首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Y Hao  J Peng  Y Ao  J Li  M Zhai 《Carbohydrate polymers》2012,90(4):1629-1633
The radiation processing of cellulose in ionic liquids (ILs) demands a comprehensive knowledge of radiation effects on cellulose in ILs. Herein, gamma radiation-induced degradation kinetics of microcrystalline cellulose (MCC) in 1-butyl-3-methylimidazolium chloride ([Bmim]Cl) was studied by viscometry. The intrinsic viscosity of MCC in [Bmim]Cl decreased slightly with increasing dose; while chemical structure and crystalline state of cellulose has no obvious change up to 300kGy. The radiation degradation rate constant (k) of MCC in [Bmim]Cl was 2.60×10(-7)/kGy, lower than that of solid cellulose, but higher than that in N-methylmorpholine-N-oxide (NMMO) solvent. Furthermore, k value decreased to 1.12×10(-7)/kGy in dimethyl sulfoxide (DMSO)/[Bmim]Cl system due to the free radicals scavenging of DMSO. The radicals generated during irradiation play main role in the radiation degradation of MCC in [Bmim]Cl. This work provides a new way to control the average molecular weight of cellulose by radiation-induced degradation of cellulose in ILs.  相似文献   

2.
Monochloramine disinfection kinetics were determined for the pure-culture ammonia-oxidizing bacterium Nitrosomonas europaea (ATCC 19718) by two culture-independent methods, namely, Live/Dead BacLight (LD) and propidium monoazide quantitative PCR (PMA-qPCR). Both methods were first verified with mixtures of heat-killed (nonviable) and non-heat-killed (viable) cells before a series of batch disinfection experiments with stationary-phase cultures (batch grown for 7 days) at pH 8.0, 25°C, and 5, 10, and 20 mg Cl2/liter monochloramine. Two data sets were generated based on the viability method used, either (i) LD or (ii) PMA-qPCR. These two data sets were used to estimate kinetic parameters for the delayed Chick-Watson disinfection model through a Bayesian analysis implemented in WinBUGS. This analysis provided parameter estimates of 490 mg Cl2-min/liter for the lag coefficient (b) and 1.6 × 10−3 to 4.0 × 10−3 liter/mg Cl2-min for the Chick-Watson disinfection rate constant (k). While estimates of b were similar for both data sets, the LD data set resulted in a greater k estimate than that obtained with the PMA-qPCR data set, implying that the PMA-qPCR viability measure was more conservative than LD. For N. europaea, the lag phase was not previously reported for culture-independent methods and may have implications for nitrification in drinking water distribution systems. This is the first published application of a PMA-qPCR method for disinfection kinetic model parameter estimation as well as its application to N. europaea or monochloramine. Ultimately, this PMA-qPCR method will allow evaluation of monochloramine disinfection kinetics for mixed-culture bacteria in drinking water distribution systems.As a result of stage 1 and stage 2 disinfectant and disinfection by-product rules, chloramination for secondary disinfection in the United States is predicted to increase to 57% of all surface and 7% of all groundwater treatment systems (49). A recent survey reported that 30% of the respondents currently chloraminate to maintain distribution system residual, and other recent surveys suggest that between 8 and 12% of drinking water utilities are contemplating a future switch to chloramination (3, 43).Although chloramines are considered weaker disinfectants than chlorine for suspended bacteria, chloramines are perceived as more effective disinfectants for a biofilm (25, 53). As a result of their lower reactivity, chloramines are believed to penetrate a biofilm further and thereby to more effectively disinfect biofilm bacteria with depth than chlorine (53).Chloramination comes with the risk of distribution system nitrification (2, 21, 22). Based on utility surveys, 30 to 63% of utilities practicing chloramination for secondary disinfection experience nitrification episodes (3, 21, 43, 54). Nitrification in drinking water distribution systems is undesirable and may result in water quality degradation (e.g., disinfectant depletion, coliform occurrences, or nitrite/nitrate formation) and subsequent noncompliance with existing regulations (e.g., surface water treatment rule or total coliform rule) (2). Thus, nitrification control is a major issue in practice and is likely to become increasingly important as chloramination increases.Unfortunately, our understanding of distribution system nitrification and its control is incomplete, which has made this a topic of considerable ongoing research. Recently, Fleming et al. (12) proposed nitrification potential curves as a possible strategy to prevent nitrification in chloraminated drinking water distribution systems. Use of this concept or other modeling approaches inherently requires knowledge of both the growth and disinfection kinetic parameters of nitrifiers, specifically ammonia-oxidizing bacteria (AOB), inhabiting the distribution system.Several chloramine disinfection studies have been reported for nitrifier cultures (2). However, only one study contains a detailed determination of chloramine disinfection kinetics, having investigated the pure-culture AOB Nitrosomonas europaea (33). In contrast to this pure-culture study, AOB are present as mixed cultures in chloraminated drinking water distribution systems, with Nitrosomonas oligotropha rather than N. europaea representing the dominant AOB found (33, 37, 38). Therefore, determination of disinfection kinetics of mixed-culture AOB likely present in chloraminated drinking water (i.e., N. oligotropha) represents a significant knowledge gap in our understanding of nitrification episodes.Disinfection kinetic parameter determination inherently depends on the method used to quantify viable bacteria. In general, there are two classes of viability determinations, i.e., (i) culture-dependent and (ii) culture-independent methods (5, 16, 27). Culture-dependent methods rely on bacterial growth and include plate counts and most-probable-number (MPN) techniques. Culture-independent methods include activity measures (e.g., substrate uptake or oxygen utilization) and other methods that rely on cell membrane integrity as a viability measure. In general, culture-dependent methods result in faster disinfection kinetics than culture-independent methods.As a first step toward gaining more information on AOB disinfection in chloraminated drinking water distribution systems, a culture-independent method with future applicability to mixed-culture AOB was implemented. In the current research, N. europaea was used. Even though N. europaea has not been found to be the dominant AOB in chloraminated systems, its use in the current research provides a comparison to existing literature. The culture-independent method combines the use of propidium monoazide (PMA), which selectively removes DNA from membrane-compromised cells and/or inhibits its amplification by PCR (29-31), with a quantitative PCR (qPCR) method developed for detection of AOB in chloraminated drinking water distribution systems (36). The results using PMA-qPCR were compared with those obtained using another culture-independent membrane integrity-based technique, the Live/Dead BacLight (LD) method. Furthermore, the experimental conditions were selected (pH 8.0 and a chlorine-to-nitrogen mass ratio of 4:1) such that monochloramine was the dominant chloramine species present, and the results are reported as monochloramine disinfection kinetics. The magnitude of the reported disinfection kinetics was closely related to the respective method used for viability determination. For example, in this research a cell was considered viable or nonviable based on the ability of propidium iodide (PI) or PMA to penetrate its membrane and on subsequent processing according to the respective method.LD was previously used to determine detailed N. europaea disinfection kinetics (33) and provides a baseline comparison for the current research. Oldenburg et al. (33) provided a comparison of estimated disinfection kinetic parameters, using both the culture-dependent AOB MPN technique and LD as viability measures. The estimated disinfection kinetic parameters based on the AOB MPN method were 3 orders of magnitude greater than those obtained with the culture-independent LD method, and the lower disinfection kinetics based on LD were more consistent with AOB persistence in chloraminated drinking water distribution systems. Based on this previous research and because the AOB MPN method requires an incubation period of 21 to 30 days, it was not evaluated in the current research (2).Initially, control experiments were conducted with various proportions of heat-killed cells to verify that both the PMA-qPCR and LD methods detected only viable cells. After the control experiments, a series of batch disinfection experiments were conducted where both PMA-qPCR and LD were utilized to quantify viable bacteria, providing two data sets for disinfection kinetic parameter estimation. Ultimately, the PMA-qPCR method used in this research will be applied to mixed-culture AOB typically present in drinking water distribution systems (i.e., N. oligotropha) (36-38).  相似文献   

3.
Fuculose-1-phosphate aldolase (FucA) is a useful biocatalyst with potential applications in chiral synthesis. In this study, the overall kinetic mechanism of FucA from the archaeon Methanococcus jannaschii was studied. The K(m) values of dihydroxyacetone phosphate (DHAP) and dl-glyceraldehyde were 0.09 and 0.74 mM, respectively. Dead-end inhibition by trimethyl phosphonoacetate and dl-threose were competitive and uncompetitive with respect to DHAP and dl-glyceraldehyde. Inhibition patterns obtained using reaction products were noncompetitive vs. DHAP and competitive vs. dl-glyceraldehyde. The equilibrium constant was 8.309×10(-3) M as assessed by varying the [DHAP]/[product] ratio at a fixed dl-glyceraldehyde concentration and by measuring the change in DHAP concentration after equilibrium was reached. This constant is consistent with the K(eq) value obtained from (13)C NMR (15.625×10(-3) M). The resultant inhibition kinetics may suggest the insights of kinetic mechanism of the FucA catalyzed reaction.  相似文献   

4.
The nitroxide tempol (4-hydroxy-2,2,6,6-tetramethyl piperidine-1-oxyl) reduces tissue injury in animal models of inflammation by mechanisms that are not completely understood. MPO (myeloperoxidase), which plays a fundamental role in oxidant production by neutrophils, is an important target for anti-inflammatory action. By amplifying the oxidative potential of H2O2, MPO produces hypochlorous acid and radicals through the oxidizing intermediates MPO-I [MPO-porphyrin?+-Fe(IV)=O] and MPO-II [MPO-porphyrin-Fe(IV)=O]. Previously, we reported that tempol reacts with MPO-I and MPO-II with second-order rate constants similar to those of tyrosine. However, we noticed that tempol inhibits the chlorinating activity of MPO, in contrast with tyrosine. Thus we studied the inhibition of MPO-mediated taurine chlorination by tempol at pH 7.4 and re-determined the kinetic constants of the reactions of tempol with MPO-I (k=3.5×105 M-1·s-1) and MPO-II, the kinetics of which indicated a binding interaction (K=2.0×10-5 M; k=3.6×10-2 s-1). Also, we showed that tempol reacts extremely slowly with hypochlorous acid (k=0.29 and 0.054 M-1·s-1 at pH 5.4 and 7.4 respectively). The results demonstrated that tempol acts mostly as a reversible inhibitor of MPO by trapping it as MPO-II and the MPO-II-tempol complex, which are not within the chlorinating cycle. After turnover, a minor fraction of MPO is irreversibly inactivated, probably due to its reaction with the oxammonium cation resulting from tempol oxidation. Kinetic modelling indicated that taurine reacts with enzyme-bound hypochlorous acid. Our investigation complements a comprehensive study reported while the present study was underway  相似文献   

5.
The bis(terpyridine)cobalt(II), Co(terpy)2(2+), reduction of cytochrome c peroxidase compound I, CcP-I, has been investigated using stopped-flow techniques as a function of ionic strength in pH 7.5 buffers at 25 degrees C. Co(terpy)2(2+) initially reduces the Trp191 radical site in CcP-I with an apparent second-order rate constant, k2, equal to 6.0+/-0.4x10(6) M(-1)s(-1) at 0.01 M ionic strength. A pseudo-first-order rate constant of 480 s(-1) was observed for the reduction of CcP-I by 79 microM Co(terpy)2(2+) at 0.01 M ionic strength. The one-electron reduction of CcP-I produces a second enzyme intermediate, CcP compound II (CcP-II), which contains an oxyferryl, Fe(IV), heme. Reduction of the Fe(IV) heme in CcP-II by Co(terpy)2(2+) shows saturation kinetics with a maximum observed rate constant, k3max, of 24+/-2 s(-1) at 0.01 M ionic strength. At low reductant concentrations, the apparent second-order rate constant for Co(terpy)2(2+) reduction of CcP-II, k3, is 1.2+/-0.5x10(6) M(-1) s-1. All three rate constants decrease with increasing ionic strength. At 0.10 M ionic strength, values of k2, k3, and k3max decrease to 6.0+/-0.8x10(5) M(-1) s(-1), 1.2+/-0.5x10(5) M(-1) s(-1), and 11+/-3 s(-1), respectively. Both the product, Co(terpy)2(3+), and ferricytochrome c inhibit the rate of Co(terpy)2(2+) reduction of CcP-I and CcP-II. Gel-filtration studies show that a minimum of two Co(terpy)2(3+) molecules bind to the native enzyme in low ionic strength buffers.  相似文献   

6.
Adenosine 5-phosphosulfate (APS) kinase from Penicillium chrysogenum is irreversibly inactivated by trinitrobenzene sulfonate in a pseudo-first order process. Under standard assay conditions kapp was 1.9 X 10(-3) s-1. Saturating MgATP or MgADP decreased Kapp to a limit of 4.1 X 10(-4) s-1. There are several explanations for the partial protection, including the presence of two essential lysyl side chains, only one of which is at the active site. Analysis of the inactivation kinetics by means of linear plots derived for partial protection yielded dissociation constants for E X MgATP (Kia) and E X MgADP (Kiq) of 2.9 mM and 1.8 mM, respectively. Low concentrations of APS alone provided no protection against trinitrobenzene sulfonate inactivation, but in the presence of 1 mM MgADP, as little as 2 microM APS provided additional protection while 100 microM APS reduced kapp to the limit of 4.1 X 10(-4) s-1. The results confirm the formation of a dead end E X MgADP X APS proposed earlier as the cause of the potent substrate inhibition by APS. Linear plots of 1/delta k versus 1/[MgADP] at different fixed [APS] and of 1/delta k versus 1/[APS] at different fixed [MgADP] were characteristic of the ordered binding of MgADP before APS (or the highly synergistic random binding of the two ligands). The true APS dissociation constant of the dead end E X MgADP X APS complex (K'ib) was determined to be 1.9 microM. From the value of K'ib and the previously reported value of KIB (apparent inhibition constant of APS as a substrate inhibitor of the catalytic reaction at saturating MgATP), the ratio of the MgADP and PAPS release rate constants (k4/k3) was calculated to be 11. Inactivation kinetics was used to study the effects of Mg2+ and high salt on ADP and APS binding. The results indicated that free ADP binds to the enzyme more tightly than does MgADP at low ionic strength. High salt decreased free ADP binding, but had little effect on MgADP binding. APS binds more tightly to E X MgADP in the absence or presence of salt than to E X ADP.  相似文献   

7.
Bovine pancreatic ribonuclease A (RNase A) catalyzes the cleavage of the P-O(5') bond in RNA. Although this enzyme has been the object of much landmark work in bioorganic chemistry, the nature of its rate-limiting transition state and its catalytic rate enhancement had been unknown. Here, the value of k(cat)/K(m) for the cleavage of UpA by wild-type RNase A was found to be inversely related to the concentration of added glycerol. In contrast, the values of k(cat)/K(m) for the cleavage of UpA by a sluggish mutant of RNase A and the cleavage of the poor substrate UpOC(6)H(4)-p-NO(2) by wild-type RNase A were found to be independent of glycerol concentration. Yet, UpA cleavage by the wild-type and mutant enzymes was found to have the same dependence on sucrose concentration, indicating that catalysis of UpA cleavage by RNase A is limited by desolvation. The rate of UpA cleavage by RNase A is maximal at pH 6.0, where k(cat) = 1.4 × 10(3) s(-1) and k(cat)/K(m) = 2.3 × 10(6) M(-1)s(-1) at 25°C. At pH 6.0 and 25°C, the uncatalyzed rate of [5,6-(3)H]Up[3,5,8-(3)H]A cleavage was found to be k(uncat) = 5 × 10(-9) s(-1) (t(1/2) = 4 years). Thus, RNase A enhances the rate of UpA cleavage by 3 × 10(11)-fold by binding to the transition state for P-O(5') bond cleavage with a dissociation constant of <2 × 10(-15) M.  相似文献   

8.
The stimulation of inositol phosphate generation by bombesin and GTP analogues was studied in Swiss 3T3 cells permeabilized by electroporation. Bombesin-stimulated inositol phosphate generation is potentiated by guanosine 5'-[gamma-thio]triphosphate (GTP[S]) and inhibited by guanosine 5'-[beta-thio]diphosphate at all peptide concentrations tested, with no change in the EC50 value (concn. giving half-maximal response) for the agonist. Kinetic analysis showed that, although bombesin-stimulated [3H]InsP3 generation in [3H]inositol-labelled cells was rapid (maximal by 5-10 s), the response to GTP[S] alone displayed a distinct lag time of 20-30 s. This lag time was significantly decreased by the addition of bombesin, suggesting that in this system agonist-stimulated GTP/GDP exchange occurs. In addition, bombesin-stimulated generation of Ins(1,4,5)P3 mass at 10 s was enhanced by GTP[S] in the absence of a nucleotide response alone, a result consistent with this proposal. Pretreatment of the cells with phorbol 12-myristate 13-acetate (PMA) resulted in a dose-dependent inhibition of bombesin-, but not GTP[S]-, stimulated inositol phosphate generation. Furthermore, although PMA pretreatment did not affect the lag time for InsP3 formation in response to GTP[S] alone, the degree of synergy between bombesin and the nucleotide was severely decreased at early time points. The results therefore demonstrate that the high-affinity bombesin receptor is coupled via a G-protein to phospholipase C in a manner consistent with a general model for receptor-G-protein interactions and that this coupling is sensitive to phosphorylation by protein kinase C.  相似文献   

9.
A radioisotope method was devised to study bacterial respiratory reduction of arsenate in sediments. The following two arsenic-rich soda lakes in California were chosen for comparison on the basis of their different salinities: Mono Lake (approximately 90 g/liter) and Searles Lake (approximately 340 g/liter). Profiles of arsenate reduction and sulfate reduction were constructed for both lakes. Reduction of [73As]arsenate occurred at all depth intervals in the cores from Mono Lake (rate constant [k] = 0.103 to 0.04 h(-1)) and Searles Lake (k = 0.012 to 0.002 h(-1)), and the highest activities occurred in the top sections of each core. In contrast, [35S]sulfate reduction was measurable in Mono Lake (k = 7.6 x10(4) to 3.2 x 10(-6) h(-1)) but not in Searles Lake. Sediment DNA was extracted, PCR amplified, and separated by denaturing gradient gel electrophoresis (DGGE) to obtain phylogenetic markers (i.e., 16S rRNA genes) and a partial functional gene for dissimilatory arsenate reduction (arrA). The amplified arrA gene product showed a similar trend in both lakes; the signal was strongest in surface sediments and decreased to undetectable levels deeper in the sediments. More arrA gene signal was observed in Mono Lake and was detectable at a greater depth, despite the higher arsenate reduction activity observed in Searles Lake. A partial sequence (about 900 bp) was obtained for a clone (SLAS-3) that matched the dominant DGGE band found in deeper parts of the Searles Lake sample (below 3 cm), and this clone was found to be closely related to SLAS-1, a novel extremophilic arsenate respirer previously cultivated from Searles Lake.  相似文献   

10.
The ionic species of cytosol estrogen receptors from mature rat uteri have been compared by HPIEC on a SynChrom AX-1000 column when complexed with either [125I]iodoestradiol, [3H]estradiol or [3H]4-hydroxy tamoxifen. Three species of receptors (isoforms) each suppressible by excess competitor were fractionated at identical salt concentrations regardless of ligand employed. One species eluted in the column void volume (10 mM) and the others at congruent to 90 mM and congruent to 155 mM phosphate. Activation of receptor complexes by increasing time of incubation with ligand from 1 to 24 h at 4 degrees C or addition of 10 mM GTP increased the proportion of the congruent to 90 mM species for all 3 ligands. The addition of 10 mM molybdate to homogenization and HPIEC buffers resulted in only two species being resolved at 10 and 110 mM phosphate. These species were again identical regardless of ligand employed. Increasing concentrations of estradiol (1-40 nM) tamoxifen (20 nM-4 microM) and 4-hydroxy tamoxifen (2-400 nM) were able to compete for binding of [125I]iodoestradiol to each of the three ionic species. Binding to each species was inhibited equally by each concentration of competitor. There was no preferential or unique association of estrogen or antiestrogen with any of the ionic species and all ligands gave identical ionic species of non-activated and activated receptor complexes.  相似文献   

11.
Beta-Adrenergic receptors were studied in intact cells of chick, rat and mouse embryo brain in primary cultures, by the specific binding of [3H]dihydro-L-alprenolol ([3H]DHA). The results were compared to the receptor binding of broken cell preparations derived from the cell cultures or from the forebrain tissues used for the preparation of the cultures. Detailed analysis of [3H]DHA binding to living chick brain cells revealed a high-affinity, stereoselective, beta-adrenergic-type binding site. Equilibrium measurements indicated the apparent positive cooperativity of the binding reaction. By direct fitting of the Hill equation to the measured data, values of Bmax = 12.01 fmol/10(6) cells (7200 sites/cell), Kd = 60.23 pM and the Hill coefficient n = 2.78 were found. The apparent cooperative character of the binding was confirmed by the kinetics of competition with L-alprenolol, resulting in maximum curves at low ligand concentrations. The rate constants of the binding reaction were estimated as k+ = 8.31 X 10(7) M-1 X min-1 and k- = 0.28 min-1 from the association results, and k- = 0.24 min-1 from the dissociation data. The association kinetics supported the cooperativity of the binding, providing a Hill coefficient n = 1.76; Kd, as (k-/k+)1/n was found to be 101 pM. Analysis of the equilibrium binding of [3H]DHA to rat and mouse living brain cells resulted in values of Bmax = 13.04 fmol/10(6) cells (7800 sites/cell), Kd = 43.85 pM and n = 2.52, and Bmax = 8.08 fmol/10(6) cells (4800 sites/cell), Kd = 46.70 pM and n = 1.63, respectively, confirming the apparent cooperativity of the beta-receptor in mammalian objects, too. The [3H]DHA equilibrium binding to broken cell preparations of either chick, rat or mouse brain cultures or forebrain tissues was found to be non-cooperative, with a Hill coefficient n = 1, Kd in the range 1-2 nM, and a Bmax of 10(3) - 10(4) sites/cell. Our findings demonstrate that cell disruption causes marked changes in the kinetics of the beta-receptor binding and in the affinity of the binding site, although the number of receptors remains unchanged.  相似文献   

12.
We explored the intra- and extracellular processes governing the kinetics of extracellular ATP (ATPe) in human erythrocytes stimulated with agents that increase cAMP. Using the luciferin-luciferase reaction in off-line luminometry we found both direct adenylyl cyclase activation by forskolin and indirect activation through β-adrenergic stimulation with isoproterenol-enhanced [ATP]e in a concentration-dependent manner. A mixture (3V) containing a combination of these agents and the phosphodiesterase inhibitor papaverine activated ATP release, leading to a 3-fold increase in [ATP]e, and caused increases in cAMP concentration (3-fold for forskolin + papaverine, and 10-fold for 3V). The pannexin 1 inhibitor carbenoxolone and a pannexin 1 blocking peptide ((10)Panx1) decreased [ATP]e by 75-84%. The residual efflux of ATP resulted from unavoidable mechanical perturbations stimulating a novel, carbenoxolone-insensitive pathway. In real-time luminometry experiments using soluble luciferase, addition of 3V led to an acute increase in [ATP]e to a constant value of ~1 pmol × (10(6) cells)(-1). A similar treatment using a surface attached luciferase (proA-luc) triggered a rapid accumulation of surface ATP levels to a peak concentration of 2.4 pmol × (10(6) cells)(-1), followed by a slower exponential decay (t(½) = 3.7 min) to a constant value of 1.3 pmol × (10(6) cells)(-1). Both for soluble luciferase and proA-luc, ATP efflux was fully blocked by carbenoxolone, pointing to a 3V-induced mechanism of ATP release mediated by pannexin 1. Ecto-ATPase activity was extremely low (~28 fmol × (10(6) cells min)(-1)), but nevertheless physiologically relevant considering the high density of erythrocytes in human blood.  相似文献   

13.
The kinetics of the decomposition of H(2)O(2) catalyzed by Cu(II) has been studied by the initial-rate method in aqueous phosphate media at near physiological pH. The activity of the catalyst is increased by [Fe(CN)(6)](3-) and decreased by VO(3)(-), CrO(4)(2-) and Zn(II). Three reaction pathways are involved in the Cu(II)-H(2)O(2) reaction, the kinetic orders of the catalyst being 1 (rate constant k1), 2 (rate constant k2) and 3 (rate constant k3). The three pathways present fractional apparent orders (>1) in H(2)O(2) and base catalysis. The apparent activation energies associated to rate constants k1, k2 and k3 are 102+/-4, 65+/-8 and 61+/-5 kJ mol(-1). Free-radical chain mechanisms are proposed for the three pathways.  相似文献   

14.
The reaction between cytochrome c1 and cytochrome c   总被引:3,自引:0,他引:3  
The kinetics of electron transfer between the isolated enzymes of cytochrome c1 and cytochrome c have been investigated using the stopped-flow technique. The reaction between ferrocytochrome c1 and ferricytochrome c is fast; the second-order rate constant (k1) is 3.0 . 10(7) M-1 . s-1 at low ionic strength (I = 223 mM, 10 degrees C). The value of this rate constant decreases to 1.8 . 10(5) M-1 . s-1 upon increasing the ionic strength to 1.13 M. The ionic strength dependence of the electron transfer between cytochrome c1 and cytochrome c implies the involvement of electrostatic interactions in the reaction between both cytochromes. In addition to a general influence of ionic strength, specific anion effects are found for phosphate, chloride and morpholinosulphonate. These anions appear to inhibit the reaction between cytochrome c1 and cytochrome c by binding of these anions to the cytochrome c molecule. Such a phenomenon is not observed for cacodylate. At an ionic strength of 1.02 M, the second-order rate constants for the reaction between ferrocytochrome c1 and ferricytochrome c and the reverse reaction are k1 = 2.4 . 10(5) M-1 . s-1 and k-1 = 3.3 . 10(5) M-1 . s-1, respectively (450 mM potassium phosphate, pH 7.0, 1% Tween 20, 10 degrees C). The 'equilibrium' constant calculated from the rate constants (0.73) is equal to the constant determined from equilibrium studies. Moreover, it is shown that at this ionic strength, the concentrations of intermediary complexes are very low and that the value of the equilibrium constant is independent of ionic strength. These data can be fitted into the following simple reaction scheme: cytochrome c2+1 + cytochrome c3+ in equilibrium or formed from cytochrome c3+1 + cytochrome c2+.  相似文献   

15.
Escherichia coli DNA photolyase (photoreactivating enzyme) is a flavoprotein. The enzyme binds to DNA containing pyrimidine dimers in a light-independent step and, upon illumination with 300-600 nm radiation, catalyzes the photosensitized cleavage of the cyclobutane ring thus restoring the integrity of the DNA. We have studied the binding reaction using the techniques of nitrocellulose filter binding and flash photolysis. The enzyme binds to dimer-containing DNA with an association rate constant k1 estimated by two different methods to be 1.4 X 10(6) to 4.2 X 10(6) M-1 S-1. The dissociation of the enzyme from dimer-containing DNA displays biphasic kinetics; for the rapidly dissociating class of complexes k2 = 2-3 X 10(-2) S-1, while for the more slowly dissociating class k2 = 1.3 X 10(-3) to 6 X 10(-4) S-1. The equilibrium association constant KA, as determined by the nitrocellulose filter binding assay and the flash photolysis assay, was 4.7 X 10(7) to 6 X 10(7) M-1, in reasonable agreement with the values predicted from k1 and k2. From the dependence of the association constant on ionic strength we conclude that the enzyme contacts no more than two phosphodiester bonds upon binding; this strongly suggests that the pyrimidine dimer is the main structural determinant of specific photolyase-DNA interaction and that nonspecific ionic interactions do not contribute significantly to substrate binding.  相似文献   

16.
The kinetics of electron transfer from reduced high-potential iron-sulfur protein (HiPIP) to the photooxidized tetraheme cytochrome c subunit (THC) bound to the photosynthetic reaction center (RC) from the purple sulfur bacterium Allochromatium vinosum were studied under controlled redox conditions by flash absorption spectroscopy. At ambient redox potential Eh = +200 mV, where only the high-potential (HP) hemes of the THC are reduced, the electron transfer from HiPIP to photooxidized HP heme(s) follows second-order kinetics with rate constant k = (4.2 +/- 0.2) 10(5) M(-1) s(-1) at low ionic strength. Upon increasing the ionic strength, k increases by a maximum factor of ca. 2 at 640 mM KCl. The role of Phe48, which lies on the external surface of HiPIP close to the [Fe4S4] cluster and presumably on the electron transfer pathway to cytochrome heme(s), was investigated by site-directed mutagenesis. Substitution of Phe48 with arginine, aspartate, and histidine completely prevents electron donation. Conversely, electron transfer is still observed upon substitution of Phe48 with tyrosine and tryptophan, although the rate is decreased by more than 1 order of magnitude. These results suggest that Phe48 is located on a key protein surface patch essential for efficient electron transfer, and that the presence of an aromatic hydrophobic residue on the putative electron-transfer pathway plays a critical role. This conclusion was supported by protein docking calculations, resulting in a structural model for the HiPIP-THC complex, which involves a docking site close to the LP heme farthest from the bacteriochlorophyll special pair.  相似文献   

17.
M R Mauk  L S Reid  A G Mauk 《Biochemistry》1982,21(8):1843-1846
The interaction between cytochrome c and the tryptic fragment of cytochrome b5 has been found to produce a difference spectrum in the Soret region with a maximum absorbance at 416 nm. The intensity of this difference has been used to determine the stoichiometry of complex formation and the stability of the complex formed. At pH 7.0 [25 degrees C (phosphate), mu = 0.01 M], the two proteins were found to form a 1:1 complex with an association constant, KA, of 8(3) x 10(4) M-1. The stability of the complex was found to be strongly dependent on ionic strength with KA increasing to 4(3) x 10(6) M-1 at mu = 0.001 M [25 degrees C, pH 7.0 (phosphate)]. Analysis of the dependence of KA on pH from pH 6.5 to 8 demonstrated that this complex is maximally stable between pH 7 and 8 or about midway between the isoelectric points of the two proteins. Analysis of the temperature dependence of KA revealed that formation of the complex between the two proteins is largely entropic in origin with delta Ho = 1 +/- 3 kcal/mol and delta So = 33 +/- 11 eu [pH 7.0 (phosphate), mu = 0.001 M]. This result may be explained either by the model of Clothia and Janin [Clothia, C., & Janin, J. (1975) Nature (London) 256, 705] in terms of extensive solvent reorganization upon complexation or by the model of Ross and Subramanian [Ross, P. D., & Subramanian, S. (1981) Biochemistry 20, 3096] in which the negative enthalpic and entropic contributions of short-range protein-protein interactions are offset by proton release.  相似文献   

18.
The redox potential of the Rieske [2Fe-2S] cluster of the bc1 complex from bovine heart mitochondria was determined by cyclic voltammetry of a water-soluble fragment of the iron/sulfur protein. At the nitric-acid-treated bare glassy-carbon electrode, the fragment gave an immediate and stable quasireversible response. The midpoint potential at pH 7.2, 25 degrees C and I of 0.01 M was Em = +312 +/- 3 mV. This value corresponds within 20 mV to results of an EPR-monitored dye-mediated redox titration. With increasing ionic strength, the midpoint potential decreased linearly with square root of I up to I = 2.5 M. From the cathodic-to-anodic peak separation, the heterogeneous rate constant, k degrees, was calculated to be approximately 2 x 10(-3) cm/s at low ionic strength; the rate constant increased with increasing ionic strength. From the temperature dependence of the midpoint potential, the standard reaction entropy was calculated as delta S degrees = -155 J.K-1.mol-1. The pH dependence of the midpoint potential was followed over pH 5.5-10. Above pH 7, redox-state-dependent pK changes were observed. The slope of the curve, -120 mV/pH above pH9, indicated two deprotonations of the oxidized protein. The pKa values of the oxidized protein, obtained by curve fitting, were 7.6 and 9.2, respectively. A group with a pKa,ox of approximately 7.5 could also be observed in the optical spectrum of the oxidized protein. Redox-dependent pK values of the iron/sulfur protein are considered to be essential for semiquinone oxidation at the Qo center of the bc1 complex.  相似文献   

19.
The reduction by dithionite ion (in excess) of methemerythrin-anion adducts, Hr+X-, to deoxyhemerythrin, Hr degree, has been examined at 25 degrees and pH 6.3 and 8.2. The results accord with the scheme: S2O42- in equilibrium 2SO2- rapid Hr+X- in equilibrium Hr++X- k-1, k1 Hr++SO2- leads to PRODUCT k2 with X- = Br-, HCO2-, CNO-, and F-, k2[SO2-] greater than k1[X-], and the pseudo first-order rate constant, kobs (= k-1), is independent of [X-] and [S2O42-]. Only with X- = NCS- is k2[SO2-] approximately k1[X-] and kobs = a[S2O42-]1/2 (b[NCS-] + [S2OR2-]1/2)-1. Values at pH 6.3 of k-1 (sec-1) and k1 (M-1 sec-1), obtained by anation and anion displacement reactions, are 2.3 x 10(-3), 1.6 x 10(-2) (Br-); 1.5 x 10(-3), 1.2 x 10(-2) (HCO2-); 1.3 x 10(-4), 0.52 (CNO-) and approximately 2 x 10(-4), 3.3 x 10(-3) (CN-, pH 7.0). Values of k-1 from reduction and displacement methods are in good agreement with each other. The value of k2 (1.6 x 10(5) M-1 sec-1, pH 6.3) in somewhat smaller than that for reduction of the met form of hemoproteins. There is only a small effect of pH on rates. Direct reduction of Hr+CN- does not occur, in contrast with Mb+CN-.  相似文献   

20.
We previously showed, during quasi-steady-state exposures, that the rate of inhaled NO2 uptake displays reaction-mediated characteristics (J. Appl. Physiol. 68: 594-603, 1990). In vitro kinetic studies of pulmonary epithelial lining fluid (ELF) demonstrated that NO2 interfacial transfer into ELF exhibits first-order kinetics with respect to NO2, attains [NO2]-dependent rate saturation, and is aqueous substrate dependent (J. Appl. Physiol. 71: 1502-1510, 1991). We have extended these observations by evaluating the kinetics of NO2 gas phase disappearance in isolated ventilating rat lungs. Transient exposures (2-3/lung at 25 degrees C) employed rebreathing (NO2-air) from a non-compliant continuously stirred closed chamber. We observed that 1) NO2 uptake rate is independent of exposure period, 2) NO2 gas phase disappearance exhibited first-order kinetics [initial rate (r*) saturation occurred when [NO2] > 11 ppm], 3) the mean effective rate constant (k*) for NO2 gas phase disappearance ([NO2] < or = 11 ppm, tidal volume = 2.3 ml, functional residual capacity = 4 ml, ventilation frequency = 50/min) was 83 +/- 5 ml/min, 4) with [NO2] < or = 11 ppm, k* and r* were proportional to tidal volume, and 5) NO2 fractional uptakes were constant across [NO2] (< or = 11 ppm) and tidal volumes but exceeded quasi-steady-state observations. Preliminary data indicate that this divergence may be related to the inspired PCO2. These results suggest that NO2 reactive uptake within rebreathing isolated lungs follows first-order kinetics and displays initial rate saturation, similar to isolated ELF.(ABSTRACT TRUNCATED AT 250 WORDS)  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