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 125 毫秒
1.
The quinones 1,4-naphthoquinone (NQ), methyl-1,4-naphthoquinone (MNQ), trimethyl-1,4-benzoquinone (TMQ) and 2,3-dimethoxy-5-methyl-1,4-benzoquinone (UQ-0) enhance the rate of nitric oxide (NO) reduction by ascorbate in nitrogen-saturated phosphate buffer (pH 7.4). The observed rate constants for this reaction were determined to be 16±2,215±6,290±14 and 462±18?M-1?s-1, for MNQ, TMQ, NQ and UQ-0, respectively. These rate constants increase with an increase in quinone one-electron redox potential at neutral pH, E71. Since NO production is enhanced under hypoxia and under certain pathological conditions, the observations obtained in this work are very relevant to such conditions.  相似文献   

2.
In the present study, the first quantum chemical calculations of structures and vibrational spectra of radicals of 1,4-naphthoquinone and 2-methoxy-1,4-benzoquinone that account for electron correlation are presented. In the case of 1,4-naphthoquinone a good agreement between calculated vibrational frequencies and 18O-shifts of the 1,4-naphthoquinone radical (protonated radical anion) with experimental data of a species detected after irradiation of vitamin K1 in solution is found. Our calculations, thus, support the previous assignment. In the case of 2-methoxy-1,4-benzoquinone we have localized the stable conformations with respect to the orientation of the methoxy group and we have determine the harmonic force fields for these structures. Our calculation suggest that, while the frequencies of the two conformers are similar, the 18O-shift of the most intensive absorptions in the spectral region between 1400 and 1700 cm–1 of the two conformers differ significantly and might serve as a tool to distinguish between the two conformers. The applied DFT method is shown to predict electron affinities which are systematically underestimated by 10%.  相似文献   

3.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


4.
The Reaction of no With Superoxide   总被引:35,自引:0,他引:35  
The rate constant for the reaction of NO with ·O2- was determined to be (6.7 ± 0.9) × 109 1 mol-1 s-1, considerably higher than previously reported. Rate measurements were made from pH 5.6 to 12.5 both by monitoring the loss of ·O2- and the formation of the product -OONO. The decay rate of -OONO, in the presence of 0.1 moll-1 formate, ranges from 1.2s-1 at pH 5 to about 0.2s-1 in strong base, the latter value probably reflecting catalysis by formate.  相似文献   

5.
Under physiological pH conditions (pH 7.2-7.4) the rate constant of the reaction NO + O2 yielding peroxonitrite (ONOO) was determined as k = (3.7 ± 1.1) × 107 M 1 s 1. The decay of peroxonitrite at this pH follows first order kinetics with a rate constant of 1.4 s 1. At alkaline pH peroxonitrite is practically stable.

Possible consequences of these reactions for the biological lifetime of EDRF will be discussed.  相似文献   

6.
Thylakoid membranes isolated from halophytic species showed differences in their interactions with ionic and lipophilic electron acceptors when compared to thylakoids from non-halophytes. FeCN was considerably less efficient as electron acceptor with halophyte thylakoids, supporting much lower rates of O2 evolution and having a lower affinity. FeCN accepted electrons at a different, DMMIB insensitive, site with these thylakoids. 1,4-Benzo-quinones with less positive midpoint potentials were less effective in accepting electrons from halophyte thylakoids compared to nonhalophyte thylakoids, also reflected in lower rates of O2 evolution and lower affinity. Considering the lipolphilic nature and the fact that there was no apparent change in the site donating electrons to the quinones, an alteration in the midpoint potential of this site by about +100mV is postulated for the halophyte thylakoids.Abbreviations AMPD 2-amino-2-methyl-1,3-propanediol - Cyt b6/f cytochrome b6/f complex - DBMIB 2,5-dibromo-6-isopropyl-3-methyl-1,4-benzoquinone - DCBQ 2,6-dichloro-1,4-benzoquinone - DCIP 2,6-dichlorophenol-indolphenol - DMBQ 2,5-dimethyl-1,4-benzoquinone - Em7 midpoint redox potential at pH 7.0, FeCN-K3Fe(CN)6 - HNQ 5-hydroxy-1,4-naphthoquinone - MV methylviologen - NQ 1,4-naphthoquinone - PBQ phenyl-1,4-benzoquinone - PC plastocyanin - PQ plastoquinone  相似文献   

7.
Direct evidence obtained by means of the technique of pulse radiolysis-kinetic spectrometry, with measurements in the time range 10−6 to 1 s, is presented that, consequent upon reaction of a single H-atom with a single molecule of ferricytochrome c, a reducing equivalent is transmitted via the protein structure to the ferriheme moiety. Such transmission accounts for at least 70% of the total reduction of the ferri to the ferro state of cytochrome c. The remainder of the total reduction takes place without stages resolvable on the time scale of these experiments. Reduction brought about by H atoms appears to follow a different course than reduction by hydrated electrons. In the latter case, intramolecular transmission of reducing equivalents could not be demonstrated (Lichtin, N. N., Shafferman, A. and Stein, G. (1973) Biochim. Biophys. Acta 314, 117–135).

Not every H-atom reacts with ferricytochrome c at a site which results in conversion of the Fe(III) state to the Fe(II) state. Approximately half of reacting H-atoms do not produce reduction.

The following second order rate constants have been determined in solutions of low ionic strength at 20±2 °C: k[H+ferricytochrome c] = (1.0±0.2) · 1010 M−1 · s−1 at pH 3.0 and 6.7; k[H+ferrocytochrome c] = (1.3±0.2) · 1010 M−1 · s−1 at pH 3.0; k[eaq + ferrocytochrome c] = (1.9±0.4) · 1010 M−1 · s−1 at pH 6.7.  相似文献   


8.
利用7-diethylamino-3-(4‘-maleimidylphenyl)-4-methylcoumarin(CPM)甲基化钙通道CPM能明显阻碍由1,4NQ诱导的对RyR1的激活,但对由1,4NQ产生的抑制没有明显作用。用经CPM预处理过和没有处理过的样品与1,4NQ温育手离心清洗,发现未经CPM处理的样品经激活和抑制浓度的1,4NQ温育后,离心清洗不能消除1,4NQ对RyR1的激活和抑制。而经CPM处理后的样品,在1,4NQ的激活浓度和抑制浓度下,离心清洗后 有观测到1,4NQ诱导的激活态,但抑制态仍然明显存在,以上结果说明,CPM所识别的特异活性硫醇参与了1,4NQ激活作用:CPM与此硫醇发生甲基化作用,因此阻断了1,4NQ对RyR1上的激活部位硫醇发生作用,使钙通道不能被激活。  相似文献   

9.
The photoreduction of 1,4-benzoquinone (BQ), 1,4-naphthoquinone (NQ), 9,10-anthraquinone (AQ) and several derivatives, e.g. dimethylBQ, trimethylBQ, duroquinone, bromoNQ, methoxyNQ, methylAQ and dimethylAQ in acetonitrile-water by ascorbate was studied by time-resolved UV-vis spectroscopy using 20 ns laser pulses at 308 nm and continuous 254 nm irradiation. The semiquinone radical (*QH/Q*(-)) is formed after H-atom transfer from ascorbate to the quinone triplet state. The rate constant for quenching is k(q)=(2-9) x 10(9) M(-1) s(-1). Termination of the radicals takes place in the micros-ms range. The results are compared with those initiated by electron transfer from DABCO under similar conditions, where the k(q) values are similar, but the termination of Q*(-) takes place by electron back transfer not yielding hydroquinones. Specific properties of the quinone triplet state, e.g. self-quenching, nucleophilic water addition and the effects of structure are discussed.  相似文献   

10.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


11.
Hypochlorous acid (HOCl) is an oxygen-derived species involved in physiological processes related to the defence of the organism that may cause adverse effects when its production is insufficiently controlled. In order to examine its reactivity with potential scavenging molecules from the non steroidal anti-inflammatory drugs (NSAIDs) family, a competition assay based on para-aminobenzoic acid (PABA) chlorination was developed. The original optimised in vitro fluorimetric procedure offered the possibility to determine rate constants (ks) for the reaction with HOCl in physiologically relevant conditions. The specificity of the system was improved by a liquid chromatography (LC) which allows the separation of the drugs and their oxidation products. After determination of the rate constant for PABA chlorination by HOCl (mean±SD in M-1 s-1: 4.3±0.3×103), the applied mathematical model for a chemical competition permits to obtain linear curves from competition studies between several NSAIDs and PABA. Their slopes provided the following rate constants for the different studied drugs: tenoxicam: 4.0±0.7×103, piroxicam: 3.6±0.7×103, lornoxicam: 4.3±0.7×103, meloxicam: 1.7±0.3×104, nimesulide: 2.3±0.6×102. Meloxicam therefore reacted significantly faster than the other oxicams and nimesulide, which is the weakest scavenger of the studied series. The identification of some of the oxidation products by NMR or MS permitted to explore the reaction mechanism and to examine some aspects of the structure/activity relationships for the molecules of the same chemical family.  相似文献   

12.
The role of microzooplankton in waters adjacent to Australia's North West Cape (21°49'S 114°14'E) was studied during the austral summers 1997/1998 and 1998/1999. We estimated microzooplankton abundance and biomass at a shallow (∼20 m) shelf station and at a shelf break station (∼80 m). Microzooplankton were placed into six categories: four ciliate groups (strombidiids, strobilidiids, tintinnids, “other ciliates”), dinoflagellates, and sarcodines. Total microzooplankton abundances ranged between 0.14×103 l-1 and 3.4×103 l-1. The most abundant groups were the dinoflagellates (mean 459±73 standard error l-1) and strombidiids (mean 334±42 standard error l-1). Total microzooplankton biomass ranged between 0.03 and 1.70 µg C l-1 (mean 0.33±0.05 standard error l-1). Redundancy analysis indicated differences in microzooplankton community composition between stations and sampling years but no differences with sampling depth. The microzooplankton community showed considerable variability between adjacent sampling dates, reinforcing the conclusion of earlier studies that this area is a dynamic environment. Ciliate production on the shelf was estimated to be 1.05 µg C l-1day-1 (∼20 mg C m-2 day-1) and 0.79 µg C l-1 day-1(∼70 mg C m-2 day-1) at the shelf break. Ciliate production near North West Cape was two- to six-fold higher than the rate of secondary production by juvenile copepods. Despite this, ciliate grazing appears to account for only ∼5% of primary production and ciliates do not appear to be a major conduit between primary producers and higher trophic levels in these waters.  相似文献   

13.
High-pressure liquid-chromatography and microcalorimetry have been used to determine equilibrium constants and enthalpies of reaction for the disproportionation reaction of adenosine 5′-diphosphate (ADP) to adenosine 5′-triphosphate (ATP) andadenosine 5′-monophosphate (AMP). Adenylate kinase was used to catalyze this reaction. The measurements were carried out over the temperature range 286 to 311 K, at ionic strengths varying from 0.06 to 0.33 mol kg−1, over the pH range 6.04 to 8.87, and over the pMg range 2.22 to 7.16, where pMg = -log a(Mg2+). The equilibrium model developed by Goldberg and Tewari (see the previous paper in this issue) was used for the analysis of the measurements. Thus, for the reference reaction: 2 ADp3− (ao) AMp2− (ao)+ ATp (ao), K° = 0.225 ± 0.010, ΔG° = 3.70 +- 0.11 kJ mol −1, ΔH° = −1.5 ± 1. 5 kJ mol −1, °S ° = −17 ± 5 J mol−1 K−1, and ACPp°≈ = −46 J mo1l−1 K−1 at 298.15 K and 0.1 MPa. These results and the thermodynamic parameters for the auxiliary equilibria in solution have been used to model the thermodynamics of the disproportionation reaction over a wide range of temperature, pH, ionic strength, and magnesium ion morality. Under approximately physiological conditions (311.15 K, pH 6.94, [Mg2+] = 1.35 × 10−3 mol kg−1, and I = 0.23 mol kg−1) the apparent equilibrium constant (KA′ = m(ΣAMP)m(ΣATP)/[ m(ΣADP)]2) for the overall disproportionation reaction is equal to 0.93 ± 0.02. Thermodynamic data on the disproportionation reaction and literature values for this apparent equilibrium constant in human red blood cells are used to calculate a morality of 1.94 × 10−4 mol kg−1 for free magnesium ion in human red blood cells. The results are also discussed in relation to thermochemical cycles and compared with data on the hydrolysis of the guanosine phosphates.  相似文献   

14.
A differential pH-termal titration apparatus is described which can detect pH differences with a sensitivity of ±0.0001 pH units and a thermal sensitivity of ±0.00002°C at a time constant of 0.1 s. With a reaction which yields 1 kcal mol−1, the current system can detect concentrations as low as 4×10−6 M or, in a 2 ml volume, a total amount of 40 nmol. With a time constant of 0.1 s, the sensitivity is 20±4 μ°C. The experimental protocol is specified by a microprocessor and three modes of operation are possible: titration at constant rate of reagent addition, titration at variable rates of addition so that the contents of both cells are at either constant pH or at a constant temperature and variable rate when a rate of change is specified. Experimental data are collected in files, corrected for heat loss, initial baseline drift, and changes in volume. The final corrected from the standardized run of 0.01338 M HCl in 0.2 M KCl at 25°C calibrate the pH scale yielded the calorimetric conversion constants and pKw which are calculated and stored for subsequent corrections for the titration of an unknown acid or the measurement of bindin constants and heats.  相似文献   

15.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


16.
The free radical scavenging properties of retinyl ascorbate (RA-AsA) were determined by monitoring the decomposition of 2,2-diphenyl-1-picrylhydrazyl (DPPH) as a function of time and in comparison with ascorbic acid (AsA), ascorbic acid palmitate (AsA-Pal), retinoic acid (RA), retinol (ROL) and retinol palmitate (Rol-Pal). The rate constant of RA-AsA (mean3±SD) was 4.9±0.3 M-1 s-1, and indicated greater potency as an antioxidant compared to the rest of the test compounds (AsA 3.4±0.4 M-1 s-1, AsA-Pal, 2.9±0.2 M-1 s-1, RA 1.4±0.3 M-1 s-1, ROL 1.3±0.1 M-1 s-1, Rol-Pal exhibited insignificant activity). The decomposition rate constant of DPPH, 5±0.6 × 10-8 M-1 s-1, in ethanol and BHA, 154±3 M-1 s-1 were both used as control. The compound RA-2-carboxy-2-hydroxy-ethanoate was isolated by prep-TLC and was identified, by 13C and 1HNMR spectroscopy, as the major by-product from the reaction of RA-AsA with DPPH, which was also found to be potent antioxidant, 2.1±0.2 M-1 s-1. This suggests that oxidation of AsA moiety did not lead to the production of erythrulose species, which could cause deleterious modifications of cellular proteins.  相似文献   

17.
The effect of Hb-I* phenotype on white muscle lactate dehydrogenease (LDH, E. C. 1.1.1.27) activity and buffering capacity was studied in Atlantic cod (Gadus morhua), acclimated and measured at temperatures near their behavioral temperature preference. It was hypothesized that these conditions would optimize biochemical processes but no difference was found in LDH activity between the Hb-I* phenotype after 56 d of acclimation to 6 and 14°C. However, LDH activity was both mass- and temperature-dependent; mean activity was 162.2±5.0 and 275.9±6.4 IU g-1 wet mass (mean±SEM) at 6 and 14°C respectively and larger fish had the highest rate of enzyme activity. White muscle buffer capacity was unaffected by Hb-I* phenotype but higher in cod held at 14°C.  相似文献   

18.
A Bacillus subtilis strain isolated from a hot-spring was shown to produce xylanolytic enzymes. Their associative/synergistic effect was studied using a culture medium with oat spelts xylan as xylanase inducer. Optimal xylanase production of about 12 U ml−1 was achieved at pH 6.0 and 50°C, within 18 h fermentation. At 50°C, xylanase productivity obtained after 11 h in shake-flasks, 96,000 U l−1 h−1, and in reactor, 104,000 U l−1 h−1 was similar. Increasing temperature to 55°C a higher productivity was obtained in the batch reactor 45,000 U l−1 h−1, compared to shake-flask fermentations, 12,000 U l−1 h−1. Optimal xylanolytic activity was reached at 60°C on phosphate buffer, at pH 6.0. The xylanase is thermostable, presenting full stability at 60°C during 3 h. Further increase in the temperature caused a correspondent decrease in the residual activity. At 90°C, 20% relative activity remains after 14 min. Under optimised fermentation conditions, no cellulolytic activity was detected on the extract. Protein disulphide reducing agents, such as DTT, enhanced xylanolytic activity about 2.5-fold. When is used xylan as substrate, xylanase production decreased as function of time in contrast, with trehalose as carbon source, xylanase production in maintained constant for at least 80 h fermentation.  相似文献   

19.
M.-E. Koller  I. Romslo  T. Flatmark 《BBA》1976,449(3):480-490
The mitochondrial ferrochelatase activity has been studied in coupled rat liver mitochondria using deuteroporphyrin IX (incorporated into liposomes of lecithin) and Fe(III) or Co(II) as the substrates.

1. 1. It was found that respiring mitochondria catalyze the insertion of Fe(II) and Co(II) into deuteroporphyrin. When Fe(III) was used as the metal donor, the reaction revealed an absolute requirement for a supply of reducing equivalents supported by the respiratory chain.

2. 2. A close correlation was found between the disappearance of porphyrin and the formation of heme which allows an accurate estimate of the extinction coefficient for the porphyrin to heme conversion. The value Δ (mM−1 · cm−1) = 3.5 for the wavelength pair 498 509 nm, is considerably lower than previously reported.

3. 3. The maximal rate of deuteroheme synthesis was found to be approx. 1 nM · min−1 · mg−1 of protein at 37 °C, pH 7.4 and optimal substrate concentrations, i.e. 75 μM Fe(III) and 50 μM deuteroporphyrin.

4. 4. Provided the mitochondria are supplemented with an oxidizable substrate, the presence of oxygen has no effect on the rate of deuteroheme synthesis.

Abbreviations: EPPS, (4-(2-hydroxyethyl)-1-piperazine propane sulphonic acid); HEPES, N-2-hydroxyethylpiperazine-N′-2-ethanesulphonic acid; PIPES, piperazine-N,N′-2-bis(2-ethanesulphonic acid)  相似文献   


20.
Rate constants determined by the stopped-flow method for four protein-protein reactions at 25°C, pH's in the range 5.8–7.5. I = 0.10 M (NaCI), are as follows: cytochrome c(II) with plastocyanin, PCu(II). 1.5 × 106 M−1 sec−1, pH 7.6; high-potential iron-sulfur protein (Hipip) with PCu(II), 3.7 × 105 M−1sec−1. pH 5.8; cytochrome c(II) with azurin, ACu(ll). 6.4 × 103 M−1sec−1, pH 6.1; Hipip with ACu(II), 2.2 × 105 M−1sec−1, pH 5.8. Activation parameters have been determined for all four reactions; they indicate higher enthalpy requirements and less negative entropy requirements for the PCu(II) as opposed to ACu(II) reactions. Equilibrium constants K for association prior to electron transfer are < 150 M−1 for the cytochrome c(II) reduction of PCu(II) (estimated charges 8 + and 9-,respectively), and < 300 M−1 for the other reactions, indicating no favorable interactions. Rate constants have been analyzed in terms of the simple Marcus theory, which has previously given an excellent fit to thirteen protein-protein reactions considered by Wherland and Pecht. No similar correlation exists in the present studies, and calculated rate constants differ by orders of magnitude from experimentally determined values.  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