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 31 毫秒
1.
Mercuric ion, a well-known nephrotoxin, promotes oxidative tissue damage to kidney cells. One principal toxic action of Hg(II) is the disruption of mitochondrial functions, although the exact significance of this effect with regard to Hg(II) toxicity is poorly understood. In studies of the effects of Hg(II) on superoxide (O) and hydrogen peroxide (H2O2) production by rat kidney mitochondria, Hg(II) (1–6 μM), in the presence of antimycin A, caused a concentration-dependent increase (up to fivefold) in mitochondrial H2O2 production but an apparent decrease in mitochondrial O production. Hg(II) also inhibited O-dependent cytochrome c reduction (IC50 ≈?2–3 μM) when O was produced from xanthine oxidase. In contrast, Hg(I) did not react with O in either system, suggesting little involvement of Hg(I) in the apparent dismutation of O by Hg(II). Hg(II) also inhibited the reactions of KO2 (i.e., O) with hemin or horseradish peroxidase dissolved in dimethyl sulfoxide (DMSO). Finally, a combination of Hg(II) and KO2 in DMSO resulted in a stable UV absorbance spectrum [currently assigned Hg(II)-peroxide] distinct from either Hg(II) or KO2. These results suggest that Hg(II), despite possessing little redox activity, enhances the rate of O dismutation, leading to increased production of H2O2 by renal mitochondria. This property of Hg(II) may contribute to the oxidative tissue-damaging properties of mercury compounds.  相似文献   

2.
Maximal rates of superoxide (O) release, and the cytochemical locales of peroxide staining in resident, elicited, and activated macrophages have been determined. Macrophages elicited into the peritoneum with either casein (1.2% w/v) or proteose-peptone (10.0% w/v) release about twice as much O as macrophages activated by infection of the animals with either Listeria monocytogenes, or Bacille Calmette-Guerin (BCG) followed by immune boosting with Purified Protein Derivative (PPD) (i.e., about 35 vs. 14–18 nmol O/min/107 cells). Macrophages elicited with thioglycollate (3.0% w/v) and resident macrophages produce negligible amounts of O upon stimulation with PMA. These data are compared with those reported by other investigators who used different procedures. A cytochemical procedure for localizing peroxide has been modified for use with murine macrophages. No production of H2O2 by macrophages is detected cytochemically in the absence of stimulation. Upon exposure to PMA, resident macrophages are still largely unresponsive. Approximately 20% of the casein elicited macrophages and BCG-PPD activated macrophages exhibit H2O2 staining, which is largely restricted to the cytoplasmic vesicles and channels induced by PMA in these cells. The only exception to this staining pattern is a small population (about 2%) of activated macrophages which exhibits H2O2 staining in the cytoplasmic vesicles and channels and on the plasmalemma as well.  相似文献   

3.
If the collagen triple helix is so built as to have one set of NH ? O hydrogen bonds of the type N3H3(A) ? O2(B), then it is possible to have a linkage between N1H1(B) and O1(A) through the intermediary of a water molecule with an oxygen O leading to the formation of the hydrogen bonds N1(B) ? O and O (A). In the same configuration, another water molecule with an oxygen O can link two earbonyl oxygens of chains A and B forming the hydrogen bonds O O1(A) and O O0 (B). The two water oxygens also become receptors at the same time for CH ? O hydrogen bonds. Thus, the neighboring chains in the triple helix are held together by secondary valence bond linkages occurring regularly sit intervals of about 3 Å along the length of the protofibril. The additional water molecules occur on the periphery of the proto-fibril and will contribute their full share towards stabilizing the structure in the solid state. In solution, they will be disturbed by the medium unless they are protected by long side groups. It appears that this type of two-bonded structure, in which one NH ? O bond is to a water molecule, can explain several observations on the stability and hydrogen exchange properties of collagen itself and related synthetic polypeptides. The nature of the water bonds and their strength are found to be better in the one-bonded structure proposed from Madras than in the one having the coordinates of Rich and Crick.  相似文献   

4.
In this note it is shown that the block design with incidence matrix Ñ = [NNN], where N = c1hNh + coh (11′–Nh). coh and c1h are any non-negative integers and Nh,h = 1, 2,…,p, are incidence matrices of balanced incomplete block designs with the same number of treatments t, is a balanced block design with the block sizes exceeding the number of treatments. In derivation the matrix M0, introduced by CALIński (1971) is utilized.  相似文献   

5.
The influence of proline cis-trans isomerization on the kinetics of lysozyme unfolding was examined carefully according to the theory of Hagerman and Baldwin [(1976) Biochemistry 15, 1462–1473]. As a result, the kinetics of lysozyme unfolding was found to follow the two-state transition model well. The temperature dependencies of kuf and kf over a wide temperature range showed that ΔC = 0 and ΔC = ?6.7 kJ K?1 mol?1 in solutions of different concentrations of GuHCl. The data observed in solutions containing other denaturants also supported the conclusion that ΔC is nearly equal to zero. The activation enthalpies of unfolding (ΔH) were observed at various concentrations of several kinds of denaturants. They were independent of species and concentrations of denaturants ΔH = 200 kJ mol?1). These facts indicate that the aspect of interaction between protein and different kinds of solvent molecules varies only slightly during the unfolding to the transition state, that is, the transition state is at compact as the native one. Therefore, it is also suggested that ΔH of 200 kJ mol?1 is primarily required for the disruption of long-range interactions among different structural domains through a subtle conformational change. We compared the effects of several kinds of denaturants on the unfolding rate. The addition of PrOH more remarkably increases the unfolding rate than do other hydrophilic denaturants. This is probably because PrOH molecules can penetrate into the hydrophobic core of lysozyme, but hydrophilic reagents cannot because of the compactness of the transition state.  相似文献   

6.
Poly(L -arginine) assumes the α-helix in the presence of the tetrahedral-type anions or some polyanions by forming the “ringed-structure bridge” between guanidinium groups and anions which is stabilized by a pair of hydrogen bonds and electrostatic interaction [Ichimura, S., Mita, K. & Zama, M. (1978) Biopolymers 17 , 2769–2782; Mita, K., Ichimura, S. & Zama, M. (1978) Biopolymers 17 , 2783–2798]. This paper describes the parallel CD studies on the conformational effects on poly (L -homoarginine) of various mono-, di-, polyvalent anions and some polyanions, as well as alcohol and sodium dodecylsulfate. The random coil to α-helix transition of poly(L -homoarginine) occurred only in NaClO4 solution or in the presence of high content of ethanol or methanol. The divalent and polyvalent anions of the tetrahedral type (SO, HPO, and P2O), which are strong α-helix-forming agents for poly(L -arginine), failed to induce the α-helical conformation of poly(L -homoarginine). By complexing with poly(L -glutamic acid) or with polyacrylate, which is also a strong α-helix-forming agent for poly(L -arginine), poly(L -homoarginine) only partially formed the α-helical conformation. Monovalent anions (OH?, Cl?, F?, and H2PO) did not change poly(L -homoarginine) to the α-helix, and in the range of pH 2–11, the polypeptide remained in an unordered conformation. In sodium dodecylsulfate, poly(L -homoarginine) exhibited the remarkably enlarged CD spectrum of an extended conformation, while poly(L -arginine) forms the α-helix by interacting with the agent. Thus poly(L -homoarginine), compared with poly(L -arginine), has a much lower ability to form the α-helical conformation by interacting with anions. The stronger hydrophobicity of homoarginine residue in comparison with the arginine residue would provide unfavorable conditions to maintain the α-helical conformation.  相似文献   

7.
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   

8.
The oxygen free radical scavenging activities of 15 chromonyl‐thiazolidine‐2,4‐dione compounds (CTDs) were examined in chemical systems producing superoxide anion radicals, O (potasium superoxide–18‐crown‐6 ether–DMSO), and hydroxyl radicals, HO? (a Fenton reaction: Fe(II)–H2O2–sodium trifluoroacetate, pH 6.15). Chemiluminescence and electron spin resonance (ESR) spectroscopy using 5,5‐dimethyl‐1‐pyrroline‐1‐oxide (DMPO) as spin trap were applied to evaluate antioxidant behaviour of CTDs towards the oxygen radicals. The results indicated that 11 of the 15 tested compounds showed a significant inhibitory effect on the chemiluminescence generated from the O‐generating system, ranging from 41 to 86%, and 13 CTDs quenched the ESR signal of the DMPO–OH spin adduct by 33–86%, at a concentration of 1 mmol L?1. Our findings demonstrate that CTDs could be good free radical scavengers. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

9.
Proton magnetic resonance spectra of model dipeptide molecules R1–C′1O1–N2H2–CHR2–C′2O2–N3H3–R3 in CCl4 solutions exhibit splited signals when investigating on mixtures of L and D enantiomers differing from the racemic composition. The major effect is observed on amide proton signals which are the ones most sensitive to the ratio of aggregation. The stereoselective dimerization of enantiomeric molecules in the so-called C5 conformational state is shown to be responsible for such a phenomenon, the intensity of which depends on the bulkiness of the side chain R2. A theoretical approach is proposed which gives predictions in close agreement with our own experimental findings.  相似文献   

10.
For a balanced one-way classification, where the normally distributed observations obey a random model yij=μ+bi+cij with two variance components var (bi) = δ and var (cij) = δ, the probability is given that the analysis of variance estimate of δ will be negative. This probability depends on δ/δ and the degrees of freedom in the ANOVA table. Tables for this probability are given. If the normally distributed observations obey an intra-class correlation model, the probability that the Mean Square between groups is smaller than the Mean Square within groups can also be evaluated from the given tables.  相似文献   

11.
Maximal rates of O and H2O2 production by human bloodstream monocytes activated during the respiratory burst by phorbol ester were only about 10% of those of neutrophils. Furthermore, monocytes possess only about 5% of the myeloperoxidase activity of neutrophils and so can only produce low levels of HOCI and related compounds. These combined reductions in O generating ability and lower myeloperoxidase levels result in low luminol chemiluminescence stimulated during the respiratory burst of monocytes. However, although monocytes generate much lower levels of O and H2O2 than neutrophils, these cells produce comparable rates of PMA-stimulated lucigenin chemiluminescence. Hence, this assay does not accurately reflect the production of either of these two oxidants by activated phagocytes, and further lucigenin must react with some other oxidant(s) via a process which leads to photon emission. This oxidant(s) is not O, H2O2, · OH, 1O2 or NO, but is derived from O generated during the respiratory burst and is generated in greater quantities by activated monocytes compared with neutrophils. Thus, lucigenin chemiluminescence is an indirect measure of superoxide release.  相似文献   

12.
The starch–iodine blue complex formation does not involve negatively charged iodine species like I, I, or I; rather, neutral iodine units are involved. The heat of reaction is determined to be about ?110 kJ for every mole of I-I unit in the amylose helix, which suggests that the dissociation of I2 (binding energy 149 kJ/mol) does not take place during the complex formation. Quantum mechanical (INDO CI) calculations indicate that the linear as well as nonlinear polyiodine units, I6, with interiodine distance of 3.0 Å are responsible for characteristic absorbance bands of the starch–iodine complex. Based on our previous article [(1989) J. Polym. Sci. A 27 , 4161] and the present studies we identify (C6H10O5)16.5I6 to be the polymeric unit responsible for the characteristic blue color of the complex.  相似文献   

13.
The relationship between cytosolic concentrations of Ca2+ (Ca) and Na+ (Na) were studied in preparations of rat submandibular and pancreatic acini loaded with the Ca2+-sensitive dye Fura-2 or the Na+-sensitive dye SBFI. Pancreatic acini showed no changes in Na during either transient or persistent changes in Ca. Increases in Ca produced by exposure of submandibular gland acini to carbachol, a muscarinic cholinergic agonist, were followed by an increase in Na after a delay of 5–10 s. When Ca2+ stores were mobilized without Ca2+ influx Na also increased, but in acini loaded with BAPTA, a nonfluorescent Ca2+ chelator, the transient increase in Ca2+ caused by mobilization of stored Ca2+ was virtually abolished, as was the increase in Na. In the presence of ionomycin, increases in Ca were followed by increases in Na. Ca2+-dependent increases in Na were abolished in Na+-free buffer and by the presence of furosemide, a blocker of Na+-K+-2Cl cotransport. In other studies, extracellular ATP (ATPo) produced an increase in Ca and Na. The steady-state increase in Ca was reduced by increasing extracellular Na+ concentrations (Na) in dose-dependent fashion (IC50 = 16.4 ± 4.7 mM Na+). Likewise, increasing Na reduced ATPo-stimulated 45Ca2+ uptake at steady state (IC50 = 15.8 ± 9.2 mM Na+). Changing Na had no effect on carbachol-stimulated increases in Ca. We conclude that, in rat submandibular gland acini, ATPo promotes an increase in Ca and Na via a common influx pathway and that, under physiologic conditions, Na+ significantly limits the ATPo-stimulated increase in Ca. In the presence of carbachol, however, Na rises in Ca-dependent fashion in submandibular gland acini via stimulation of Na+-K+-2Cl cotransport. © 1996 Wiley-Liss, Inc.  相似文献   

14.
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   

15.
Spectrophotometric techniques have been employed to study the binding of bromophenol red (BPR) to hen egg white lysozyme and the consequent inhibition of enzyme activity. Experimental evidence is given from the dye binding studies in the presence of hexasaccharide and from the studies on activity that BPR binds at a site outside the proposed cleft region (A–F) in such a way that it inhibits the lytic activity towards cell walls but does not inhibit the activity towards hexasaccharide. These observations are consistent with the kinetics of binding [studied using temperature-jump (T-jump)] in the presence of Co++ or chitotriose in large concentrations and the experiments with acetylated lysozyme which suggest that the binding site of BPR is closer to a lysine residue near the cleft. It is suggested that the binding site of BPR could be important in positioning the peptide segment of the cell walls, which are cleaved in the cleft. Evidence for the statement that this binding takes place at least by a two-step process, in which the bimolecular step is followed by a slower monomolecular step, is given from the observations of two types of 1:1 complexes at 24°C in equilibrium studies and from the concentration dependence of the relaxation observed at 605 nm in the T-jump experiments. The binding process is examined by analyzing the T-jump data obtained between 18 and 33°C in the pH range 5.2–9.2 and ionic strength 0.01–01. The ionic strength and pH dependences of the equilibrium constant associated with the bimolecular step k2/k1 and the forward rate constant associated with monomolecular step k3 have been given as evidence for the suggestion that a Coulombic interaction is involved in the first step of binding. However, the final state of binding is hydrophobic in nature. The enthalpy of activation ΔH and the entropy of activation ΔS associated with kf[= k3(k1/k2)] showed compensation behavior with pH variation, with maxima around pH ~ 7.5 in H2O. This has been interpreted as a maximal disordering of water structure in a region of the enzyme at this pH during the monomolecular step. However, the binding of chitotriose or Co++ in the cleft reduces the ΔH and ΔS associated with the monomolecular step of BPR binding, probably by disordering the structured water during their binding in the cleft. The differences in the kinetic parameters obtained in H2O and in D2O probably arise due to subtle differences in the conformation of the enzyme in the two solvents and apart from isotope effects. The correlation between the pH (or pD) dependence of the “intrinsic activity” towards cell walls and ΔH or ΔS indicates that ordered water structure could be playing a role in controlling the catalytic activity. It is also suggested that this factor is associated with the rate constant k3s of the monomolecular step leading to the formation of the final bound state of the substrate in cell lysis, which is also a factor controlling kcat.  相似文献   

16.
Empirical force-field calculations and ir and 1H-nmr spectra indicate that five-membered (C5) and seven-membered (C) hydrogen-bonded rings are the preferred conformations of acetyl-L -Phe p-acetyl and p-valeryl anilides in nonpolar media. The C5/C ratio was found to be dependent on the dryness of the solute and the solvent. This fact and the results from conformational-energy calculations suggest that a molecule of water participates in the stabilization of the C conformation.  相似文献   

17.
Equilibrium unfolding (folding) studies reveal that the autoregulatory RNA pseudoknots derived from the bacteriophage T2 and T4 gene 32 mRNAs exhibit significant stabilization by increasing concentrations of divalent metal ions in solution. In this report, the apparent affinities of exchange inert trivalent Co(NH3) have been determined, relative to divalent Mg2+, for the folded, partially folded (Kf), and fully unfolded (Ku) conformations of these molecules. A general nonspecific, delocalized ion binding model was developed and applied to the analysis of the metal ion concentration dependence of individual two‐state unfolding transitions. Trivalent Co(NH3) was found to associate with the fully folded and partially unfolded pseudoknotted forms of these RNAs with a Kf of 5–8 × 104 M−1 in a background of 0.10 M K+, or 3‐ to 5‐fold larger than the Kf obtained for two model RNA hairpins and hairpin unfolding intermediates, and ≈ 40–50‐fold larger than Kf for Mg2+. The magnitude of Kf was found to be strongly dependent on the monovalent salt concentration in a manner qualitatively consistent with polyelectrolyte theory, with Kf reaching 1.2 × 105 M−1 in 50 mM K+. Two RNA hairpins were found to have affinities for Co(NH3) and Ru(NH3) of 1–2 ×104 M−1, or ≈ 15‐fold larger than the Kf of ∼ 1000 M−1 observed for Mg2+. Additionally, the Ku of 4,800 M−1 for the trivalent ligands is ≈ 8‐fold larger than the Ku of 600 M−1 observed for Mg2+. These findings suggest that the T2 and T4 gene 32 mRNA pseudoknots possess a site(s) for Mg2+ and Co(NH3) binding of significantly higher affinity than a “duplexlike” delocalized ion binding site that is strongly linked to the thermodynamic stability of these molecules. Imino proton perturbation nmr spectroscopy suggests that this site(s) lies near the base of the pseudoknot stem S2, near a patch of high negative electrostatic potential associated with the region where the single loop L1 adenosine crosses the major groove of stem S2. © 1999 John Wiley & Sons, Inc. Biopoly 50: 443–458, 1999  相似文献   

18.
In this paper it is shown that if N= \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop \sum \limits_{i = 1}^{S_h} $\end{document} cihNih, where cih are some non-negative integer numbers and Nih are such incidence matrices that Ah = \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop \sum \limits_{i = 1}^{S_h} $\end{document} i Nih is a balanced matrix defined by SHAH (1959), for h = 1, 2,…, p, then a block design with an incidence matrix Ñ = [N, N,…,N] is an equi-replicated balanced block design. Here the balance of a block design is defined in terms of the matrix M0 introduced by CALI?SKI (1971).  相似文献   

19.
Modes of aggregation fo alanine-, norvaline- and valine-contaiing dpepetides with the general formula R1? C1O1? N2H2? CHR2? C2O2? N3H3? R3 have been studied in CCl4 solution by using infrared and nuclear magnetic resonance spectroscopies. Solutions of the pure L isomer and of the racemic mixture do not give identical data. At a given concentration, the racemic mixtrue is more aggregated than the pure enantiomer, and the difference, negligible in the case of alanine derivative, increases wiht the bulkiness of the side cahin R2. The results show that a selective interaction takes place between enantiomeric molecules, resulting ina dimer associating tow inverse configurated C5 conformers. The stabilizaion of this dimer proceeds from two symmetrical and intermolecular H3 … O1 hydrogen bondings.  相似文献   

20.
John A. Schellman 《Biopolymers》1994,34(8):1015-1026
A model for solvation in mixed solvents, which was developed for the free energy and preferential interaction [J. A. Schellman (1987), Biopolymers, Vol. 26, pp. 549–559; (1990), Biophysical Chemistry, Vol. 37, pp. 121–140; (1993), Biophysical Chemistry, Vol. 45, pp. 273–279], is extended in this paper to cover the thermal properties: enthalpy, entropy, and heat capacity. An important result is that the enthalpy of solvation H? responds directly to the fraction of site occupation. This differs from the free energy ? and preferential interaction Γ32, which are measures of the excess binding above a random distribution of solvent molecules. In other words, the enthalpy is governed by K while ? and Γ32 are governed by (K ? 1) where K is the equilibrium constant on a mole fraction scale [Schellman (1987)]. The solvation heat capacity C?p consists of two term: (1) the intrinsic heat capacity of species in solution with no change in composition, and (2) a term that accounts for the change in composition that accompanies solvent exchange. Binding to biological macromolecules is heterogeneous but experiementalists must use binding isotherms that assume the homogeneity of sites. Equations are developed for the interpretation of the experimental parameters (number of sites nexp, equilibrium constant Kexp, and enthalpy, Δhexp), when homogeneous formulas are applied to the heterogeneous case. It is shown that the experimental parameters for the occupation and enthalpy are simple functions of the moments of the distribution of equilibrium constants over the sites. In general, nexp is greater than the true number of sites and Kexp is greater than the average of the equilibrium constants. The free energy and preferential interaction can be fit to a homogenious formula, but the parameters of the curve are not easily represented in terms of the moments of distributions over the sites. The strengths and deficiencies of this type of thermodynamic model are discussed. © 1994 John Wiley & Sons, Inc.  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