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 62 毫秒
1.
The effect of pretreatment with ultraviolet-B (UV-B) light (280-320 nanometers) on the enzymatic conversion of the diepoxyxanthophyll violaxanthin to the epoxy-free zeaxanthin occurring in thylakoid membranes was investigated. When isolated chloroplasts of pea (Pisum sativum) were exposed to UV-B, a biologically effective fluence of 7000 joules per square meter caused about 50% inhibition of the activity of the violaxanthin deepoxidase, measured as the first order rate constant of the absorbance change at 505 nanometers. The dose requirement for the inhibition of the deepoxidase in intact leaves, however, was about 2 orders of magnitude higher. The inhibition of the rate constant was observed for both the dark deepoxidation at pH 5, and for the light-driven deepoxidation induced by the lumen acidification due to electron transport from H2O to methylviologen or due to a photosystem I partial reaction with duroquinol as the electron donor. The availability of violaxanthin was not directly affected by UV-B radiation, as shown for UV-B-treated chloroplasts by the final extent of the 505 nanometer change measured in the dark at pH 5 or by the partial photosystem I reaction. A significant decrease in the violaxanthin availability was observed when lumen acidification was caused by electron transport from H2O to methylviologen. That effect was probably caused by the wellknown UV-B inhibition of photosystem II with a subsequent decreased ability to reduce the plastoquinone pool, the redox state of which is believed to regulate the final amount of converted violaxanthin.  相似文献   

2.
The effects of dithiothreitol on absorbance changes at 505 and 515 nm in isolated lettuce chloroplasts were investigated. Dithiothreitol inhibited the ascorbate-dependent 505-nm change that is due to the de-epoxidation of violaxanthin to zeaxanthin. Dithiothreitol was effective for both light-induced de-epoxidation at pH 7 and dark de-epoxidation at pH 5. Titration of de-epoxidase activity with dithiothreitol resulted in complete inhibition at about 5 μmoles dithiothreitol per mg chlorophyll. Removal of dithiothreitol restored de-epoxidase activity. These results are consistent with the view that dithiothreitol inhibits violaxanthin de-epoxidation and the corresponding 505-nm change by reducing a disulfide that is required for de-epoxidase activity.

Dithiothreitol was effective in resolving absorbance changes due to violaxanthin de-epoxidation and other changes that were superimposed under some conditions. At 515 nm and in the presence of 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), phenazine methosulfate, and ascorbate, dithiothreitol inhibited the large, slow and irreversible change which was due to de-epoxidation but not the fast and reversible so-called 515-nm change. At 505 nm and under similar conditions, dithiothreitol revealed the presence of a slow reversible change in addition to the one from de-epoxidation. Results with dithiothreitol showed that the absorbance change at 505 nm in the presence of DCMU, 2,6-dichlorophenolindophenol and ascorbate was due entirely to de-epoxidation. Similarly, absorbance changes at 515 nm also appeared to be mainly from de-epoxidation but with the presence of a small transient change due to some other components. It is suggested that dithiothreitol may be useful in resolving complex light-induced absorbance changes in other photosynthetic systems as well as in enabling new studies on reversible absorbance changes in the 500-nm region.  相似文献   


3.
When cotton (Gossypium hirsutum L., cv Acaia SJC-1) leaves kept in weak light were suddenly exposed to strong red actinic light a spectral absorbance change took place having the following prominent characteristics. (a) It was irreversible within the first four minute period after darkening. (b) The difference in leaf absorbance between illuminated and predarkened leaves had a major peak at 505 nanometers, a minor peak at 465 nanometers, a shoulder around 515 nanometers, and minor troughs at 455 and 480 nanometers. (c) On the basis of its spectral and kinetic characteristics this absorbance change can be readily distinguished from the much faster electrochromic shift which has a peak at 515 nanometers, from the slow, so-called light-scattering change which has a broad peak centered around 535 nanometers and is reversed upon darkening, and from absorbance changes associated with light-induced chloroplast rearrangements. (d) The extent and time course of this absorbance change closely matched that of the deepoxidation of violaxanthin to zeaxanthin in the same leaves. (e) Both the absorbance change and the ability to form zeaxanthin were completely blocked in leaves to which dithiothreitol (DTT) had been provided through the cut petlole. DTT treatment also caused strong inhibition of that component of the 535-nanometer absorbance change which is reversed in less than 4 minutes upon darkening and considered to be caused by increased light scattering. Moreover, DTT inhibited a large part of nonphotochemical quenching of chlorophyll fluorescence in the presence of excessive light. However, DTT had no detectable effect on the photon yield of photosynthesis measured under strictly rate-limiting photon flux densities or on the light-saturated photosynthetic capacity, at least in the short term. We conclude that it is possible to monitor light-induced violaxanthin de-epoxidation in green intact leaves by measurement of the absorbance change at 505 nanometers. Determination of absorbance changes in conjunction with measurements of photosynthesis in the presence and absence of DTT provide a system well suited for future studies of meachanisms of dissipation of excessive excitation energy in intact leaves.  相似文献   

4.
A new ascorbate-induced chloroplast absorbance change which has the characteristics of a carotenoid shift is described. The absorbance change was light-dependent at pH 7 but not at pH 5. The difference spectra for the light and dark changes were similar, showing a large absorbance peak at 505 nanometers, smaller peaks near 468 and 437 nanometers, and a sharp valley around 483 nanometers. The absorbance change is assigned to violaxanthin de-epoxidation because various conditions affected the absorbance change and violaxanthin de-epoxidation similarly, and the difference spectrum resembled the spectrum of zeaxanthin minus violaxanthin in organic solvent.  相似文献   

5.
Kinetics of the NO binding and removal reaction of recombinant Prolixin-S (rProlixin-S) were analyzed using stopped-flow spectrophotometry. The reaction was observed as a biphasic process. The rate constant of the fast phase increased linearly as NO concentration increased. The rate constant at the slow phase increased as NO concentrations increased at low NO concentration, then reached a plateau at high NO concentration. These NO dependencies of the reaction are characteristic of a bimolecular two-step consecutive reaction. The reaction consisted of the fast NO binding reaction of rProlixin-S and the following slow structural change of NO-protein complex. Kinetic studies revealed that the NO binding rate constant was independent of pH, but the rate constant of the NO removal reaction increased as pH increased. The apparent NO dissociation constant (Kd) of rProlixin-S was also calculated from the values of the kinetic parameters obtained in this work. The Kd value increased as pH and temperature increased. The Kd value of rProlixin-S and NO was 10-300 nM in regular physiological condition, which is 103 higher and 103 lower than those of the other ferric and ferrous hemoproteins and NO, respectively. These results indicate that Prolixin-S is one of NO transport proteins regulating blood pressure.  相似文献   

6.
Frommolt R  Goss R  Wilhelm C 《Planta》2001,213(3):446-456
In vivo the prasinophyceaen alga Mantoniella squamata Manton et Parke uses an incomplete violaxanthin (Vx) cycle, leading to a strong accumulation of antheraxanthin (Ax) under conditions of high light. Here, we show that this zeaxanthin (Zx)-depleted Vx/Ax cycle is caused by an extremely slow second de-epoxidation step from Ax to Zx, and a fast epoxidation from Ax back to Vx in the light. The rate constant of Ax epoxidation is 5 to 6 times higher than the rate constant of Zx formation, implying that Ax is efficiently converted back to Vx before it can be de-epoxidated to Zx. It is, however, only half the rate constant of the first de-epoxidation step from Vx to Ax, thus explaining the observed net accumulation of Ax during periods of strong illumination. When comparing the rate constant of the second de-epoxidation step in M. squamata with Zx formation in spinach (Spinacia oleracea L.) thylakoids, we find a 20-fold reduction in the reaction kinetics of the former. This extremely slow Ax de-epoxidation, which is also exhibited by the isolated Mantoniella violaxanthin de-epoxidase (VDE), is due to a reduced substrate affinity of M. squamata VDE for Ax compared with the VDE of higher plants. Mantoniella VDE, which has a similar Km value for Vx, shows a substantially increased Km for the substrate Ax in comparison with spinach VDE. Our results furthermore explain why Zx formation in Mantoniella cells can only be found at low pH values that represent the pH optimum of VDE. A pH of 5 blocks the epoxidation reaction and, consequently, leads to a slow but appreciable accumulation of Zx.  相似文献   

7.
The soluble cytochrome o from Vitreoscilla contains two identical subunits and two hemes. The reduced form binds 2 mol of CO in a cooperative manner with a Hill coefficient near 2 (Tyree, B., and Webster, D. A. (1978) J. Biol. Chem. 253, 6988-6991). This carbonyl compound was photolysed with a dye laser and recombination followed at 437 or 420 nm where maximal absorbance changes were registered. Recombination kinetics were biphasic, and the fast phase was approximately 10 times the rate of the slow phase. Apparent rate constants of both phases showed a nonlinear dependence on CO concentration, respectively, in conformity with a reaction scheme which assumes the transient formation of an intermediate species in both slow and fast reactions. A study of temperature dependence of the reactions gave EA = 2.7 kcal/mol for the slow reaction and EA = 3.2 kcal/mol for the fast reaction below 23 degrees C; above this temperature the slope of the Arrhenius plot for the fast reaction became positive. Maximal rates for both phases were around pH 6.5 and fell to approximately 40% of maximal at pH 12. The binding reaction was affected by even a low concentration of sodium dodecyl sulfate (0.0025%), which changed both the kinetic constant of each phase and the relative contribution of each phase to the reaction. A model which assumes the existence of fast and slow reaction conformers in equilibrium is proposed.  相似文献   

8.
Light-driven violaxanthin deepoxidation was measured in isolated pea (Pisum sativum) chloroplasts without ATP synthesis (basal conditions) and with ATP synthesis (coupled conditions). Thylakoids stored in high salt (HS) or low salt (LS) storage medium were tested. In previous experiments, HS thylakoids and LS thylakoids were related to delocalized and localized proton coupling, respectively.Light-driven deepoxidase activity was compared to the pH dependence of deepoxidase activity established in dark reactions. At an external pH of 8, light-driven deepoxidation indicated effective pH values close to pH 6 for all reaction conditions. Parallel to deepoxidation, the thylakoid lumen pH was estimated by the fluorescent dye pyranine.In LS thylakoids under coupled conditions the lumen pH did not drop below pH 6.7. At pH 6.7, no deepoxidase activity is expected based on the pH dependence of enzyme activity. The results suggest that deepoxidation activity is controlled by the pH in sequestered membrane domains, which, under localized proton coupling, can be maintained at pH 6.0 when the lumen pH is far above pH 6.0. The extent of violaxanthin conversion (availability), however, appeared to be regulated by lumenal pH. Dithiothreitol-sensitive nonphotochemical quenching of chlorophyll fluorescence was dependent on zeaxanthin and not related to lumenal pH. Thus, zeaxanthin-dependent quenching[mdash]known to be pH dependent[mdash]appeared to be triggered by the pH of localized membrane domains.  相似文献   

9.
pH dependence of folding of iso-2-cytochrome c   总被引:4,自引:0,他引:4  
B T Nall  J J Osterhout  L Ramdas 《Biochemistry》1988,27(19):7310-7314
Starting from a standard unfolded state (3.0 M guanidine hydrochloride, pH 7.2), the kinetics of refolding of iso-2-cytochrome c have been investigated as a function of final pH between pH 3 and pH 10. Absorbance in the ultraviolet and visible spectral regions and tryptophan fluorescence are used to monitor folding. Over most of the pH range, fast and slow folding phases are detected by both fluorescence and absorbance probes. Near neutral pH, the rate of fast folding appears to be the same when monitored by absorbance and fluorescence probes. At higher and lower pH, there are two fast folding reactions, with absorbance-detected fast folding occurring in a slightly faster time range than fluorescence-detected fast folding. The rates of both fast folding reactions pass through broad minima near neutral pH, indicating involvement of ionizable groups in rate-limiting steps. The rates of slow folding also depend on the final pH. At acid pH, there appears to be a single slow folding phase for both fluorescence and absorbance probes. At neutral pH, the absorbance-detected and fluorescence-detected slow folding phases separate into distinct kinetic processes which differ in rate and relative amplitude. At high pH, absorbance-detected slow folding is no longer observed, while fluorescence-detected slow folding is decreased in amplitude. In contrast, the equilibrium and kinetic properties of proline imide bond isomerization, believed to be involved in the slow folding reactions, are largely independent of pH. The results suggest that the pH dependence of slow folding involves coupling of pH-sensitive structure to proline imide bond isomerization.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The binding of cyanide to both oxidized and ascorbate-reduced forms of Pseudomonas cytochrome c-551 oxidase was investigated. Spectral studies on the oxidized enzyme and its apoprotein showed that the ligand can bind to both the c and d, haem components of the molecule, and kinetic observations indicated that both chromophores reacted, under a variety of conditions, with very similar rates. Cyanide combination velocities were dependent on ligand concentration, and increasing the pH also accelerated the reaction; the second-order rate constant was estimated as approx. 0.2M-1 . s-1 at pH 7.0. The binding of cyanide to the protein was observed to have a considerable influence on reduction of the enzyme by ascorbate. Spectral and kinetic observations have revealed that the species haem d13+-cyanide and any unbound haem c may react relatively rapidly with the reductant, but the behaviour of cyanide-bound haem c indicates that it may not be reduced without prior dissociation of the ligand, which occurs relatively slowly. The reaction of reduced Pseudomonas cytochrome oxidase with cyanide is radically different from that of the oxidized protein. In this case the ligand only binds to the haem d1 component and reacts much more rapidly. Stopped-flow kinetic measurements showed the binding to be biphasic in form. Both the rates of these processes were dependent on cyanide concentration, with the fast phase having a second-order rate constant of 9.3 X 10(5) M-1 . s-1 and the slow phase one of 2.3 X 10(5) M-1 . s-1. The relative proportions of the two phases also showed a dependency on cyanide concentration, the slower phase increasing as the cyanide concentration decreased. Computer simulations indicate that a reaction scheme originally proposed for the reaction of the enzyme with CO is capable of providing a reasonable explanation of the experimental results. Static-titration data of the reduced enzyme with with cyanide indicated that the binding was non-stoicheiometric, the ligand-binding curve being sigmoidal in shape. A Hill plot of the results yielded a Hill coefficient of 2.6.  相似文献   

11.
A kinetic analysis of the ATP-dependent dissociation of wild-type GroEL and mutants from immobilized GroES was carried out using surface plasmon resonance. Excellent fits of the data were obtained using a double-exponential equation with a linear drift. Both the fast and slow observed dissociation rate constants are found to have a sigmoidal dependence on the concentration of ATP. The values of the Hill coefficients corresponding to the fast and slow observed rate constants of dissociation of wild-type GroEL and the Arg197-->Ala mutant are in good agreement with the respective values of the Hill coefficients previously determined for these proteins from plots of initial rates of ATP hydrolysis as a function of ATP concentration, in the presence of GroES. Our results are consistent with a kinetic mechanism for dissociation of the GroEL-GroES complex according to which GroES release takes place after an ATP-induced conformational change in the trans ring that is preceded by ATP hydrolysis and a subsequent conformational change in the cis ring. It is shown that the rate of complex dissociation increases with increasing positive cooperativity in ATP binding by the GroEL ring distal to GroES in the GroEL-GroES complex.  相似文献   

12.
Cytochrome c-554 of the ammonia-oxidizing chemolithoautotropic bacteria is thought to mediate electron transfer from hydroxylamine oxidoreductase to a terminal oxidase and/or to ammonia monooxygenase. The cytochrome has four c hemes which interact magnetically and have the same redox potential. We report that the kinetics of reduction of ferric cytochrome c-554 by dithionite or the oxidation of ferrous cytochrome c-554 by O2 or H2O2 are complex and multiphasic. Transient rapid-scan difference spectra indicate discrete maxima at approximately 418 nm, 425 nm and 432 nm. Absorbance changes at all three difference maxima appear to occur in all kinetic phases, although not in equal amounts for each wavelength. Reduction by 20 mM dithionite was biphasic. At pH 7.5 the first phase, which involved approximately 50% of the total absorbance change, had a rate constant (20 degrees C) of 140 s-1 and energy of activation of 20 kJ X mol-1. The slow phase had a rate constant 0.43 s-1 and a relatively high energy of activation, 87 kJ X mol-1, suggesting that a change in protein configuration accompanied the reaction. As the pH of the solution increased, the rate constant for both phases decreased and the fraction of absorbance change in the rapid phase increased. Oxidation of ferrous cytochrome c-554 by O2 involved a discrete rapid phase with a rate constant of 14 s-1, accounting for 6% of the absorbance. The remainder of the reaction was multiphasic with rate constants in the range 0.1-0.01 s-1. With H2O2 as the oxidant, the rapid phase involved 39% of the change in absorbance with a rate constant of 19 s-1. The remainder of the reoxidation was multiphasic with rate constants ranging over 0.4-0.01 s-1.  相似文献   

13.
The reaction of trimethylamine dehydrogenase with trimethylamine   总被引:1,自引:0,他引:1  
The reductive half-reaction of trimethylamine dehydrogenase with its physiological substrate trimethylamine has been examined by stopped-flow spectroscopy over the pH range 6.0-11.0, with attention focusing on the fastest of the three kinetic phases of the reaction, the flavin reduction/substrate oxidation process. As in previous work with the slow substrate diethylmethylamine, the reaction is found to consist of three well resolved kinetic phases. The observed rate constant for the fast phase exhibits hyperbolic dependence on the substrate concentration with an extrapolated limiting rate constant (klim) greater than 1000 s-1 at pH above 8.5, 10 degrees C. The kinetic parameter klim/Kd for the fast phase exhibits a bell-shaped pH dependence, with two pKa values of 9.3 +/- 0.1 and 10. 0 +/- 0.1 attributed to a basic residue in the enzyme active site and the ionization of the free substrate, respectively. The sigmoidal pH profile for klim gives a single pKa value of 7.1 +/- 0. 2. The observed rate constants for both the intermediate and slow phases are found to decrease as the substrate concentration is increased. The steady-state kinetic behavior of trimethylamine dehydrogenase with trimethylamine has also been examined, and is found to be adequately described without invoking a second, inhibitory substrate-binding site. The present results demonstrate that: (a) substrate must be protonated in order to bind to the enzyme; (b) an ionization group on the enzyme is involved in substrate binding; (c) an active site general base is involved, but not strictly required, in the oxidation of substrate; (d) the fast phase of the reaction with native enzyme is considerably faster than observed with enzyme isolated from Methylophilus methylotrophus that has been grown up on dimethylamine; and (e) a discrete inhibitory substrate-binding site is not required to account for excess substrate inhibition, the kinetic behavior of trimethylamine dehydrogenase can be readily explained in the context of the known properties of the enzyme.  相似文献   

14.
Kinetics and mechanism of bilirubin binding to human serum albumin   总被引:3,自引:0,他引:3  
The kinetics of bilirubin binding to human serum albumin at pH 7.40, 4 degrees C, was studied by monitoring changes in bilirubin absorbance. The time course of the absorbance change at 380 nm was complex: at least three kinetic events were detected including the bimolecular association (k1 = 3.8 +/- 2.0 X 10(7) M-1 S-1) and two relaxation steps (52 = 40.2 +/- 9.4 s-1 and k3 = 3.8 +/- 0.5 s-1). The presence of the two slow relaxations was confirmed under pseudo-first order conditions with excess albumin. Curve-fitting procedures allowed the assignment of absorption coefficients to the intermediate species. When the bilirubin-albumin binding kinetics was observed at 420 nm, only the two relaxations were seen; apparently the second order association step was isosbestic at this wavelength. The rate of albumin-bound bilirubin dissociation was measured by mixing the pre-equilibrated human albumin-bilirubin complex with bovine albumin. The rate constant for bilirubin dissociation measured at 485 nm was k-3 = 0.01 s-1 at 4 degrees C. A minimum value of the equilibrium constant for bilirubin binding to human albumin determined from the ratio k1/k-3 is therefore approximately 4 X 10(9) M-1.  相似文献   

15.
The carotenoid content of thylakoids and envelopes isolated from dark-or light-treated spinach (Spinacia oleracea L.) chloroplasts was compared. In thylakoids, light induced a decrease of violaxanthin parallel with a stoichiometric increase of zeaxanthin due to violaxanthin deepoxidation. In envelopes, violaxanthin was also decreased and the relative decrease was similar to thylakoids, but zeaxanthin increase was small resulting in an over-all decrease of the amount of envelope carotenoids. When violaxanthin deepoxidation in thylakoids was partly inhibited by 10 nm nigericin, violaxanthin decrease in the envelope was inhibited to a similar degree.  相似文献   

16.
The chaperonin CCT (chaperonin containing t-complex polypeptide 1 (TCP-1)) from bovine testis was mixed rapidly with different concentrations of ATP and the time-resolved change in fluorescence emission, upon excitation at 280 nm, was followed. Two kinetic phases were observed and assigned by (i) analyzing the dependence of the corresponding observed rate constants on ATP concentration; and (ii) by carrying out mixing experiments also with ADP, ATPgammaS and ATP without K(+). The values of the observed rate constants corresponding to both phases are found to be dependent on ATP concentration. The observed rate constant corresponding to the fast phase displays a bi-sigmoidal dependence on ATP concentration with Hill coefficients that are similar to those determined in steady-state ATPase experiments. This phase most likely reflects ATP binding-induced conformational changes. The rate constant of the conformational change in the presence of excess ATP is about 17s(-1) (at 25 degrees C) and is tenfold slower than the corresponding rate constant of GroEL. The observed rate constant corresponding to the second slower phase displays a hyperbolic dependence on ATP concentration. This phase is not observed in mixing experiments of CCT with ADP, ATPgammaS or ATP without K(+) and it, therefore, reflects a conformational change associated with ATP hydrolysis. Taken together, our results indicate that the kinetic mechanism of the allosteric transitions of CCT differs considerably from that of GroEL.  相似文献   

17.
The decrease of the intrinsic tryptophan fluorescence intensity of purified influenza (X31 strain) hemagglutinin (HA) was used to monitor the low pH-induced conformational change of this protein. The kinetics of the fluorescence decrease depended strongly on the pH. At pH optimal for fusion, the change in tryptophan fluorescence was fast and could be fitted to a monoexponential function. We measured a rate constant of 5.78 s-1 (t1/2 = 120 ms) at pH 4.9 using rapid stopped-flow mixing. Under suboptimal conditions (higher pH), the rate constant was decreased by an order of magnitude. In addition, a slow component appeared and the fluorescence decrease followed a sum of two exponentials. The kinetics of conformational changes were compared with those of the fusion of influenza virus with red blood cell membranes as assessed by the R18-dequenching assay. At optimal pH the HA conformational change was not rate-limiting for the fusion process. However, at sub-optimal pH, the slow transition to the fusogenic conformational of HA resulted in slower kinetics and decreased extent of fusion.  相似文献   

18.
The complex oligomeric assembly of the hemoglobin subunits may influence the autoxidation rate. To understand this relation, the rate of autoxidation was studied at pH 9.0, where the Glossoscolex paulistus Hemoglobin (GpHb) dissociates. At alkaline pH, this hemoglobin is dissociated into monomers, trimers and tetramers, allowing the study of the integral protein and monomer subunit autoxidation on independent experiments. The autoxidation rate was evaluated in the presence and absence of cyanide (CN(-)), a strong field ligand to the ferric ion. The oxidation kinetic was monitored using the UV-vis absorption at 415 nm, and resulted in: i) bi-exponential kinetics for the whole hemoglobin (indicating a fast and a slow oxidative process) and ii) mono-exponential for the monomer (indicating a single process). To understand the specific characteristics of each autoxidation process, Arrhenius plots allowed the determination of the activation energy. The experimental results indicate for the whole hemoglobin in the absence of CN(-) an activation energy of 150 +/- 10 kJ mol(-1) for the fast and the slow processes. Under the same conditions the monomer displayed an activation energy of 160 +/- 10 kJ mol(-1), very close to the value obtained for the integral protein. The pseudo-second order rate constant for the whole protein autoxidation by CN(-) showed two different behaviors characterized by a rate constant k(CN1)' = 0.11 +/- 0.02 s(-1) mol(-1) L for CN(-) concentrations lower than 0.012 mol L(-1); and k(CN1)" = 0.76 +/- 0.04 s(-1) mol(-1) L at higher concentrations for the fast process, while the slow process remain constant with k(CN2) = 0.033 +/- 0.002 s(-1) mol(-1) L. The monomer has a characteristic rate constant of 0.041 +/- 0.002 s(-1) mol(-1) L for all cyanide concentrations. Comparing the results for the slow process of the whole hemoglobin and the oxidation of the monomer, it is possible to infer that the slow process has a strong contribution of the monomer in the whole hemoglobin kinetic. Moreover, as disulfide linkers sustain the trimer assembly, cooperativity may explain the higher kinetic constant for this subunit.  相似文献   

19.
Light-induced absorbance change at 515 nm in spinach chloroplastswas studied in the temperature range from –2?C to 27?C.Lowering of temperature had no marked effect on the extentsof initial "light-on" spike and the steady-state change overthe temperature range examined, whereas the rate of recoveryof the 515-nm change was significantly reduced at lower temperatures.Above 15?C, recovery of the 515-nm change after continuous illuminationshowed a first-order kinetics. In contrast, the recovery wascomposed of a fast and a slow phases at lower temperatures. The fast phase of the recovery of the 515-nm change was acceleratedby carbonyl cyanide m-chlorophenylhydrazone, valinomycin plusK+ or sodium tetraphenylboron, while the slow phase was completelyeliminated in glutaraldehyde-fixed chloroplasts. Light-inducedchange in absorbance at 546 nm, an indicator of structural changesof membrane, showed almost the same dependency on temperatureas the slow phase of the recovery of the 515-nm change. Theseresults suggest that not only electric field formation acrossthe thylakoid membrane but also structural or conformationalchanges in the membrane participate in the 515-nm absorbancechange observed under steady illumination. (Received July 5, 1976; )  相似文献   

20.
R L Blakley  L Cocco 《Biochemistry》1985,24(17):4704-4709
Stopped-flow spectrophotometry and stopped-flow fluorometry have been used to study the binding of methotrexate (MTX) and 3-deazamethotrexate (3-deazaMTX) to dihydrofolate reductase (DHFR) isoenzymes from Streptococcus faecium and from Lactobacillus casei. The absorbance change and fluorescence quenching that occur when MTX binds to DHFR isoenzyme II from S. faecium (SFDHFR II) are both biphasic and give similar apparent rate constants for both phases. The faster phase has an apparent rate constant that is dependent on MTX concentration and therefore corresponds to the initial binding reaction. From the concentration dependence it has been calculated that the association rate constant is 3.0 X 10(5) M-1 s-1 at 20 degrees C and pH 7.3, and the association constant (equilibrium constant) under these conditions is 5.8 X 10(5) M-1. By examination of the amplitude of the fast-phase absorbance change at various wavelengths, it has been determined that the absorbance change occurring in the fast phase is due to MTX protonation. Within the limits of the method it was thus not possible to detect a difference in the rates of binding and of protonation of MTX. The MTX association rate constant is pH dependent, decreasing 330-fold as the pH is decreased from 5.0 to 9.0. The data fit well to a curve generated by assuming a single ionization with a pKa of 6.0 and a pH-independent association rate constant 1000-fold greater for binding of protonated MTX to SFDHFR II than for binding of unprotonated MTX.(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