on the activity of the conjugated cellobiase has been investigated.  相似文献   

7.
Ca2+ (1-5 mM) and lanthanide (20-250 microM) ions enhance the rate of polymerization of purified calf skin collagen (1.5 mg/ml) at pH 7.0 in the presence of 30mM-Tris/HCl and 0.2 M-NaCl. Both the nucleation phase and the growth phase of polymerization are accelerated. The activation energy of the growth phase, 239.3 +/- 24.3 kJ/mol (57.2 +/- 5.8 kcal/mol), is decreased to 145.6 +/- 9.6 kJ/mol (34.8 +/- 2.3 kcal/mol) by 5 mM-Ca2+ and to 75.3 +/- 4.6 kJ/mol (18.0 +/- 1.1 kcal/mol) by 25 microM-Sm3+. In contrast, the activation energy of the nucleation phase, 191.6 +/- 23.4 kJ/mol (45.8 +/- 5.6 kcal/mol), is only slightly decreased by Ca2+ or Sm3+. Collagen fibrils formed in the presence of Sm3+ are thinner than control fibrils, and more thermoresistant.  相似文献   

8.
The gene (nprM) for the highly thermostable neutral protease of Bacillus stearothermophilus MK232 was cloned in Bacillus subtilis using pTB53 as a vector. The nucleotide sequence of nprM and its flanking regions was determined. The DNA sequence revealed only one large open reading frame, composed of 1656 base pairs and 552 amino acid residues. A Shine-Dalgarno (SD) sequence was found 12 bases upstream from the translation start site (ATG). A possible promotor sequence (TTTTCC for the -35 region and TATTGT for the -10 region), which was nearly identical to the promoter for another thermostable neutral protease gene, nprT, was also found about 40 bases upstream of the SD sequence. The deduced amino acid sequence contained a signal sequence in its amino-terminal region. The sequence of the first five amino acids of purified extracellular protease completely matched residues 237-241 of the open reading frame. This suggests that the enzyme is translated as a large polypeptide containing a pre-pro structure as is known for other neutral proteases. The amino acid sequence of the extracellular form of this protease (316 amino acids, molecular mass 34,266 Da) was identical to that of the thermostable neutral protease (thermolysin) from Bacillus thermoproteolyticus except for two amino acid substitutions (Asp37 to Asn37 and Glu119 to Gln119). The G + C content of the coding region of nprM was 42 mol%, while that of the third letter of the codons was lower (36 mol%). This extremely low content is an exceptional case for genes from thermophiles. When the protease genes, nprM and nprT, were cloned on pTB53 in B. subtilis, the expression of nprM was about 20 times higher than that of nprT. The reason for the difference between the two systems is discussed.  相似文献   

9.
The observed equilibrium constants (Kobs) of the P-choline hydrolysis reaction have been determined under physiological conditions of temperature (38 degrees) and ionic strength (0.25 M) and physiological ranges of pH and free [Mg2+]. Using sigma and square brackets to indicate total concentrations: (see article.) The value of Kobs has been found to be relatively insensitive to variations in pH and free [Mg2+]. At pH 7.0 and taking the standard state of liquid water to have unit activity ([H2O] = 1), Kobs = 26.6 M at free [Mg2+] = 0 [epsilon G0obs = -2.03 kcal/mol(-8.48 kJ/mol)], 26.8 M at free [Mg2+] = 10(-3) M, and 28.4 M at free [Mg2+] = 10(-2) M. At pH 8.0, Kobs = 18.8 M at free [Mg2+] = 0, 19.2 M at free [Mg2+] = 10(-3), and 22.2 M at free [Mg2+] = 10(-2) M. These values apply only to situations where choline and Pi concentrations are both relatively low (such as the conditions found in most tissues). At higher concentrations of phosphate and choline, the value of Kobs becomes significantly increased since HPO42- complexes choline weakly (association constant = 3.3 M-1). The value of K at 38 degrees and I = 0.25 M is calculated to be 16.4 +/- 0.3 M [epsilonG0 = 1.73 kcal/mol (-7.23 kJ/mol)]. The K for the P-choline hydrolysis reaction has been combined with the K for the ATP hydrolysis reaction determined previously under physiological conditions to calculate a value of 4.95 X 10(-3 M [deltaG0 j.28 kcal/mol (13.7 kJ/mol] for the K of the choline kinase reaction (EC 2.7.1.32), an important step in phospholipid metabolism: (see article.) Likewise, values for Kobs for the choline kinase reaction at 38 degrees, pH 7.0, and I = 0.25 M have been calculated to be 5.76 X 10(4) [deltaG0OBS = -6.77 KCAL/MOL (-28.3 KJ/mol)] at [Mg2+] = 0; 1.24 X 10(4) [deltaG0obs = -5.82 kcal/mol (-24.4 kJ/mol)] at [Mg2+] = 10(-3) M and 8.05 X 10(3) [delta G0obs = -5.56 kcal/mol (-23.3 kJ/mol)] at [Mg2+ = 10(-2) M. Attempts to determine the Kobs of the choline kinase reaction directly were unsuccessful because of the high value of the constant. The results indicate that in contrast to the high deltaG0obs for the hydrolysis of the ester bond of acetylcholine, the deltaG0obs for the hydrolysis of the ester bond of P-choline is quite low, among the lowest known for phosphate ester bonds of biological interest.  相似文献   

10.
Conditions are described which allow the isolation of rat adipose-cell plasma membranes retaining a large part of the stimulatory effect of insulin in intact cells. In these membranes, the magnitude of glucose-transport stimulation in response to insulin was compared with the concentration of transporters as measured with the cytochalasin-B-binding assay or by immunoblotting with an antiserum against the human erythrocyte glucose transporter. Further, the substrate- and temperature-dependencies of the basal and insulin-stimulated states were compared. Under carefully controlled homogenization conditions, insulin-treated adipose cells yielded plasma membranes with a glucose transport activity 10-15-fold higher than that in membranes from basal cells. Insulin increased the transport Vmax. (from 1,400 +/- 300 to 15,300 +/- 3,400 pmol/s per mg of protein; means +/- S.E.M.; assayed at 22 degrees C) without any significant change in Km (from 17.8 +/- 4.4 to 18.9 +/- 1.4 nM). Arrhenius plots of plasma-membrane transport exhibited a break at 21 degrees C, with a higher activation energy over the lower temperature range. The activation energy over the higher temperature range was significantly lower in membranes from basal than from insulin-stimulated cells [27.7 +/- 5.0 kJ/mol (6.6 +/- 1.2 kcal/mol) and 45.3 +/- 2.1 kJ/mol (10.8 +/- 0.5 kcal/mol) respectively], giving rise to a larger relative response to insulin when transport was assayed at 37 degrees C as compared with 22 degrees C. The stimulation of transport activity at 22 degrees C was fully accounted for by an increase in the concentration of transporters measured by cytochalasin B binding, if a 5% contamination of plasma membranes with low-density microsomes was assumed. However, this 10-fold stimulation of transport activity contrasted with an only 2-fold increase in transporter immunoreactivity in membranes from insulin-stimulated cells. These data suggest that, in addition to stimulating the translocation of glucose transporters to the plasma membrane, insulin appears to induce a structural or conformational change in the transporter, manifested in an altered activation energy for plasma-membrane transport and possibly in an altered immunoreactivity as assessed by Western blotting.  相似文献   

11.
Thermal inactivation and reactivation of pantothenate hydrolase were studied in whole cells of Pseudomonas fluorescens. The enzyme is susceptible to thermal inactivation in whole cells at 37-40 degrees C, and is reactivated when the temperature is lowered again. Chloramphenicol does not prevent reactivation. The activation energy of enzyme inactivation in vivo is about 540kJ/mol. This activation energy is 220kJ/mol in vitro, but it is increased to 550-630kJ/mol by several metabolites, such as succinate, glyoxylate and oxalate. Generally, good carbon sources, causing rapid growth, protect the enzyme from thermal inactivation in vivo, and enable reactivation to occur at a fast rate. The enzyme is also inactivated below 35 degrees C, showing an activation energy of about 35kJ/mol. Good carbon sources prevent this inactivation as well, and cause slight reactivation. Glycine, although not utilized for growth, protects the enzyme well from this inactivation but not from inactivation at 37-40 degrees C, and prevents reactivation totally. From the activation energies of inactivation and the effects of the various carbon sources, it appears possible that changes in the concentrations of intracellular metabolites may be responsible for the changes in inactivation and reactivation.  相似文献   

12.
The subunit composition, metal content, substrate-analogue binding and thermal stability of Aspergillus flavus uricase were determined. A. flavus uricase is a tetramer and contains no copper, iron or any other common prosthetic group. Analytical-gel-filtration and equilibrium-dialysis experiments showed one binding site per subunit for urate analogues. The free energy of xanthine binding was -30.5 kJ (-7.3 kcal)/mol of subunit by equilibrium dialysis and -30.1 kJ (-7.2 kcal)/mol of subunit by microcalorimetry. The enthalpy change for xanthine binding was -15.9 kJ (-3.8 kcal)/mol of subunit when determined from the temperature-dependence of the equilibrium constant and -18.0 kJ (-4.3 kcal)/mol of subunit when measured microcalorimetrically. The thermal inactivation rate of A. flavus uricase increases as protein concentration is decreased. This concentration-dependent instability is not due to subunit dissociation.  相似文献   

13.
14.
The interaction of the synthetic 21 amino acid peptide (AcE4K) with 1-oleoyl-2-[caproyl-7-NBD]-sn-glycero-3-phosphocholine membranes is used as a model system for the pH-sensitive binding of fusion peptides to membranes. The sequence of AcE4K (Ac-GLFEAIAGFIENGWEGMIDGK) is based on the sequence of the hemagglutinin HA2 fusion peptide and has similar partitioning into phosphatidylcholine membranes as the viral peptide. pH-dependent partitioning in the membrane, circular dichroism, tryptophan fluorescence, change of membrane area, and membrane strength, are measured to characterize various key aspects of the peptide-membrane interaction. The experimental results show that the partitioning of AcE4K in the membrane is pH dependent. The bound peptide inserts in the membrane, which increases the overall membrane area in a pH-dependent manner, however the depth of insertion of the peptide in the membrane is independent of pH. This result suggests that the binding of the peptide to the membrane is driven by the protonation of its three glutamatic acids and the aspartic acid, which results in an increase of the number of bound molecules as the pH decreases from pH 7 to 4.5. The transition between the bound state and the free state is characterized by the Gibbs energy for peptide binding. This Gibbs energy for pH 5 is equal to -30.2 kJ/mol (-7.2 kcal/mol). Most of the change of the Gibbs energy during the binding of AcE4K is due to the enthalpy of binding -27.3 kJ/mol (-6.5 kcal/mol), while the entropy change is relatively small and is on the order of 6.4 J/mol.K (2.3 cal/mol.K). The energy barrier separating the bound and the free state, is characterized by the Gibbs energy of the transition state for peptide adsorption. This Gibbs energy is equal to 51.3 kJ/mol (12.3 kcal/mol). The insertion of the peptide into the membrane is coupled with work for creation of a vacancy for the peptide in the membrane. This work is calculated from the measured area occupied by a single peptide molecule (220 A(2)) and the membrane elasticity (190 mN/m), and is equal to 15.5 kJ/mol (3.7 kcal/mol). The comparison of the work for creating a vacancy and the Gibbs energy of the transition state shows that the work for creating a vacancy may have significant effect on the rate of peptide insertion and therefore plays an important role in peptide binding. Because the work for creating a vacancy depends on membrane elasticity and the elasticity of the membrane is dependent on membrane composition, this provides a tool for modulating the pH for membrane instability by changing membrane composition. The insertion of the peptide in the membrane does not affect the membrane permeability for water, which shows that the peptide does not perturb substantially the packing of the hydrocarbon region. However, the ability of the membrane to retain solutes in the presence of peptide is compromised, suggesting that the inserted peptide promotes formation of short living pores. The integrity of the membrane is substantially compromised below pH 4.8 (threshold pH), when large pores are formed and the membrane breaks down. The binding of the peptide in the pore region is reversible, and the pore size varies on the experimental conditions, which suggests that the peptide in the pore region does not form oligomers.  相似文献   

15.
T P Pitner  J D Glickson 《Biochemistry》1975,14(14):3083-3087
Kinetics of internal rotation about the C(6)-N(6) bond of N-6,N-6-dimethyladenine (M2-6A) was investigated by -1H nuclear magnetic resonance line-shape analysis of the methyl resonances (220 MHz). Rates of rotation were determined for M2-6A deuterated at N(1) and for neutral M2-6A. Activation parameters for monodeuterated M2-6A at 22 degrees are Ea = 13.8kcal/mol, log A = 12.6, incrementG++=14.9 kcal/mol, incrementH++ = 13.1 kcal/mol, incrementS++ = minus 5.8 eu; for neutral M2-6A: Ea = 15.5 kcal/mol, log A = 14.9, incrementG++ = 12.6 kcal/mol, incrementH++ = 14.9 kcal/mol, incrementS++ =7.8 eu. Vertical stacking of bases interferes with internal rotation of the dimethylamino group.  相似文献   

16.
The cellular autolytic reaction system in Streptococcus faecalis ATCC 9790 was analyzed for relative increases in reaction rates with increasing temperature by determination of Arrhenius activation energies (E). The systems examined were: (i) an isolated wall-enzyme complex in 0.01 M sodium phosphate, pH 6.9; (ii) exponential-phase cells suspended in 0.01 or o.3 M sodium phosphate pH 6.8, or in 0.04 M ammonium acetate, pH 6.8, (iii) growing cultures deprived of glucose or lysine; and (iv) cultures treated in growth media with the nonionic detergent, Triton X-100. For detergent-treated cells, E values were between 23.9 and 27.4 kcal/mol (ca. 100.1 to 174.7 kJ/mol) at concentrations of Triton X-100 between about 0.03 and 0.072 mg/ml. E values dropped sharply to 11.5 to 13.0 kcal/m-l (ca. 48.2 to 54.4 kJ/mol) at Triton X-100 concentrations of 0.12 mg/ml or higher. For the remaining systems, E values ranged from 16 to 20 kcal/mol (ca. 67.0 to 83.7 kJ/mol) (wall lysis, cellular autolysis in 0.01 M sodium phosphate or in 0.04 M ammonium acetate, and autolysis of glucose-starved cells) to 31 to 38 kcal/mol (ca 129.8 to 159.1 kJ/mol) (cellular autolysis in 0.3 M sodium phosphate or autolysis of lysine-starved cells). High concentrations of Triton X-100 appear to lower the E values below the 16 to 20 kcal/mol observed for the autolysis of isolated walls. This effect may be related to disruption by the detergent of a hydrophobic complex regulating cellular autolysis in vivo.  相似文献   

17.
This paper addresses the similarities and differences in the topology of the catalytic centres of human liver cytosolic beta-glucosidase and placental lysosomal glucocerebrosidase, and utilizes well-documented reversible active-site-directed inhibitors. This comparative kinetic study was performed mainly to decipher the chemical and structural nature of the active site of the cytosolic beta-glucosidase, whose physiological function is unknown. Specifically, analysis of the effects of a family of alkyl beta-glucosides consistently displayed 100-250-fold lower inhibition constants with the cytosolic broad-specificity beta-glucosidase compared with the placental glucocerebrosidase; for example, with octyl beta-D-glucoside the Ki values were 10 microM and 1490 microM for the cytosolic and lysosomal beta-glucosidases respectively. Furthermore the higher affinity of the cytosolic beta-glucosidase than glucocerebrosidase for the amphipathic alkyl beta-D-glucosides was validated by the greater increase in the free energy of binding with increasing alkyl chain length [delta delta G0 (K,)/CH2: lysosomal enzyme, 2.01 kJ/mol (480 cal/mol); cytosolic enzyme, 3.05 kJ/mol (730 cal/mol)]. The implications of the presence of highly non-polar domains in the active site of the cytosolic beta-glucosidase are discussed with regard to its potential physiological substrates.  相似文献   

18.
The stabilizing effects of dangling ends and terminal base pairs on the core helix GCGC are reported. Enthalpy and entropy changes of helix formation were measured spectrophotometrically for AGCGCU, UGCGCA, GGCGCCp, CGCGCGp, and the corresponding pentamers XGCGCp and GCGCYp containing the GCGC core plus a dangling end. Each 5' dangling end increases helix stability at 37 degrees C roughly 0.2 kcal/mol and each 3' end from 0.8 to 1.7 kcal/mol. The free energy increments for dangling ends on GCGC are similar to the corresponding increments reported for the GGCC core [Freier, S. M., Alkema, D., Sinclair, A., Neilson, T., & Turner, D. H. (1985) Biochemistry 24, 4533-4539], indicating a nearest-neighbor model is adequate for prediction of stabilization due to dangling ends. Nearest-neighbor parameters for prediction of the free energy effects of adding dangling ends and terminal base pairs next to G.C pairs are presented. Comparison of these free energy changes is used to partition the free energy of base pair formation into contributions of "stacking" and "pairing". If pairing contributions are due to hydrogen bonding, the results suggest stacking and hydrogen bonding make roughly comparable favorable contributions to the stability of a terminal base pair. The free energy increment associated with forming a hydrogen bond is estimated to be -1 kcal/mol of hydrogen bond.  相似文献   

19.
Geometry optimization, at the B3LYP/6-311++G** level of theory, was carried out on 4C1 and 1C4 chairs, (3,O)B and B(3,O) boats, and skew-boat conformations of alpha- and beta-D-glucopyranose. Similar calculations on 1,5-anhydro-D-glucitol allowed examination of the effect of removal of the 1-hydroxy group on the energy preference of the hydroxymethyl rotamers. Stable minimum energy boat conformers of glucose were found, as were stable skew boats, all having energies ranging from approximately 4-15 kcal/mol above the global energy 4C1 chair conformation. The 1C4 chair electronic energies were approximately 5-10 kcal/mol higher than the 4C1 chair, with the 1C4 alpha-anomers being lower in energy than the beta-anomers. Zero-point energy, enthalpy, entropy, and relative Gibbs free energies are reported at the harmonic level of theory. The alpha-anomer 4C1 chair conformations were found to be approximately 1 kcal/mol lower in electronic energy than the beta-anomers. The hydroxymethyl gt conformation was of lowest electronic energy for both the alpha- and beta-anomers. The glucose alpha/beta anomer ratio calculated from the relative free energies is 63/37%. From a numerical Hessian calculation, the tg conformations were found to be approximately 0.4-0.7 kcal/mol higher in relative free energy than the gg or gt conformers. Transition-state barriers to rotation about the C-5-C-6 bond were calculated for each glucose anomer with resulting barriers to rotation of approximately 3.7-5.8 kcal/mol. No energy barrier was found for the path between the alpha-gt and alpha-gg B(3,O) boat forms and the equivalent 4C1 chair conformations. The alpha-tg conformation has an energy minimum in the 1S3 twist form. Other boat and skew-boat forms are described. The beta-anomer boats retained their starting conformations, with the exception of the beta-tg-(3,O)B boat that moved to a skew form upon optimization.  相似文献   

20.
N Sone  A Naqui  C Kumar    B Chance 《The Biochemical journal》1984,221(2):529-533
Reaction of O2 and CO with a caa3-type terminal cytochrome oxidase (EC 1.9.3.1) from the thermophilic bacterium PS3 grown with high aeration was studied at low temperatures. The CO recombination at the temperature range studied (-50 degrees C to -80 degrees C) followed first-order kinetics with an activation energy of 29.3 kJ/mol (7.0 kcal/mol). In the presence of O2 at -113 degrees C the photolysed reduced form binds O2 to form an 'oxy' intermediate similar to Compound A. At a higher temperature (-97 degrees C) another intermediate, similar to Compound B, is formed as a result of electron transfer from the enzyme to the liganded O2.  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