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 265 毫秒
1.
Direct enantiomer separation of hypericin, pseudohypericin, and protohypericin was accomplished by high‐performance liquid chromatography (HPLC) using immobilized polysaccharide‐type chiral stationary phases (CSPs). Enantioselectivities up to 1.30 were obtained in the polar‐organic elution mode whereby for hypericin and pseudohypericin Chiralpak IC [chiral selector being cellulose tris(3,5‐dichlorophenylcarbamate)] and for protohypericin Chiralpak IA (chiral selector being the 3,5‐dimethylphenylcarbamate of amylose) gave favorable results. Enantiomers were distinguished by on‐line electronic circular dichroism detection. Optimized enantioselective chromatographic conditions were the basis for determining stereodynamic parameters of the enantiomer interconversion process of hypericin and pseudohypericin. Rate constants delivered by computational simulation of dynamic HPLC elution profiles (stochastic model, consideration of peak tailing) were used to calculate averaged enantiomerization barriers (ΔG) of 97.6–99.6 kJ/mol for both compounds (investigated temperature range 25–45°C). Complementary variable temperature off‐column (i.e., in solution) racemization experiments delivered ΔG = 97.1–98.0 kJ/mol (27–45°C) for hypericin and ΔG = 98.9–101.4 kJ/mol (25–55°C) for pseudohypericin. An activation enthalpy of ΔH# = 86.0 kJ/mol and an activation entropy of ΔS# = ?37.7 J/(K mol) were calculated from hypericin racemization kinetics in solution, whereas for pseudohypericin these figures amounted to 74.1 kJ/mol and ?82.6 J/(K mol), respectively. Although the natural phenanthroperylene quinone pigments hypericin and pseudohypericin as well as their biological precursor protohypericin are chiral and can be separated by enantioselective HPLC low enantiomerization barriers seem to prevent the occurrence of an excess of one enantiomer under typical physiological conditions—at least as long as stereoselective intermolecular interactions with other chiral entities are absent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
In rare but nevertheless important cases it is of practical interest to decrease the thermostability of an enzyme, that is, to increase thermolability in a controlled manner. In the present model study, this unconventional goal has been reached by applying directed evolution to the lipase from Pseudomonas aeruginosa (PAL). By utilizing the B‐factor iterative test (B‐FIT), previously developed to increase the thermostability of enzymes, it was possible to reduce the value from 71.6°C in the case of wild type (WT‐PAL) to 35.6°C (best mutant) without affecting the catalytic profile in terms of substrate acceptance or enantioselectivity at room temperature. Accordingly, saturation mutagenesis was performed at sites in PAL, which on the basis of its X‐ray structure, have the lowest B‐factors indicative of high rigidity. Focused mutations were introduced which can be expected to decrease rigidity, the ensuing increased flexibility leading to higher thermolability without changing the actual catalytic profile. Biotechnol. Bioeng. 2009;102: 1712–1717. © 2008 Wiley Periodicals, Inc.  相似文献   

3.
Integral enthalpies of solution of several dipeptides and tripeptides in water at low concentrations have been determined at 25 and 35°C. These data have been used to derive the changes in heat capacity on dissolution at infinite dilution ΔC at 30°C. Limiting partial molal heat capacities ΔC have been determined by combining ΔC with Cp2 (heat capacity of pure solid peptides). Using the data on ω-amino acids and these peptides, the partial molal heat capacity of a peptide group ? CONH? was semiquantitatively estimated.  相似文献   

4.
The low-temperature heat capacity of collagen (in the hydrated and dehydrated states) and the large entropy of collagen in the coiled state relative to the same protein in the helical state were investigated. The heat capacity for collagen in the solid state in the temperature range 4°–50° K changes proportionally to the square of temperature (CpT2). Above 50°K there is a linear dependence (CpT). The differences in the character of temperature dependence of heat capacity for the hydrated and dehydrated collagen show the importance of the specific interaction of water molecules with polypeptide chains of this protein. The peculiarities of the temperature dependence of the heat capacity difference (ΔCp) of hydrated denatured (random coiled) and hydrated native (helical) collagen are observed at 15°, 120°, and 240°K. These differences are caused by the varying degree of ordering of the hydrate water molecules in native and denatured collagen macromolecules. At all temperatures (4°–300°K) the entropy of the random coiled state is higher than that of collagen in the native state and at 298°K ΔS = ∫ (ΔCp/T)dT = 0.8 cal/100 g °K.  相似文献   

5.
6.
The structural and energetic perturbations caused by cavity-creating mutations (Leu-41 → Val and Leu-41 → Ala) in the dimeric 4-α-helical-bundle protein ROP have been characterized by CD spectroscopy and differential scanning calorimetry (DSC). Deconvolution of the CD spectra showed a decrease in α -helicity as a result of the amino acid exchanges that follows qualitatively the overall decrease in conformational stability. Transition enthalpies are sensitive probes of the energetic change associated with point mutations. ΔH0 values at the respective transition temperatures, T1/2 (71.0, 65.3, and 52.9°C at 0.5 mg/ml) decrease from 580 ± 20 to 461 ± 20 kJ/(mol of dimmer) and 335 ± 20 kJ/(mol of dimmer) for wildtype ROP (Steif, C., Weber, P., Hinz, H.-J., Flossdorf, J., Cesareni, G., Kokkinidis, M. Biochemistry 32:3867-3876, 1993), L41V, and L41A, respectively. The conformational stabilities at 25°C expressed by the standard Gibbs energies of denaturation, ΔG, are 71.7, 61.1, and 46.1 kJ/(mol of dimmer). The corresponding transition enthalpies have been obtained from extrapolation using the c(T)and c(T) functions. Their values at 25°C are 176.3, 101.9, and 141.7 kJ/(mol of dimmer) for wild-type ROP, L41V, and L41A, respectively. When the stability perturbation resulting from the cavity creating mutations is referred to the exchange of 1 mol of CH2 group, the average ΔΔG value is ?5.0 ± 1 kJ/(mol of CH2 group). This decrease in conformation stability suggests that dimeric ROP exhibits the same susceptibility to Leu → Yal and Leu → Ala exchanges as small monomeric proteins. Careful determinations of the partial specific heat capacities of wild-type and mutated protein solutions suggest that the mutational effects are predominantly manifested in the native rather than the unfolded state. © 1995 Wiley-Liss, Inc.  相似文献   

7.
The kinetics of ethidium's intercalative binding to DNA packaged in bacteriophage T7 and two T7 deletion mutants have been determined, using enhancement of fluorescence to quantitate binding. At a constant ethidium concentration, the results can be described as first-order binding with two different rate constants, k (= k1 + k?1) and k (= k2 + k?2). The larger rate constant (k) was at least four orders of magnitude smaller than the comparable first-order forward rate constant for binding to DNA released from its capsid. At 25°C values of k decreased as the amount of DNA packaged per internal volume increased. This latter observation indicates that the rate of ethidium's binding to packaged T7 DNA is limited by an event that occurs inside of the DNA-containing region of T7, not by the crossing of T7 capsid's outer shell. Arrhenius plots of kM are biphasic, indicating a transition for packaged DNA at a temperature of 20°C. The data indicate that k s are limited by either sieving of ethidium during its passage through the packaged DNA or subsequent hindered intercalation.  相似文献   

8.
Cystathionine β‐lyase (CBL) catalyzes the hydrolysis of L ‐cystathionine (L ‐Cth) to produce L ‐homocysteine, pyruvate, and ammonia. A series of active‐site mutants of Escherichia coli CBL (eCBL) was constructed to investigate the roles of residues R58, R59, D116, W340, and R372 in catalysis and inhibition by aminoethoxyvinylglycine (AVG). The effects of these mutations on the kcat/K for the β‐elimination reaction range from a reduction of only 3‐fold for D116A and D116N to 6 orders of magnitude for the R372L and R372A mutants. The order of importance of these residues for the hydrolysis of L ‐Cth is: R372 >> R58 > W340 ≈ R59 > D116. Comparison of the kinetic parameters for L ‐Cth hydrolysis with those for inhibition of eCBL by AVG demonstrates that residue R58 tethers the distal carboxylate group of the substrate and confirms that residues W340 and R372 interact with the α‐carboxylate moiety. The increase in the pKa of the acidic limb and decrease in the pKa of the basic limb of the kcat/K versus pH profiles of the R58K and R58A mutants, respectively, support a role for this residue in modulating the pKa of an active‐site residue.  相似文献   

9.
Incorporation of noncanonical disulfide linkages into single‐domain antibodies (sdAbs) has been shown to enhance thermostability and other properties. Here, we evaluated the effects of introducing a novel disulfide linkage formed between Cys residues at IMGT positions 40 and 55 on the melting temperatures (T ms), reversibility of thermal unfolding, solubility, and antigen‐binding affinities of three types of sdAbs (VHH, VH, and VL domains). The Cys40‐Cys55 disulfide linkage was tolerated by 9/9 VHHs, 12/12 VHs, and 2/11 VLs tested and its formation was confirmed by mass spectrometry. Using circular dichroism, we found that the Cys40‐Cys55 disulfide linkage increased sdAb T m by an average of 10.0°C (range: 0–21.8°C). However, enhanced thermostability came at the cost of a partial loss of refolding ability upon thermal denaturation as well as, for some sdAbs, significantly decreased solubility and antigen‐binding affinity. Thus, Cys40/Cys55 can be added to the panel of known locations for introducing stabilizing noncanonical disulfide linkages into antibody variable domains, although its effects should be tested empirically for individual sdAbs.  相似文献   

10.
Cystathionine γ‐synthase (CGS) catalyzes the condensation of O‐succinyl‐L ‐homoserine (L ‐OSHS) and L ‐cysteine (L ‐Cys), to produce L ‐cystathionine (L ‐Cth) and succinate, in the first step of the bacterial transsulfuration pathway. In the absence of L ‐Cys, the enzyme catalyzes the futile α,γ‐elimination of L ‐OSHS, yielding succinate, α‐ketobutyrate, and ammonia. A series of 16 site‐directed variants of Escherichia coli CGS (eCGS) was constructed to probe the roles of active‐site residues D45, Y46, R48, R49, Y101, R106, N227, E325, S326, and R361. The effects of these substitutions on the catalytic efficiency of the α,γ‐elimination reaction range from a reduction of only ~2‐fold for R49K and the E325A,Q variants to 310‐ and 760‐fold for R361K and R48K, respectively. A similar trend is observed for the kcat/K of the physiological, α,γ‐replacement reaction. The results of this study suggest that the arginine residues at positions 48, 106 and 361 of eCGS, conserved in bacterial CGS sequences, tether the distal and α‐carboxylate moieties, respectively, of the L ‐OSHS substrate. In contrast, with the exception of the 13‐fold increase observed for R106A, the K is not markedly affected by the site‐directed replacement of the residues investigated. The decrease in kcat observed for the S326A variant reflects the role of this residue in tethering the side chain of K198, the catalytic base. Although no structures exist of eCGS bound to active‐site ligands, the roles of individual residues is consistent with the structures inhibitor complexes of related enzymes. Substitution of D45, E325, or Y101 enables a minor transamination activity for the substrate L ‐Ala.  相似文献   

11.
Wei Liu  Takashi Norisuye 《Biopolymers》1988,27(10):1641-1654
Weight-average molecular weights Mw, second virial coefficients, and z-average radii of gyration 〈S2〉 were determined by light scattering as a function of temperature T for four sodium salt samples of xanthan in 0.01M aqueous NaCl, in which the polysaccharide undergoes an order–disorder conformation change with increasing T. The data for 〈S2〉 and Mw at 25 and 80°C, the lowest and highest temperatures studied, confirmed the previous conclusion that the predominant conformation at the former T, i.e., in the ordered state, is a double helix, while that at the latter T, i.e., in the disordered state, is a dimerized coil expanded by electrostatic repulsions between charged groups of the polymer. As T was increased from 25 to 80°C, 〈S2〉 sigmoidally decreased or increased depending on the dimer's molecular weight. This temperature dependence of 〈S2〉 and that determined elsewhere for a high molecular weight sample were found to be described almost quantitatively by a simple dimer model in which the double helix melts from both ends, when the double-helical fraction in the dimer at a given T estimated previously from optical rotation data was used.  相似文献   

12.
Intensity fluctuations of laser light scattered from filamentous viruses Pf1 [length L (Å) × diameter d (Å) = 20,000 × 90], M13 (9000 × 90), potato virus X (5150 × 130), and tobacco mosaic virus (3000 × 180) in sucrose density gradients were measured with a photon correlation spectrometer over a range of scattering angles from 15° to 120°. The experimental data can be approximated by two exponential decays, “slow” and “fast.” The slow decay rate constant t corresponds to the translational diffusion D of the virus, i.e., t = K2D, where K is the magnitude of the scattering vector. The amplitude of the slow component, i.e., translational diffusion, remains greater than that of the fast component, even at high KL. The fast decay rate constant t is also proportional to K2 for viruses such as Pf1, M13, and even potato virus X. In the companion paper, we shall attribute the amplitude enhancement of the translational diffusion to the coupling of its anisotropy to the rotational diffusion modes. In order to explain the excessive decay rates in the fast component, we need to consider the bending mode of rodlike viruses, especially in the longer viruses such as M13 and Pf1, in addition to the usually expected rotational diffusion modes.  相似文献   

13.
Purple non‐sulfur (PNS) bacteria produce hydrogen by photofermentation of organic acids in wastewater. However, NH in wastewater may inhibit hydrogen synthesis by repressing the expression and activity of nitrogenase, the enzyme catalyzing hydrogen production in PNS bacteria. In this study, the Rhodobacter sphaeroides 6016 glnA gene encoding glutamine synthetase (GS) was knocked out by homologous recombination, and the effects on hydrogen production and nitrogenase activity were examined. Using 3 mM glutamine as the nitrogen source, hydrogen production (1,245–1,588 mL hydrogen/L culture) and nitrogenase activity were detected in the mutant in the presence of relatively high NH concentrations (15–40 mM), whereas neither was detected in the wild‐type strain under the same conditions. Further analysis indicated that high NH concentrations greatly inhibited the expression of nifA and nitrogenase gene in the wild‐type strain but not in the glnA1? mutant. These observations suggest that GS is essential to NH repression of nitrogenase and that deletion of glnA1 results in the complete derepression of nitrogenase by preventing NH assimilation in vivo, thus relieving the inhibition of nifA and nitrogenase gene expression. Knocking out glnA1 therefore provides an efficient approach to removing the inhibitory effects of ammonium ions in R. sphaeroides and possibly in other hydrogen‐producing PNS bacteria. Biotechnol. Bioeng. 2010;106: 564–572. © 2010 Wiley Periodicals, Inc.  相似文献   

14.
Excited‐state intramolecular proton transfer (ESIPT) and dual luminescence behaviour of 3‐hydroxyflavone (3‐HF) have been utilized to monitor its binding to liposomal membranes prepared from egg yolk phosphatydilcholine (EYPC). Additionally, absorption spectrophotometric assay has been performed to evaluate the antioxidant activity of 3‐HF against lipid peroxidation in this membrane system. When 3‐HF molecules are partitioned into EYPC liposomes, a weak long‐wavelength absorption band with λ ~410 nm appears in addition to the principal absorption at ~λ = 345 nm. Selective excitation of the 410 nm band produces the characteristic emission (λ~460 nm) of the ground‐state anionic species, whereas excitation at the higher energy absorption band leads to dual emission with predominatly ESIPT tautomer fluorescence (λ = 528 nm). Both ESIPT tautomer and the anionic species exhibit fairly high fluorescence anisotropy (r) values (r = 0.122 and 0.180, respectively). Biexponential fluorescence decay kinetics are observed for the ESIPT tautomer as well as the ground‐state anionic forms, indicating heterogeneity in the microenvironments of the corresponding emitting species. Furthermore, we demonstrate that lipid peroxidation of EYPC liposomes is significantly inhibited upon 3‐HF binding, suggesting that 3‐HF can be potentially useful as an inhibitor of peroxidative damage of cell membranes. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

15.
Glucose-limited chemostat cultures of Candida utilis were cultivated at various pH levels (3.0–7.5), temperatures (15–37.5°C), dilution rates (0.06–0.42 hr?1), and with different nitrogen sources (NH and NO). The ratio of total nucleic acid to protein increased with increase in dilution rate at constant temperature and decreased with increase in temperature at constant dilution rate. The pattern of these variations is consistent with the hypothesis that the nucleic acid to protein ratio is a function of the ratio of the actual dilution rate to the critical dilution rate corresponding to each one of the cultivation temperatures. This ratio is called “reduced dilution rate.” A basis is proposed on which various microorganisms may be compared with respect to the ratios of cell protein to nucleic acid, RNA, ribosomal RNA, and polysomes.  相似文献   

16.
The limited treatment option for recurrent prostate cancer and the eventual resistance to conventional chemotherapy drugs has fueled continued interest in finding new anti‐neoplastic agents of natural product origin. We previously reported anti‐proliferative activity of deoxypodophyllotoxin (DPT) on human prostate cancer cells. Using the PC‐3 cell model of human prostate cancer, the present study reveals that DPT induced apoptosis via a caspase‐3‐dependent pathway that is activated due to dysregulated mitochondrial function. DPT‐treated cells showed accumulation of the reactive oxygen species (ROS), intracellular Ca surge, increased mitochondrial membrane potential (MMP, ΔΨm), Bax protein translocation to mitochondria and cytochrome c release to the cytoplasm. This resulted in caspase‐3 activation, which in turn induced apoptosis. The antioxidant N‐acetylcysteine (NAC) reduced ROS accumulation, MMP and Ca surge, on the other hand the Ca2+ chelator BAPTA inhibited the Ca overload and MMP without affecting the increase of ROS, indicating that the generation of ROS occurred prior to Ca2+ flux. This suggested that both ROS and Ca signaling play roles in the increased MMP via Ca‐dependent and/or ‐independent mechanisms, since ΔΨm elevation was reversed by NAC and BAPTA. This study provides the first evidence for the involvement of both ROS‐ and Ca‐activated signals in the disruption of mitochondrial homeostasis and the precedence of ROS production over the failure of Ca2+ flux homeostasis. J. Cell. Biochem. 114: 1124–1134, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
Molecular conformation of some polysaccharides in aqueous solution in evidenced by changes in the optical rotation and in the elution pattern of gel filtration. The changes in the specific rotation against that in water are expressed as a molar conformational value [K]: ?495° for colominic acid in 1.0 N NaOH solution, and ?180° for hyaluronate (HA), +85° for corneal keratin sulfate, and +234° for amylose in 8 M urea solution. The gel filtration of amylose and HA dissolved in 8 M urea solution shows elution patterns distinctly different from those dissolved in water. The phenomena are attributable to a molecular conformational transition of polysaccharide molecules in aqueous solution.  相似文献   

18.
19.
20.
4‐Hydroxyisoleucine, a promising drug, has mainly been applied in the clinical treatment of type 2 diabetes in the pharmaceutical industry. l ‐Isoleucine hydroxylase specifically converts l‐ Ile to 4‐hydroxyisoleucine. However, due to its poor thermostability, the industrial production of 4‐hydroxyisoleucine has been largely restricted. In the present study, the disulfide bond in l ‐isoleucine hydroxylase protein was rationally designed to improve its thermostability to facilitate industrial application. The half‐life of variant T181C was 4.03 h at 50°C, 10.27‐fold the half‐life of wild type (0.39 h). The specific enzyme activity of mutant T181C was 2.42 ± 0.08 U/mg, which was 3.56‐fold the specific enzyme activity of wild type 0.68 ± 0.06 U/mg. In addition, molecular dynamics simulation was performed to determine the reason for the improvement of thermostability. Based on five repeated batches of whole‐cell biotransformation, Bacillus subtilis 168/pMA5‐idoT181C recombinant strain produced a cumulative yield of 856.91 mM (126.11 g/L) 4‐hydroxyisoleucine, which is the highest level of productivity reported based on a microbial process. The results could facilitate industrial scale production of 4‐hydroxyisoleucine. Rational design of disulfide bond improved l ‐isoleucine hydroxylase thermostability and may be suitable for protein engineering of other hydroxylases.  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