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 156 毫秒
1.
A H52Q variant of yeast cytochrome c peroxidase (CcP), in which the distal histidine is replaced by glutamine, catalyzes oxidation of H(2)O(2) instead of reduction. This redirection of catalytic action is detected by protein film voltammetry. In the presence of H(2)O(2), wild-type CcP, adsorbed on a graphite electrode, shows a strong catalytic reduction wave commencing at about 0.8V (pH 5.4); by contrast, H52Q does not exhibit this activity but instead shows a catalytic oxidation current at potentials in the region of 0.9 V. The oxidation current is partly suppressed in the presence of tetranitromethane (a superoxide scavenger) and is not observed for other mutants studied, including H52A. The only significant structural change in the H52Q variant is that the Q-52 side chain occupies the space vacated by the H-52 imidazole; specifically, the N-epsilon atom that is believed to transfer a proton and induce O--O cleavage is replaced, to within 0.75 A, by the carbamide-O. Thus, while the weakly basic amide functionality is unable to serve in the reorganization of bound H(2)O(2), it is able to facilitate its oxidation, most obviously by serving as a H-bond acceptor to assist formation of a labile superoxide intermediate.  相似文献   

2.
When Escherichia coli cells are stressed by hydrogen peroxide (H2O2), synthesis of a large number of proteins is repressed, while several other proteins are induced. Since there is evidence that some lethal effects of near-UV (NUV) radiation may be directly or indirectly due to hydrogen peroxide generated by NUV light, treatment of cells with NUV radiation or H2O2 might be expected to repress and induce the same set of proteins. In this study, we compared the effects of H2O2 and NUV irradiation on patterns of protein induction and/or repression which were separate from the 4-thiouridine-dependent response using growth delay mutants (nuv). Concentrating initially on the proteins that ceased synthesis following NUV irradiation in an nuv mutant, we observed that these were not the same as those that ceased synthesis following H2O2 treatment. Inspection of two-dimensional polyacrylamide gel electrophoresis proteins indicated that NUV irradiation repressed synthesis of a different set of proteins, although there was some overlap between the two (45%). It was also observed that the new proteins which appeared after each of the two treatments were different. This suggests that the induction and/or repression of new proteins following NUV irradiation is not triggered solely via oxidative stress, although there is some overlap between the proteins that are induced or repressed following the two treatments.  相似文献   

3.
When cells are first exposed to low levels of oxidative stress, they develop a resistance to a subsequent challenge of the same stress, even at higher levels. Although some protein(s) induced by oxidative stress likely mediated this adaptive response, the nature of these proteins is unknown. In this study, the total proteins extracted from human U937 leukemia cells exposed to 50 micromM H(2)O(2) for 24 h to induce an optimal protective response were analyzed by two-dimensional polyacrylamide gel electrophoresis. H(2)O(2) treatment induced elevation of level of 34 protein spots. An analysis of these spots by a matrix associated laser desorption/ionization time-of-flight mass spectrometry identified 28 of the H(2)O(2)-induced proteins. These include proteins involved in energy metabolism, translation and RNA processing, chaperoning or mediating protein folding, cellular signaling, and redox regulation, as well as a mitochondrial channel component, and an actin-bundling protein. Therefore, it appears that the cellular adaptation to oxidative stress is a complex process, and is accompanied by a modulation of diverse cellular functions.  相似文献   

4.
Nanocrystalline In(2)O(3)-based solid solutions, with different concentration of Co, with cubic structure were successfully prepared by a simple route. The as-prepared materials were characterized by X-ray diffraction (XRD) and transmission electron microscopy (TEM). The indirect heating structure sensors based on pure and doped In(2)O(3) as sensitive materials were fabricated on an alumna tube with Au electrodes and gas-sensing properties of the sensor elements were measured as a function of concentration of dopant, operating temperature and concentrations of the test gases. The results showed that In(2)O(3) had high response towards H(2)S gas at an operating temperature 150 degrees C, while 2.5 wt.% Co-doped In(2)O(3) sensor exhibited enhance response and selectivity to H(2)S at rather lower operating temperature. Incorporation of platinum further improved the response, selectivity and response time towards H(2)S. Platinum incorporated 2.5 wt.% Co-doped In(2)O(3) (Pt: 0.5 wt.%) was able to detect 10-100 ppm of H(2)S in air at an operating temperature of 100 degrees C. The selectivity of the sensor elements for H(2)S against liquefied petroleum gas (LPG), NH(3) and H(2) gases was studied. The improved gas-sensing properties can mainly be attributed to the selectivity to oxidation of H(2)S and noble metal additive sensitization.  相似文献   

5.
The decomposition of hydrogen peroxide into singlet molecular oxygen-(1)O(2) ((1)Delta(g))-in the presence of lanthanum(iii) salts was studied by monitoring its characteristic IR luminescence at 1270 nm. The process was found to be heterogeneously catalyzed by La(III), provided that the heterogeneous catalyst is generated in situ. The yield of (1)O(2) generation was assessed as 45+/-5 % both in water and in methanol. The pH-dependence on the rate of (1)O(2) generation corresponds to a bell-shaped curve from pH 4.5 to 13 with a maximum around pH 8. The study of the influence of H(2)O(2) showed that the formation of (1)O(2) begins as soon as one equivalent of H(2)O(2) is introduced. It then increases drastically up to two equivalents and more smoothly above. Unlike all other metal salt catalyst systems known to date for H(2)O(2) disproportionation, this chemical source of (1)O(2) is able to generate (1)O(2) not only in basic media, but also under neutral and slightly acidic conditions. In addition, this La-based catalyst system has a very low tendency to induce unwanted oxygenating side reactions, such as epoxidation of alkenes. These two characteristics of the heterogeneous lanthanum catalyst system allow non-photochemical (i.e., "dark") singlet oxygenation of substrate classes that cannot be peroxidized successfully with conventional molybdate catalysts, such as allylic alcohols and alkenyl amines.  相似文献   

6.
Na2[UO2(IO3)4(H2O)] has been synthesized under mild hydrothermal conditions. Its structure consists of Na+ cations and [UO2(IO3)4(H2O)](2-) anions. The [UO2(IO3)4(H2O)](2-) anions are formed from the coordination of a nearly linear uranyl, UO2(2+), cation by four monodentate IO(3-) anions and a coordinating water molecule to yield a pentagonal bipyramidal environment around the uranium center. The water molecules form intermolecular hydrogen bonds with the terminal oxo atoms of neighboring [UO2(IO3)4(H2O)](2-) anions to yield one-dimensional chains that extend down the b axis. There are two crystallographically unique iodate anions in the structure of Na2[UO2(IO3)4(H2O)]. One of these anions is aligned so that the lone-pair of electrons is also directed along the b axis. The overall structure is therefore polar, owing to the cooperative alignment of both the hydrogen bonds and the lone-pair of electrons on iodate. The polarity of the monoclinic space group C2 (a = 11.3810(12) A, b = 8.0547(8) A, c = 7.6515(8) A, beta = 90.102(2) degrees , Z = 2, T = 193 K) found for this compound is consistent with the structure. Second-harmonic generation of 532 nm light from a 1064 nm laser source yields a response of approximately 16x alpha-SiO2.  相似文献   

7.
Abstract— Although the lex gene has been described until recently as being required for lysogenic induction, both our work and the work of others have reported Λ prophage induction in some lexA mutants. However, the characteristics of the process were not defined. We describe UV induction of prophage in a lexA1 mutant at a slightly lower level and requiring 2 times longer than the wild type. As demonstrated in some work, in cells treated with low levels of rifampicin (RIF) no new synthesis of RecA protein is needed for the prophage induction although the onset of lysis is delayed. We suggest that the lysogenic induction in lexA cells is due to the same mechanism that induces prophage in the wild type cells treated with RIF. That is, the induction is due to the cleavage of Λ represser by the basal RecA protease in the DNA-single-strand gap, since RecA protease and monomer represser both have high affinity for this type of DNA. So, LexA protein need not be cleaved for the prophage induction.
No Weigle-reactivation (WR) was detected in the lex mutant even after a long post-irradiation incubation, suggesting that unlike prophage induction, WR requires LexA protein cleavage.  相似文献   

8.
The unimolecular reactivity of M(2+)(H(2)O)(2), M = Be, Mg, Ca, Sr, and Ba, is investigated by density functional theory. Dissociation of the complex occurs either by proton transfer to form singly charged metal hydroxide, MOH(+), and protonated water, H(3)O(+), or by loss of water to form M(2+)(H(2)O) and H(2)O. Charge transfer from water to the metal forming H(2)O(+) and M(+)(H(2)O) is not favorable for any of the metal complexes. The relative energetics of these processes are dominated by the metal dication size. Formation of MOH(+) proceeds first by one water ligand moving to the second solvation shell followed by proton transfer to this second-shell water molecule and subsequent Coulomb explosion. These hydroxide formation reactions are exothermic with activation energies that are comparable to the water binding energy for the larger metals. This results in a competition between proton transfer and loss of a water molecule. The arrangement with one water ligand in the second solvation shell is a local minimum on the potential energy surface for all metals except Be. The two transition states separating this intermediate from the reactant and the products are identified. The second transition state determines the height of the activation barrier and corresponds to a M(2+)-OH(-)-H(3)O(+) "salt-bridge" structure. The computed B3LYP energy of this structure can be quantitatively reproduced by a simple ionic model in which Lewis charges are localized on individual atoms. This salt-bridge arrangement lowers the activation energy of the proton-transfer reaction by providing a loophole on the potential energy surface for the escape of H(3)O(+). Similar salt-bridge mechanisms may be involved in a number of proton-transfer reactions in small solvated metal ion complexes, as well as in other ionic reactions.  相似文献   

9.
By a systematic examination on Na(H2O)n, with n = 4-7, 9, 10, and 15, we demonstrate that a hydrogen loss reaction can be initiated by a single sodium atom with water molecules. This reaction is similar to the well-known size-dependent intracluster hydrogen loss in Mg+(H2O)n, which is isoelectronic to Na(H2O)n. However, with one less charge on Na(H2O)n than that on Mg+(H2O)n, the hydrogen loss for Na(H2O)n is characterized by a higher barrier and a more flexible solvation shell around the metal ion, although the reaction should be accessible, as the lowest barrier is around 8 kcal/mol. Interestingly, the hydroxide ion OH- produced in the process is stabilized by the solvation of H2O molecules and the formation of an ion pair Na+(H2O)4(H2O)n-l-4[OH-(H2O)l]. The activation barrier is reduced as the unpaired electron in Na(H2O)n moves to higher solvation shells with increasing cluster size, and the reaction is not switched off for larger clusters. This is in sharp contrast to the reaction for Mg+(H2O)n, in which the OH- ion is stabilized by direct coordination with Mg2+ and the reaction is switched off for n > 17, as the unpaired electron moved to higher solvation shells. Such a contrast illustrates the important link between microsolvation environment and chemical reactivity in solvation clusters.  相似文献   

10.
Near-resonant vibrational energy exchange between oxygen and water molecules is an important process in the Earth's atmosphere, combustion chemistry, and the chemical oxygen iodine laser (COIL). The reactions in question are (1) O2(1) + O2(0) --> O2(0) + O2(0); (2) O2(1) + H2O(000) --> O2(0) + H2O(000); (3) O2(1) + H2O(000) <--> O2(0) + H2O(010); (4) H2O(010) + H2O(000) --> H2O(000) + H2O(000); and (5) H2O(010) + O2(0) --> H2O(000) + O2(0). Reanalysis of the data available in the chemical kinetics literature provides reliable values for rate coefficients for reactions 1 and 4 and strong evidence that reactions 2 and 5 are slow in comparison with reaction 3. Analytical solution of the chemical rate equations shows that previous attempts to measure the rate of reaction 3 are unreliable unless the water mole fraction is higher than 1%. Reanalysis of data from the only experiment satisfying this constraint provides a rate coefficient of (5.5 +/- 0.4) x 10(-13) cm3/s at room temperature, between the values favored by the atmospheric and laser modeling communities.  相似文献   

11.
12.
This study consists of two parts: The first part comprised an experimental determination of the kinetic parameters for the exchange of water between UO2(H2O)5(2+) and bulk water, including an ab initio study at the SCF and MP2 levels of the geometry of UO2(H2O)5(2+), UO2(H2O)4(2+), and UO2(H2O)6(2+) and the thermodynamics of their reactions with water. In the second part we made an experimental study of the rate of water exchange in uranyl complexes and investigated how this might depend on inter- and intramolecular hydrogen bond interactions. The experimental studies, made by using 17O NMR, with Tb3+ as a chemical shift reagent, gave the following kinetic parameters at 25 degrees C: kex = (1.30 +/- 0.05) x 10(6) s(-1); deltaH(not equal to) = 26.1 +/- 1.4 kJ/mol; deltaS(not equal to) = -40 +/- 5J J/(K mol). Additional mechanistic indicators were obtained from the known coordination geometry of U(VI) complexes with unidentate ligands and from the theoretical calculations. A survey of the literature shows that there are no known isolated complexes of UO2(2+) with unidentate ligands which have a coordination number larger than 5. This was corroborated by quantum chemical calculations which showed that the energy gains by binding an additional water to UO2(H2O)4(2+) and UO2(H2O)5(2+) are 29.8 and -2.4 kcal/mol, respectively. A comparison of the change in deltaU for the reactions UO2(H2O)5(2+)--> UO2(H2O)4(2+) + H2O and UO2(H2O)5(2+) + H2O --> UO2(H2O)6(2+) indicates that the thermodynamics favors the second (associative) reaction in gas phase at 0 K, while the thermodynamics of water transfer between the first and second coordination spheres, UO2(H2O)5(2+) --> UO2(H2O)4(H2O)2+ and UO2(H2O)5(H2O)2+ --> UO2(H2O)6(2+), favors the first (dissociative) reaction. The energy difference between the associative and dissociative reactions is small, and solvation has to be included in ab initio models in order to allow quantitative comparisons between experimental data and theory. Theoretical calculations of the activation energy were not possible because of the excessive computing time required. On the basis of theoretical and experimental studies, we suggest that the water exchange in UO2(H2O)5(2+) follows a dissociative interchange mechanism. The rates of exchange of water in UO2(oxalate)F(H2O)2- (and UO2(oxalate)F2(H2O)2- studied previously) are much slower than in the aqua ion, kex = 1.6 x 10(4) s(-1), an effect which we assign to hydrogen bonding involving coordinated water and fluoride. The kinetic parameters for the exchange of water in UO2(H2O)52+ and quenching of photo excited *UO2(H2O)5(2+) are very near the same, indicating similar mechanisms.  相似文献   

13.
浸渍法制备了催化剂V2O5-Sb2O3-TiO2,考察了V2O5、Sb2O3负载量、pH值和焙烧温度对催化剂V2O5- Sb2O3-TiO2低温氨选择性催化还原(SCR)NO活性的影响;同时,考察了催化剂V2O5-Sb2O3-TiO2抗H2O和SO2毒化性能。结果表明,V2O5和Sb2O3负载量分别为5%和2%、焙烧温度为400℃、pH值为4时,催化剂SCR活性最好,反应温度220℃时,可达97%。Sb2O3的加入不仅能增强V2O5/TiO2的催化活性,而且能明显提高催化剂的抗H2O和SO2毒化性能。SO2、NO吸附暂态反应和TG-DTG测试表明,Sb2O3的促进机制主要是促进了催化剂在SO2存在条件下对NO的吸附,同时,减弱了硫酸铵盐与催化剂之间的相互作用,硫酸铵盐更容易分解。  相似文献   

14.
This paper describes a method for monitoring the degradation of hydrogen peroxide by cells immobilized on a beaded support. The detection is based on the voltammetric reduction of hydrogen peroxide on a mercury film working electrode, whilst combining the concept of sequential injection (SI) with the lab-on-valve (LOV) manifold allows the measurements to be carried out in real time and automatically, in well-defined conditions. The method is shown to be capable of simultaneously monitoring hydrogen peroxide in the 10-1000 microM range and oxygen in the 160-616 microM range. A correction algorithm has been used to ensure reliable H2O2 results in the presence of varying oxygen levels. The method has been successfully applied to monitoring the degradation of H2O2 by wild-type cells and by catalase-overexpressing mouse embryonic fibroblasts. Since the technique allows the monitoring of the initial response rate, it provides data not accessible by current methods that are end-point-based measurements.  相似文献   

15.
The effect of a single water molecule on the reaction between H(2)O(2) and HO has been investigated by employing MP2 and CCSD(T) theoretical approaches in connection with the aug-cc-PVDZ, aug-cc-PVTZ, and aug-cc-PVQZ basis sets and extrapolation to an ∞ basis set. The reaction without water has two elementary reaction paths that differ from each other in the orientation of the hydrogen atom of the hydroxyl radical moiety. Our computed rate constant, at 298 K, is 1.56 × 10(-12) cm(3) molecule(-1) s(-1), in excellent agreement with the suggested value by the NASA/JPL evaluation. The influence of water vapor has been investigated by considering either that H(2)O(2) first forms a complex with water that reacts with hydroxyl radical or that H(2)O(2) reacts with a previously formed H(2)O·OH complex. With the addition of water, the reaction mechanism becomes much more complex, yielding four different reaction paths. Two pathways do not undergo the oxidation reaction but an exchange reaction where there is an interchange between H(2)O(2)·H(2)O and H(2)O·OH complexes. The other two pathways oxidize H(2)O(2), with a computed total rate constant of 4.09 × 10(-12) cm(3) molecule(-1) s(-1) at 298 K, 2.6 times the value of the rate constant of the unassisted reaction. However, the true effect of water vapor requires taking into account the concentration of the prereactive bimolecular complex, namely, H(2)O(2)·H(2)O. With this consideration, water can actually slow down the oxidation of H(2)O(2) by OH between 1840 and 20.5 times in the 240-425 K temperature range. This is an example that demonstrates how water could be a catalyst in an atmospheric reaction in the laboratory but is slow under atmospheric conditions.  相似文献   

16.
Hydrogen peroxide (H(2)O(2)) acts as a signaling molecule in a wide variety of signaling transduction processes and an oxidative stress marker in aging and disease. However, excessive H(2)O(2) production is implicated with various diseases. Nitric oxide (NO) serves as a secondary messenger inducing vascular smooth muscle relaxation. However, mis-regulation of NO production is associated with various disorders. To disentangle the complicated inter-relationship between H(2)O(2) and NO in the signal transduction and oxidative pathways, fluorescent reporters that are able to display distinct signals to H(2)O(2), NO, and H(2)O(2)/NO are highly valuable. Herein, we present the rational design, synthesis, spectral properties, and living cell imaging studies of FP-H(2)O(2)-NO, the first single-fluorescent molecule, that can respond to H(2)O(2), NO, and H(2)O(2)/NO with three different sets of fluorescence signals. FP-H(2)O(2)-NO senses H(2)O(2), NO, and H(2)O(2)/NO with a fluorescence signal pattern of blue-black-black, black-black-red, and black-red-red, respectively. Significantly, we have further demonstrated that FP-H(2)O(2)-NO, a single fluorescent probe, is capable of simultaneously monitoring endogenously produced NO and H(2)O(2) in living macrophage cells in multicolor imaging. We envision that FP-H(2)O(2)-NO will be a unique molecular tool to investigate the interplaying roles of H(2)O(2) and NO in the complex interaction networks of the signal transduction and oxidative pathways. In addition, this work establishes a robust strategy for monitoring the multiple ROS and RNS species (H(2)O(2), NO, and H(2)O(2)/NO) using a single fluorescent probe, and the modularity of the strategy may allow it to be extended for other types of biomolecules.  相似文献   

17.
Diallyl disulfide (DADS) induced apoptosis through the caspase-3 dependent pathway in leukemia cells was earlier reported from this laboratory. In this study, we investigated the involvement of Ca(2+) in DADS-induced apoptotic cell death of HCT-15, human colon cancer cell line. DADS induced the elevation of cytosolic Ca(2+) by biphasic pattern; rapid Ca(2+) peak at 3 min and following slow and sustained elevation till 3 h after the addition of DADS. Production of H(2)O(2) was also observed with its peak value at 4 h. Apoptotic pathways including the sequence of caspase-3 activation, poly(ADP-ribose) polymerase cleavage, and DNA fragmentation by DADS were completely blocked by various inhibitors such as specific caspase-3 inhibitor, free radical scavenger, and intracellular Ca(2+) chelator. N-acetylcystein and catalase treatment prevented the accumulation of H2O2 and later caspase-3 dependent apoptotic pathway. However, these radical scavengers did not block the elevation of intracellular Ca(2+). Treatment of cells with 1, 2-bis (2-aminophenoxyethane)-N, N, N-tetraacetic acid tetrakis -acetoxymethyl ester (BAPTA-AM), cellular Ca(2+) chelator, resulted in a complete blockage of the caspase-3 dependent apoptotic pathway of HCT-15 cells. It abolished the elevation of intracellular Ca(2+), and furthermore, completely inhibited the production of H(2)O(2). These results indicate that cytosolic Ca(2+) elevation is an earlier signaling event in apoptosis of HCT-15 cells. Collectively, our data demonstrate that DADS can induce apoptosis in HCT-15 cells through the sequential mechanism of Ca(2+) homeostasis disruption, accumulation of H(2)O(2), and resulting caspase-3 activation.  相似文献   

18.
采用量子化学计算方法研究了H2O2 氧化N2 生成N2O 和H2O 的机理.结果发现, H2O2 氧化N2 先通过1 个四元环过渡态形成中间体H2N2O2 分子,H2N2O2 再通过一个五元环过渡态形成N2O和H2O.根据计算得到的每步反应的活化能,得知H2O2 氧化N2 生成中间体H2N2O2 分子是整个反应的控制步骤.  相似文献   

19.
One of the most important parameters to be considered when developing a biosensor based on the use of oxidase enzymes is the selective determination of H(2)O(2) in the presence of easily oxidisable interferents. In this work one approach was taken to overcome this problem-the use of metallised carbon electrodes in conjunction with polymers. Polymers, both conducting and nonconducting, have recently become very interesting materials for the suppression of interferences. They are easy to grow on any electrode surface and the extensive range of polymers available provides a huge scope for the large variety of sensors that exist today. They can be grown in organic or aqueous media. In this work, three polymers (polypyrrole, polyaniline and 1,3-diaminobenzene) were examined for their interferent-preventing potential on several types of electrodes. Previous work carried out at the Laboratory of Sensor Development has shown the co-deposition of ruthenium and rhodium on carbon to provide an electrode surface which is highly catalytic and selective towards H(2)O(2)[1]. The co-deposition of Ru, Rh and Pt, as well as Pt on Ru-Rh electrodes was investigated and all these transducers were coupled with the use of the polymers for enhanced elimination of interferences with highly promising results obtained. The best system was seen to be a Ru-Rh metallised electrode polymerised with poly(1,3-diaminobenzene). At an applied potential of +100 mV the response to H(2)O(2) was approximately 200 times greater than the response of any of the potential interferences.  相似文献   

20.
The ultrafast relaxation dynamics of Cu(H(2)O)(2) is investigated using femtosecond photodetachment-photoionization spectroscopy. In addition, stationary points on the Cu(H(2)O)(2) anion, neutral, and cation potential energy surfaces are characterized by ab initio electronic structure calculations. Electron photodetachment from Cu(-)(H(2)O)(2) initiates the dynamics on the ground-state potential energy surface of neutral Cu(H(2)O)(2). The resulting Cu(H(2)O)(2) complexes experience large-amplitude H(2)O reorientation and dissociation. The time evolution of the Cu(H(2)O)(2) fragmentation products is monitored by time-resolved resonant multiphoton ionization. The parent ion, Cu(+)(H(2)O)(2), is not detected above background levels. The rise to a maximum of the Cu(+) signal from Cu(-)(H(2)O)(2), and the decay of the Cu(+)(H(2)O) signal from Cu(-)(H(2)O)(2) have similar tau approximately 10 ps time dependences to the corresponding signals from Cu(-)(H(2)O), but display clear differences at very short and long times. The experimental observations can be understood in terms of the following picture. Prompt dissociation of H(2)O from nascent Cu(H(2)O)(2) gives rise to a vibrationally excited Cu(H(2)O) complex, which dissociates to Cu+H(2)O due to coupling of H(2)O internal rotation to the dissociation coordinate. This prompt dissociation removes all intra-H(2)O vibrational excitation from the intermediate Cu(H(2)O) fragment, which quenches the long time vibrational predissociation to Cu+H(2)O previously observed in analogous experiments on Cu(-)(H(2)O).  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