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.
Interactions of glutaraldehyde with either n-butylamine, poly(α,L -lysine), or collagen resulted in a fast release of protons in dilute aqueous solutions at various pH values, followed by much slower changes. The latter reactions, which extended over hours and days, were followed spectrophotometrically and revealed the formation of distinct absorption bands in the visible and near-ultraviolet regions in all the above systems. The visible-range bands disappeared upon treatment with sodium borohydride. A qualitative relationship between oxygen uptake by the system n-butylamine–glutaraldehyde and the slow formation of colored products has been established, while the chemical nature of the reaction products has not been determined. Sedimentation velocity, viscosity, and optical rotation measurements on the products of interaction between poly(L -lysine) and glutaraldehyde in aqueous solution indicated large conformational changes in the polyamino acid present in excess (in residues) over the dialdehyde. In particular, the intrinsic viscosity dropped considerably after interaction, indicating intramolecular crosslinking. At molar ratios of 1:1 between polylsine residues and aldehyde groups, intermolecular crosslinking of polylysine was obtained at pH 8.6. Electron microscopic examinations of collagen samples treated by glutaraldehyde at various pH values indicated changes from unordered to more ordered structures upon treatment with glutaraldehyde, in particular at pH 10. The present structural and optical investigations are considered to be relevant to tanning processes of hides and to fixation procedures.  相似文献   

2.
Ming M  Lu M  Balashov SP  Ebrey TG  Li Q  Ding J 《Biophysical journal》2006,90(9):3322-3332
The pH-dependence of photocycle of archaerhodopsin 4 (AR4) was examined, and the underlying proton pumping mechanism investigated. AR4 is a retinal-containing membrane protein isolated from a strain of halobacteria from a Tibetan salt lake. It acts as a light-driven proton pump like bacteriorhodopsin (BR). However, AR4 exhibits an "abnormal" feature--the time sequence of proton release and uptake is reversed at neutral pH. We show here that the temporal sequence of AR4 reversed to "normal"--proton release preceding proton uptake--when the pH is increased above 8.6. We estimated the pK(a) of the proton release complex (PRC) in the M-intermediate to be approximately 8.4, much higher than 5.7 of wide-type BR. The pH-dependence of the rate constant of M-formation shows that the pK(a) of PRC in the initial state of AR4 is approximately 10.4, whereas it is 9.7 in BR. Thus in AR4, the chromophore photoisomerization and subsequent proton transport from the Schiff base to Asp-85 is coupled to a decrease in the pK(a) of PRC from 10.4 to 8.4, which is 2 pK units less than in BR (4 units). This weakened coupling accounts for the lack of early proton release at neutral pH and the reversed time sequence of proton release and uptake in AR4. Nevertheless the PRC in AR4 effectively facilitates deprotonation of primary proton acceptor and recovery of initial state at neutral pH. We found also that all pK(a)s of the key amino acid residues in AR4 were elevated compared to those of BR.  相似文献   

3.
The pH dependencies of the rate constants in the photocycles of recombinant D96N and D115N/D96N bacteriorhodopsins were determined from time-resolved difference spectra between 70 ns and 420 ms after photoexcitation. The results were consistent with the model suggested earlier for proteins containing D96N substitution: BR hv----K----L----M1----M2----BR. Only the M2----M1 back-reaction was pH-dependent: its rate increased with increasing [H+] between pH 5 and 8. We conclude from quantitative analysis of this pH dependency that its reverse, the M1----M2 reaction, is linked to the release of a proton from a group with a pKa = 5.8. This suggests a model for wild-type bacteriorhodopsin in which at pH greater than 5.8 the transported proton is released on the extracellular side from this as yet unknown group and on the 100-microseconds time scale, but at pH less than 5.8, the proton release occurs from another residue and later in the photocycle most likely directly from D85 during the O----BR reaction. We postulate, on the other hand, that proton uptake on the cytoplasmic side will be by D96 and during the N----O reaction regardless of pH. The proton kinetics as measured with indicator dyes confirmed the unique prediction of this model: at pH greater than 6, proton release preceded proton uptake, but at pH less than 6, the release was delayed until after the uptake. The results indicated further that the overall M1----M2 reaction includes a second kinetic step in addition to proton release; this is probably the earlier postulated extracellular-to-cytoplasmic reorientation switch in the proton pump.  相似文献   

4.
Horseradish peroxidase was reacted with glutaraldehyde under various reaction conditions. The reaction product was, in a second step, bound covalently to aminohexyl groups attached to Sepharose particles. The influence of pH, time and the concentration ratio of enzyme:glutaraldehyde on the reaction was evaluated. A first step reaction with 100-fold molar excess of glutaraldehyde to horseradish peroxidase at pH 9.5 for 2 hr at room temperature results in a high yield of conjugated enzyme with well preserved enzymatic activity.  相似文献   

5.
Cytochrome c nitrite reductase is a homodimeric enzyme, containing five covalently attached c-type hemes per subunit. Four of the heme irons are bishistidine-ligated, whereas the fifth, the active site of the protein, has an unusual lysine coordination and calcium site nearby. A fascinating feature of this enzyme is that the full six-electron reduction of the nitrite is achieved without release of any detectable reaction intermediate. Moreover, the enzyme is known to work over a wide pH range. Both findings suggest a unique flexibility of the active site in the complicated six-electron, seven-proton reduction process. In the present work, we employed density functional theory to study the energetics and kinetics of the initial stages of nitrite reduction. The possible role of second-sphere active-site amino acids as proton donors was investigated by taking different possible protonation states and geometrical conformations into account. It was found that the most probable proton donor is His277, whose spatial orientation and fine-tuned acidity lead to energetically feasible, low-barrier protonation reactions. However, substrate protonation may also be accomplished by Arg114. The calculated barriers for this pathway are only slightly higher than the experimentally determined value of 15.2 kcal/mol for the rate-limiting step. Hence, having proton-donating side chains of different acidity within the active site may increase the operational pH range of the enzyme. Interestingly, Tyr218, which was proposed to play an important role in the overall mechanism, appears not to take part in the reaction during the initial stage.  相似文献   

6.
Hajime Noguchi 《Biopolymers》1966,4(10):1105-1113
Water-insoluble films of poly-L -lysine, crosslinked with formaldehyde, were suspended in aqueous media and their relative lengths measured as a function of pH. A sharp transition of the polymer was observed in the pH range which corresponded with that observed in polylysine solutions by optical rotation or dilatometry. In NaBr and NaCl solutions the coiled form of the polylysine film shrinks with increasing salt concentration, but in NaHCO3 solution the extent of the contraction is larger, and the coil–helix transition of polylysine occurs at lower pH when NaHCO3 is added to the medium. If one assumes the formation of amino carbamate in this case, this phenomenon can be well explained. Urea does break up the hydrogen bonds in helical polylysine film, but not completely. This result is interesting compared with that obtained for poly(L -glutamic acid). After the coil–helix transition region was found by film experiments, the volume change associated with the coil-to-helix transition was measured and found to be about 1–l.5 ml. per amino residue after taking electrostatic interaction into consideration. This value is nearly same as that obtained for poly(L -glutamic acid). By contrast, the value for poly-γ-benzyl-L -glutamate was reported to be ?0.077 ml./mole of repeating unit. So it is still necessary to determine the magnitude and direction of the volume change for various kinds of polypeptides.  相似文献   

7.
P Jahns  W Junge 《Biochemistry》1992,31(32):7398-7403
In photosynthesis of green plants water is oxidized to dioxygen. This four-step process is accompanied by the release of four protons (per molecule of dioxygen) into the lumen of thylakoids. In dark-adapted thylakoids which are excited with a series of short flashes of light, the extent of proton release oscillates with period four as a function of flash number. Noninteger and pH-dependent proton/electron ratios (e.g., 1.1, 0.25, 1.0, and 1.65 at pH 7) have been attributed to a superposition of two reactions: chemical production of protons and transient electrostatic response of peripheral amino acid side chains. Aiming at the true pattern of proton production, we investigated the relative contribution of peripheral proteins. Thylakoids with and without chlorophyll a/b binding proteins were compared. Thylakoids lacking chlorophyll a/b binding proteins were prepared from pea seedlings grown under intermittent light [Jahns, P., & Junge, W. (1992) Biochemistry (preceding paper in this issue)]. We found no oscillation of proton release in the pH range from 6 to 7.5. These and other results showed that chlorophyll a/b binding proteins, which primarily serve as light-harvesting antennas, modulate proton release by water oxidation. A nonoscillating pattern of proton release, with proton/electron ratios of 1:1:1:1 more closely represents the events in the catalytic center proper. This implies hydrogen abstraction rather than electron abstraction from water during the oxygen-evolving step S3----S0.  相似文献   

8.
ω-Amino acid: pyruvate aminotransferase of Pseudomonas sp. F-126 catalyzes stoichiometricalJy a transamination between various amines and pyruvate. Most of alkyl and aromatic monoamines served as an amino donor. The enzyme activity was affected by carbon number of straight-chain alkylmonoamines with a maximum activity at 5-carbon unit, n-amylamine. Michaelis constants for n-butylamine and pyruvate were calculated to be 66.6 mm and 5.5 mm respectively. The enzyme was active in the alkaline range with a maximum at pH 10.5 ~ 11.0, though not any activity was observed at the pH below 8.0. The optimum temperature for the reaction was at 60°C.  相似文献   

9.
Decarboxylation of pyruvate by pyruvate decarboxylase (EC 4.1.1.1) was performed in a reaction mixture containing 50% deuterium. The isolated product, acetaldehyde, was investigated directly by 1H NMR and by mass spectrometry after conversion to the 2,4-dinitrophenyl hydrazone. The protium content of 56% at acetaldehyde C1 demonstrates a specific protonation of the corresponding intermediate by the enzyme. Proton inventory studies and enzyme modification indicate the 4' amino group of the coenzyme, thiamine pyrophosphate, in an immonium structure being a possible proton donor. A 'partially concerted' mechanism is suggested for the reaction steps following the decarboxylation.  相似文献   

10.
When the extracellular pH was increased from 7.6 to 9.8, Clostridium paradoxum, a novel alkalithermophile, increased its pH gradient across the cell membrane ((Delta)pH, pH(infin) - pH(infout)) by as much as 1.3 U. At higher pH values (>10.0), the (Delta)pH and membrane potential ((Delta)(psi)) eventually declined, and the intracellular pH increased significantly. Growth ceased when the extracellular pH was greater than 10.2 and the intracellular pH increased to above 9.8. The membrane potential increased to 110 (plusmn) 8.6 mV at pH 9.1, but the total proton motive force ((Delta)p) declined from about 65 mV at pH 7.6 to 25 mV at pH 9.8. Between the extracellular pH of 8.0 and 10.3, the intracellular ATP concentration was around 1 mM and decreased at lower and higher pH values concomitantly with a decrease in growth rate.  相似文献   

11.
The activity of yeast enolase is inhibited by Li+ and Na+. At pH 7.1, inhibition by Li+ is "mixed" with respect to Mg2+; both Vmax and Km (Mg2+) are increased by Li+. The inhibition by Li+ appears to be partial, indicating that enzyme with Li+ bound is active. The step inhibited by Li+ cannot be proton abstraction since Li+ decreases the kinetic isotope effect on Vmax. At pH 9.2, where proton abstraction is no longer partially rate-limiting, inhibiton by Li+ is competitive with respect to Mg2+. The rate of enzyme-catalyzed exchange of the C-2 hydrogen with solvent is not affected by Li+. We interpret these results as follows: Li+ (and Na+) binds to enolase and decreases the rate of at least one step in the mechanism. At pH 7.1, this step is partially rate-limiting; at pH 9.2, this step is a fast step in the reaction. The step inhibited by Li+ cannot be proton abstraction but may be release of product (phosphoenol pyruvate) or Mg2+.  相似文献   

12.
Formation of a product easily converted to methylglyoxal on TLC with silica gel was observed in an early stage of the reaction mixture of sugar with an alkylamine or amino acid. NMR spectra of the ether extract of reaction mixtures indicated that methylglyoxal dialkylimme was produced mainly at an early stage of the reaction of glucose with alkylamine, and was assumed to change to methylglyoxal on the TLC. The C3 imine production in the t-butylamine system was apparently little and slow compared to that in the normal alkylamine system. A large, rapid production of C3 imine was also observed in the system of the Amadori product and n-butylamine. These results suggested that the C3 formation in the system with normal alkylamine may occur mainly via a newly proposed mechanism, though the ¿-butylamine system may possibly produce it according to the scheme proposed by Hodge.  相似文献   

13.
F Rappaport  J Lavergne 《Biochemistry》1991,30(41):10004-10012
Flash-induced absorption changes of pH-indicating dyes were investigated in photosystem II enriched membrane fragments, in order to retrieve the individual contributions to proton release of the successive transitions of the Kok cycle. These stoichiometric coefficients were found to be, in general, noninteger and to vary as a function of pH. Proton release on the S0----S1 step decreases from 1.75 at pH 5.5 to 1 at pH 8, while, on S1----S2 the stoichiometry increases from 0 to 0.5 in the same pH range and remains close to 1 for S2----S3. These findings are analyzed in terms of pK shifts of neighboring amino acid residues caused by electrostatic interactions with the redox centers involved in the two first transitions. The electrochromic shift of a chlorophyll, associated with the S transitions, responding to local electrostatic effects was investigated under similar conditions. The pH dependence of this signal upon the successive transitions was found correlated with the titration of the proton release stoichiometries, expressing the electrostatic balance between the oxidation and deprotonation processes.  相似文献   

14.
Martin SF  Hergenrother PJ 《Biochemistry》1999,38(14):4403-4408
The phosphatidylcholine-preferring phospholipase C from Bacillus cereus (PLCBc) is a 28.5 kDa enzyme with three zinc ions in its active site. Although much is known about the roles that various PLCBc active site amino acids play in binding and catalysis, there is little information about the rate-determining step of the PLCBc-catalyzed hydrolysis of phospholipids and the catalytic cycle of the enzyme. To gain insight into these aspects of the hydrolysis, solvent viscosity variation experiments were conducted to determine whether an external step (substrate binding or product release) or an internal step (hydrolysis) is rate-limiting. The data indicate that the PLCBc-catalyzed reaction is unaffected by changes in solvent viscosity. This observation is inconsistent with the notion of substrate binding or product release being rate-determining and supports the hypothesis that a chemical step is rate-limiting. Furthermore, a deuterium isotope effect of 1.9 and a linear proton inventory plot indicate one proton is transferred in the rate-determining step. These data may be used to formulate a comprehensive catalytic cycle that is for the first time based on experimental evidence. In this mechanism, Asp55 of PLCBc activates an active site water molecule for attack on the phosphodiester bond, the hydrolysis of which is rate-limiting. The phosphorylcholine product is the first to leave the active site, followed by diacylglycerol.  相似文献   

15.
We have used the stopped-flow indicator dye method to measure proton release and product formation simultaneously in the initial transient-state portion of the glutamate dehydrogenase-catalyzed oxidative deamination of L-glutamate. We observe a measurably slow release of a proton from the enzyme-NADP-L-glutamate complex. This proton release precedes the hydride transfer step, as indicated by the distinct lag in the product formation signal. We show that the proton release step corresponds to an obligatory intermediate in the reaction sequence. We also find that compounds which are competitive inhibitors of L-glutamate are capable of inducing this phenomenon. We prove that this unanticipated prehydride transfer event cannot be due to the release of an alpha-amino group proton from the substrate.  相似文献   

16.
The kinetics of the enzymatic step of the peroxidatic reaction between NAD and hydrogen peroxide, catalysed by horse liver alcohol dehydrogenase (alcohol:NAD+ oxidoreductase, EC 1.1.1.1), has been investigated at pH 7 at high enzyme concentration. Under such conditions no burst phase has been observed, thus indicating that the rate-limiting step in the process, which converts NAD into Compound I, either precedes or coincides with the chemical step responsible for the observed spectroscopic change. Kinetic analysis of the data, performed according to a simplified reaction scheme suggests that the rate-limiting step is coincident with the spectroscopic (i.e., chemical) step itself. Furthermore, the absence of a proton burst phase indicates the proton release step does not precede the chemical step, in contrast with the case of ethanol oxidation. A kinetic effect of different premixing conditions on the reaction rate has been observed and attributed to the presence of NADH formed in the 'blank reaction' between NAD and residual ethanol tightly bound to alcohol dehydrogenase. A molecular mechanism for the enzymatic peroxidation step is finally proposed, exploiting the knowledge of the much better known reaction of ethanol oxidation. Inhibition of this reaction by NADH has been investigated with respect to H2O2 (noncompetitive, Ki about 10 microM) and to NAD (competitive, Ki about 0.7 microM). The effect of temperature on the steady-state reaction state (about 65 kJ/mol activation energy) has also been studied.  相似文献   

17.
A proton electrochemical potential across the membranes of photosynthetic purple bacteria is established by a light-driven proton pump mechanism: the absorbed light in the reaction center initiates electron transfer which is coupled to the vectorial displacement of protons from the cytoplasm to the periplasm. The stoichiometry and kinetics of proton binding and release can be tracked directly by electric (glass electrodes), spectrophotometric (pH indicator dyes) and conductimetric techniques. The primary step in the formation of the transmembrane chemiosmotic potential is the uptake of two protons by the doubly reduced secondary quinone in the reaction center and the subsequent exchange of hydroquinol for quinone from the membrane quinone-pool. However, the proton binding associated with singly reduced promary and/or secondary quinones of the reaction center is substoichiometric, pH-dependent and its rate is electrostatically enhanced but not diffusion limited. Molecular details of protonation are discussed based on the crystallographic structure of the reaction center of purple bacteriaRb. sphaeroides andRps. viridis, structure-based molecular (electrostatic) calculations and mutagenesis directed at protonatable amino acids supposed to be involved in proton conduction pathways.  相似文献   

18.
The variation with pH of the kinetic parameters for the alcohol and acetaldehyde reactions were studied for the alleloenzyme AdhS from Drosophila melanogaster. The variation of Ki (KEO,I) with pH for two ethanol-competitive inhibitors, pyrazole and 2,2,2-trifluoroethanol, was also studied. Both alcohol oxidation and acetaldehyde reduction follow a compulsory ordered pathway, with coenzyme binding first. The rate-limiting step for ethanol oxidation is complex and involves at least hydride transfer and dissociation of the enzyme-NADH complex (ER). In contrast with this, the rate-limiting step for the back reaction, i.e. the reduction of acetaldehyde, is dissociation of the enzyme-NAD+ complex (EO). A rate-limiting ER dissociation appears in the oxidation of the secondary alcohol propan-2-ol, whereas for the back reaction, i.e. acetone reduction, hydride transfer in the ternary complexes is rate-limiting. There is one group in the free enzyme, with a pK of approx. 8.0, that regulates the kon velocity for NADH, whereas for NAD+ several groups seem to be involved. A group in the enzyme is drastically perturbed by the formation of the binary EO complex. Protonation of this group with a pK of 7.6 in the EO complex resulted in weakened alcohol and inhibitor binding, in addition to an increased dissociation rate of NAD+ from the binary EO complex. Neither the binding of acetaldehyde nor the dissociation rate of NADH from the binary ER complex varied within the pH region studied.  相似文献   

19.
The D and L copolymerizations of γ-benzyl glutamate N-carboxylic anhydride (NCA) were carried out by two different initiators, n-butylamine and sodium methoxide. The stereoregularity of the polymer was examined by infrared spectroscopy in the region 700–200 cm?1. A remarkable difference was found between the polymers obtained by these initiators in the region 450–400 cm?1. In the polymer obtained with sodium methoxide as initiator, the 409-cm?1 peak assigned to local α-helical conformation was not greatly affected by the amount of the L -form, but in the polymer obtained by n-butylamine this peak was much affected by the variation of L -content. This indicates that the stereo-selectivity of the polymerization in the sodium methoxide initiated system is higher than that in the n-butylamine initiated system.  相似文献   

20.
Emissions of acetaldehyde from tree leaves were investigated by proton‐transfer‐reaction mass spectrometry (PTR‐MS), a technique that allows simultaneous monitoring of different leaf volatiles, and confirmed by derivatization and high‐performance liquid chromatography analysis. Bursts of acetaldehyde were released by sycamore, aspen, cottonwood and maple leaves following light–dark transitions; isoprene emission served as a measure of chloroplastic processes. Acetaldehyde bursts were not accompanied by ethanol, but exposure of leaves to inhibitors of pyruvate transport or respiration, or anoxia, led to much larger releases of acetaldehyde, accompanied by ethanol under anoxic conditions. These same leaves have an oxidative pathway for ethanol present in the transpiration stream, resulting in acetaldehyde emissions that are inhibited in vivo by 4‐methylpyrazole, an alcohol dehydrogenase (Adh) inhibitor. Labelling of leaf volatiles with 13CO2 suggested that the pools of cytosolic pyruvate, the proposed precursor of acetaldehyde bursts, were derived from both recent photosynthesis and cytosolic carbon sources. We hypothesize that releases of acetaldehyde during light–dark transitions result from a pyruvate overflow mechanism controlled by cytosolic pyruvate levels and pyruvate decarboxylase activity. These results suggest that leaves of woody plants contribute reactive acetaldehyde to the atmosphere under different conditions: (1) metabolic states that promote the accumulation of cytosolic pyruvate, triggering the pyruvate decarboxylase reaction; and (2) leaf ethanol oxidation resulting from ethanol transported from anoxic tissues.  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