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.
The protein motive force of metabolizing Bacillus subtilis cells was only slightly affected by changes in the external pH between 5 and 8, although the electrical component and the chemical component of the proton motive force contributed differently at different external pH. The electrical component of the proton motive force was very small at pH 5, and the chemical component was almost negligible at pH 7.5. At external pH values between 6 and 7.7, swimming speed of the cells stayed constant. Thus, either the electrical component or the chemical component of the proton motive force could drive the flagellar motor. When the proton motive force of valinomycin-treated cells was quantitatively decreased by increasing the external K+ concentration, the swimming speed of the cells changed in a unique way: the swimming speed was not affected until about--100 mV, then decreased linearly with further decrease in the proton motive force, and was almost zero at about--30 mV. The rotation rate of a flagellum, measured by a tethered cell, showed essentially the same characteristics. Thus, there are a threshold proton motive force and a saturating proton motive force for the rotation of the B. subtilis flagellar motor.  相似文献   

2.
Cytoplasmic pH and periplasmic pH of Escherichia coli cells in suspension were observed with 4-s time resolution using fluorimetry of TorA-green fluorescent protein mutant 3* (TorA-GFPmut3*) and TetR-yellow fluorescent protein. Fluorescence intensity was correlated with pH using cell suspensions containing 20 mM benzoate, which equalizes the cytoplasmic pH with the external pH. When the external pH was lowered from pH 7.5 to 5.5, the cytoplasmic pH fell within 10 to 20 s to pH 5.6 to 6.5. Rapid recovery occurred until about 30 s after HCl addition and was followed by slower recovery over the next 5 min. As a control, KCl addition had no effect on fluorescence. In the presence of 5 to 10 mM acetate or benzoate, recovery from external acidification was diminished. Addition of benzoate at pH 7.0 resulted in cytoplasmic acidification with only slow recovery. Periplasmic pH was observed using TorA-GFPmut3* exported to the periplasm through the Tat system. The periplasmic location of the fusion protein was confirmed by the observation that osmotic shock greatly decreased the periplasmic fluorescence signal by loss of the protein but had no effect on the fluorescence of the cytoplasmic protein. Based on GFPmut3* fluorescence, the pH of the periplasm equaled the external pH under all conditions tested, including rapid acid shift. Benzoate addition had no effect on periplasmic pH. The cytoplasmic pH of E. coli was measured with 4-s time resolution using a method that can be applied to any strain construct, and the periplasmic pH was measured directly for the first time.  相似文献   

3.
Bacteria migrate away from an acid pH and from a number of chemicals, including organic acids such as acetate; the basis for detection of these environmental cues has not been demonstrated. Membrane-permeant weak acids caused prolonged tumbling when added to Salmonella sp. or Escherichia coli cells at pH 5.5. Tethered Salmonella cells went from a prestimulus behavior of 14% clockwise rotation to 80% clockwise rotation when 40 mM acetate was added and remained this way for more than 30 min. A low external pH in the absence of weak acid did not markedly affect steady-state tumbling frequency. Among the weak acids tested, the rank for acidity (salicylate greater than benzoate greater than acetate greater than 5,5-dimethyl-2,4-oxazolidinedione) was the same as the rank for the ability to collapse the transmembrane pH gradient and to cause tumbling. At pH 7.0, the tumbling responses caused by the weak acids were much briefer. Indole, a non-weak-acid repellent, did not cause prolonged tumbling at low pH. Two chemotaxis mutants (a Salmonella mutant defective in the chemotaxis methylesterase and an E. coli mutant defective in the methyl-accepting protein in MCP I) showed inverse responses of enhanced counterclockwise rotation in the first 1 min after acetate addition. The latter mutant had been found previously to be defective in the sensing of gradients of extracellular pH and (at neutral pH) of acetate. We conclude (i) that taxes away from acid pH and membrane-permeant weak acids are both mediated by a pH-sensitive component located either in the cytoplasm or on the cytoplasmic side of the membrane, rather than by an external receptor (as in the case of the attractants), and (ii) that both of these taxes involve components of the chemotaxis methylation system, at least in the early phase of the response.  相似文献   

4.
Escherichia coli and Salmonella typhimurium were grown in a supplemented minimal medium (SMM) at a pH of 7.0 or 5.0 or were shifted from pH 7.0 to 5.0. Two-dimensional gel electrophoretic analysis of proteins labeled with H2(35)SO4 for 20 min during the shift showed that in E. coli, 13 polypeptides were elevated 1.5- to 4-fold, whereas in S. typhimurium, 19 polypeptides were increased 2- to 14-fold over the pH 7.0 control. Upon long-term growth at pH 5.0, almost double the number of polypeptides were elevated twofold or more in S. typhimurium compared with E. coli. In E. coli, there was no apparent induction of heat shock proteins upon growth at pH 5.0 in SMM. However, growth of E. coli in a complex broth to pH 5.0, or subsequent growth of fresh E. coli cells in the filtrate from this culture, showed that a subset of five polypeptides is uniquely induced by low pH. Two of these polypeptides, D60.5, the inducible lysyl-tRNA synthetase, and C62.5, are known heat shock proteins. Measurements of the internal pH (pHi) and growth rates of both organisms were made during growth in SMM at pH 7.0, pH 5.0, and upon the pH shift. The data show that the pHi of E. coli decreases more severely than that of S. typhimurium at an external pH of 5.0; the growth rate of E. coli is about one-half that of S. typhimurium at this pH, whereas the two organisms have the same growth rate at pH 7.0. The two-dimensional gel, growth, and pHi experiments collectively suggest that, at least in SMM, S. typhimurium is more adaptive to low-pH stress than is E. coli.  相似文献   

5.
The ability of Escherichia coli and Bacillus subtilis to regulate their cytoplasmic pH is well studied in cell suspensions but is poorly understood in individual adherent cells and biofilms. We observed the cytoplasmic pH of individual cells using ratiometric pHluorin. A standard curve equating the fluorescence ratio with pH was obtained by perfusion at a range of external pH 5.0 to 9.0, with uncouplers that collapse the transmembrane pH difference. Adherent cells were acid stressed by switching the perfusion medium from pH 7.5 to pH 5.5. The E. coli cytoplasmic pH fell to a value that varied among individual cells (range of pH 6.2 to 6.8), but a majority of cells recovered (to pH 7.0 to 7.5) within 2 min. In an E. coli biofilm, cells shifted from pH 7.5 to pH 5.5 failed to recover cytoplasmic pH. Following a smaller shift (from pH 7.5 to pH 6.0), most biofilm cells recovered fully, although the pH decreased further than that of isolated adherent cells, and recovery took longer (7 min or longer). Some biofilm cells began to recover pH and then failed, a response not seen in isolated cells. B. subtilis cells were acid shifted from pH 7.5 to pH 6.0. In B. subtilis, unlike the case with E. coli, cytoplasmic pH showed no "overshoot" but fell to a level that was maintained. This level of cytoplasmic pH post-acid shift varied among individual B. subtilis cells (range of pH, 7.0 to 7.7). Overall, the cytoplasmic pHs of individual bacteria show important variation in the acid stress response, including novel responses in biofilms.  相似文献   

6.
Escherichia coli and Salmonella typhimurium were grown in a supplemented minimal medium (SMM) at a pH of 7.0 or 5.0 or were shifted from pH 7.0 to 5.0. Two-dimensional gel electrophoretic analysis of proteins labeled with H2(35)SO4 for 20 min during the shift showed that in E. coli, 13 polypeptides were elevated 1.5- to 4-fold, whereas in S. typhimurium, 19 polypeptides were increased 2- to 14-fold over the pH 7.0 control. Upon long-term growth at pH 5.0, almost double the number of polypeptides were elevated twofold or more in S. typhimurium compared with E. coli. In E. coli, there was no apparent induction of heat shock proteins upon growth at pH 5.0 in SMM. However, growth of E. coli in a complex broth to pH 5.0, or subsequent growth of fresh E. coli cells in the filtrate from this culture, showed that a subset of five polypeptides is uniquely induced by low pH. Two of these polypeptides, D60.5, the inducible lysyl-tRNA synthetase, and C62.5, are known heat shock proteins. Measurements of the internal pH (pHi) and growth rates of both organisms were made during growth in SMM at pH 7.0, pH 5.0, and upon the pH shift. The data show that the pHi of E. coli decreases more severely than that of S. typhimurium at an external pH of 5.0; the growth rate of E. coli is about one-half that of S. typhimurium at this pH, whereas the two organisms have the same growth rate at pH 7.0. The two-dimensional gel, growth, and pHi experiments collectively suggest that, at least in SMM, S. typhimurium is more adaptive to low-pH stress than is E. coli.  相似文献   

7.
Chen X  Berg HC 《Biophysical journal》2000,78(5):2280-2284
We studied changes in speed of the flagellar rotary motor of Escherichia coli when tethered cells or cells carrying small latex spheres on flagellar stubs were shifted from H(2)O to D(2)O or subjected to changes in external pH. In the high-torque, low-speed regime, solvent isotope effects were found to be small; in the low-torque, high-speed regime, they were large. The boundaries between these regimes were close to those found earlier in measurements of the torque-speed relationship of the flagellar rotary motor (, Biophys. J. 65:2201-2216;, Biophys. J., 78:1036-1041). This observation provides direct evidence that the decline in torque at high speed is due primarily to limits in rates of proton transfer. However, variations of speed (and torque) with shifts of external pH (from 4.7 to 8.8) were small for both regimes. Therefore, rates of proton transfer are not very dependent on external pH.  相似文献   

8.
9.
Rhodobacter sphaeroides can swim toward a wide range of attractants (a process known as taxis), propelled by a single rotating flagellum. The reversals of motor direction that cause tumbles in Eschericia coli taxis are replaced by brief motor stops, and taxis is controlled by a complex sensory system with multiple homologues of the E. coli sensory proteins. We tethered photosynthetically grown cells of R. sphaeroides by their flagella and measured the response of the flagellar motor to changes in light intensity. The unstimulated bias (probability of not being stopped) was significantly larger than the bias of tethered E. coli but similar to the probability of not tumbling in swimming E. coli. Otherwise, the step and impulse responses were the same as those of tethered E. coli to chemical attractants. This indicates that the single motor and multiple sensory signaling pathways in R. sphaeroides generate the same swimming response as several motors and a single pathway in E. coli, and that the response of the single motor is directly observable in the swimming pattern. Photo-responses were larger in the presence of cyanide or the uncoupler carbonyl cyanide 4-trifluoromethoxyphenylhydrazone (FCCP), consistent with the photo-response being detected via changes in the rate of electron transport.  相似文献   

10.
The study of glutathione status in aerobically grown Escherichia coli cultures showed that the total intracellular glutathione (GSHin + GSSGin) level falls by 63% in response to a rapid downshift in the extracellular pH from 6.5 to 5.5. The incubation of E. coli cells in the presence of 50 mM acetate or 10 micrograms/ml gramicidin S decreased the total intracellular glutathione level by 50 and 25%, respectively. The fall in the total intracellular glutathione level was accompanied by a significant decrease in the (GSHin:GSSGin) ratio. The most profound effect on the extracellular glutathione level was exerted by gramicidin S, which augmented the total glutathione level by 1.8 times and the (GSHout:GSSGout) ratio by 2.1 times. The gramicidin S treatment and acetate stress inhibited the growth of mutant E. coli cells defective in glutathione synthesis 5 and 2 times more severely than the growth of the parent cells. The pH downshift and the exposure of E. coli cells to gramicidin S and 50 mM acetate enhanced the expression of the sodA gene coding for superoxide dismutase SodA.  相似文献   

11.
Bacterial flagella are driven by a rotary motor that is energized by an electrochemical ion gradient across the cell membrane. In this study the torque generated by the flagellar motor was measured in tethered cells of a smooth-swimming Escherichia coli strain by using rotating electric fields to determine the relationship between the torque and speed over a wide range. By measuring the electric current applied to the sample cell and combining the data obtained at different viscosities, the torque of the flagellar motor was estimated up to 55 Hz, and also at negative rotation rates. By this method we have found that the torque of the flagellar motor linearly decreases with rotation rate from negative through positive rate of rotation. In addition, the dependence of torque upon temperature was also investigated. We showed that torque at the high speeds encountered in swimming cells had a much steeper dependence on temperature that at the low speeds encountered in tethered cells. From these results, the activation energy of the proton transfer reaction in the torque-generating unit was calculated to be about 7.0 x 10(-20) J.  相似文献   

12.
Sodium-coupled motility in a swimming cyanobacterium.   总被引:7,自引:2,他引:5       下载免费PDF全文
The energetics of motility in Synechococcus strain WH8113 were studied to understand the unique nonflagellar swimming of this cyanobacterium. There was a specific sodium requirement for motility such that cells were immotile below 10 mM external sodium and cell speed increased with increasing sodium levels above 10 mM to a maximum of about 15 microns/s at 150 to 250 mM sodium. The sodium motive force increased similarly with increasing external sodium from -120 to -165 mV, but other energetic parameters including proton motive force, electrical potential, the proton diffusion gradient, and the sodium diffusion gradient did not show such a correlation. Over a range of external sodium concentrations, cell speed was greater in alkaline environments than in neutral or acidic environments. Monensin and carbonyl cyanide m-chlorophenylhydrazone inhibited motility and affected components of sodium motive force but did not affect ATP levels. Cells were motile when incubated with 3-(3,4-dichlorophenyl)-1,1-dimethylurea and arsenate, which decreased cellular ATP to about 2% of control values. The results of this investigation are consistent with the conclusion that the direct source of energy for Synechococcus motility is a sodium motive force and that below a threshold of about -100 mV, cells are immotile.  相似文献   

13.
The bacterial flagellar motor is a tiny molecular machine that uses a transmembrane flux of H(+) or Na(+) ions to drive flagellar rotation. In proton-driven motors, the membrane proteins MotA and MotB interact via their transmembrane regions to form a proton channel. The sodium-driven motors that power the polar flagellum of Vibrio species contain homologs of MotA and MotB, called PomA and PomB. They require the unique proteins MotX and MotY. In this study, we investigated how ion selectivity is determined in proton and sodium motors. We found that Escherichia coli MotA/B restore motility in DeltapomAB Vibrio alginolyticus. Most hypermotile segregants isolated from this weakly motile strain contain mutations in motB. We constructed proteins in which segments of MotB were fused to complementary portions of PomB. A chimera joining the N terminus of PomB to the periplasmic C terminus of MotB (PotB7(E)) functioned with PomA as the stator of a sodium motor, with or without MotX/Y. This stator (PomA/PotB7(E)) supported sodium-driven motility in motA or motB E.coli cells, and the swimming speed was even higher than with the original stator of E.coli MotA/B. We conclude that the cytoplasmic and transmembrane domains of PomA/B are sufficient for sodium-driven motility. However, MotA expressed with a B subunit containing the N terminus of MotB fused to the periplasmic domain of PomB (MomB7(E)) supported sodium-driven motility in a MotX/Y-dependent fashion. Thus, although the periplasmic domain of PomB is not necessary for sodium-driven motility in a PomA/B motor, it can convert a MotA/B proton motor into a sodium motor.  相似文献   

14.
Escherichia coli batch cultures were grown under aerobic and anaerobic conditions on glucose with the substrate addition at pH 7.0. The cultures accumulated acetate in the medium at concentrations sufficient to inhibit the growth. This inhibitory effect of acetate was mediated apparently via its action on the intracellular pH. The inhibition of E. coli growth by acetate increased when the redox proton pump was switched off in the course of transition from aerobiosis to anaerobiosis and when the regulation of K+ fluxes was disordered in the presence of valinomycin. H+-ATPase was not essentially involved in maintaining the high rate of E. coli growth in the presence of acetate under aerobic conditions. If the activity of H+-ATPase was inhibited under anaerobic conditions at pH 7.0, the growth ceased after the dissipation of ionic gradients on the membrane. When CCCP was added under aerobic conditions, the growth did not stop at once if the medium had a pH of 7.6, but ceased immediately at pHout 7.0 in the glucose-salt medium.  相似文献   

15.
Streptococcus cremoris was grown in pH-regulated batch and continuous cultures with lactose as the energy source. During growth the magnitude and composition of the electrochemical proton gradient and the lactate concentration gradient were determined. The upper limit of the number of protons translocated with a lactate molecule during lactate excretion (the proton-lactate stoichiometry) was calculated from the magnitudes of the membrane potential, the transmembrane pH difference, and the lactate concentration gradient. In cells growing in continuous culture, a low lactate concentration gradient (an internal lactate concentration of 35 to 45 mM at an external lactate concentration of 25 mM) existed. The cell yield (Ymax lactose) increased with increasing growth pH. In batch culture at pH 6.34, a considerable lactate gradient (more than 60 mV) was present during the early stages of growth. As growth continued, the electrochemical proton gradient did not change significantly (from -100 to -110 mV), but the lactate gradient decreased gradually. The H+-lactate stoichiometry of the excretion process decreased from 1.5 to about 0.9. In nongrowing cells, the magnitude and composition of the electrochemical proton gradient was dependent on the external pH but not on the external lactate concentration (up to 50 mM). The magnitude of the lactate gradient was independent of the external pH but decreased greatly with increasing external lactate concentrations. At very low lactate concentrations, a lactate gradient of 100 mV existed, which decreased to about 40 mV at 50 mM external lactate. As a consequence, the proton-lactate stoichiometry decreased with increasing external concentrations of protons and lactate at pH 7.0 from 1 mM lactate to 1.1 at 50 mM lactate and at pH 5.5 from 1.4 at l mM lactate to 0.7 at 50 mM lactate. The data presented in this paper suggest that a decrease in external pH and an increase in external lactate concentration both result in lower proton-lactate stoichiometry values and therefore in a decrease of the generation of metabolic energy by the end product efflux process.  相似文献   

16.
S Khan  M Dapice    I Humayun 《Biophysical journal》1990,57(4):779-796
The effect of load and pH on the relation between proton potential and flagellar rotation has been studied in cells of a smooth-swimming Streptococcus strain. The driving potential, speeds of free-swimming bacteria, and rotation rates of bacteria tethered to glass by a single flagellum were measured. The relation between rotation rate of tethered bacteria and potential was remarkably linear up to nearly -200 mV. The relation between swimming speed and potential exhibited both saturation and threshold, as previously observed in other species. The form of these relations depended on pH. The equivalence of the electrical and chemical potential components of the proton potential in enabling swimming depended on the voltage. Our observations may be most simply accommodated by a kinetic scheme that links transmembrane proton transits to a tightly coupled work cycle. The properties of this scheme were elucidated by computer simulations of the experimental plots. These simulations indicated that the protonable groups that participate in the rate limiting reactions have a fractional electrical distance between three-fourths to all of the way toward the cytoplasm with a corresponding mean proton binding affinity of 10(-7.3)-10(-7.0) M, respectively.  相似文献   

17.
Many bacterial species swim using flagella. The flagellar motor couples ion flow across the cytoplasmic membrane to rotation. Ion flow is driven by both a membrane potential (V(m)) and a transmembrane concentration gradient. To investigate their relation to bacterial flagellar motor function we developed a fluorescence technique to measure V(m) in single cells, using the dye tetramethyl rhodamine methyl ester. We used a convolution model to determine the relationship between fluorescence intensity in images of cells and intracellular dye concentration, and calculated V(m) using the ratio of intracellular/extracellular dye concentration. We found V(m) = -140 +/- 14 mV in Escherichia coli at external pH 7.0 (pH(ex)), decreasing to -85 +/- 10 mV at pH(ex) 5.0. We also estimated the sodium-motive force (SMF) by combining single-cell measurements of V(m) and intracellular sodium concentration. We were able to vary the SMF between -187 +/- 15 mV and -53 +/- 15 mV by varying pH(ex) in the range 7.0-5.0 and extracellular sodium concentration in the range 1-85 mM. Rotation rates for 0.35-microm- and 1-microm-diameter beads attached to Na(+)-driven chimeric flagellar motors varied linearly with V(m). For the larger beads, the two components of the SMF were equivalent, whereas for smaller beads at a given SMF, the speed increased with sodium gradient and external sodium concentration.  相似文献   

18.
Excitatory signaling in bacterial probed by caged chemoeffectors.   总被引:8,自引:2,他引:6       下载免费PDF全文
Chemotactic excitation responses to caged ligand photorelease of rapidly swimming bacteria that reverse (Vibrio alginolyticus) or tumble (Escherichia coli and Salmonella typhimurium) have been measured by computer. Mutants were used to assess the effects of abnormal motility behavior upon signal processing times and test feasibility of kinetic analyses of the signaling pathway in intact bacteria. N-1-(2-Nitrophenyl)ethoxycarbonyl-L-serine and 2-hydroxyphenyl 1-(2-nitrophenyl) ethyl phosphate were synthesized. These compounds are a 'caged' serine and a 'caged' proton and on flash photolysis release serine and protons and attractant and repellent ligands, respectively, for Tsr, the serine receptor. The product quantum yield for serine was 0.65 (+/- 0.05) and the rate of serine release was proportional to [H+] near-neutrality with a rate constant of 17 s-1 at pH 7.0 and 21 degrees C. The product quantum yield for protons was calculated to be 0.095 on 308-nm irradiation but 0.29 (+/- 0.02) on 300-350-nm irradiation, with proton release occurring at > 10(5) s-1. The pH jumps produced were estimated using pH indicators, the pH-dependent decay of the chromophoric aci-nitro intermediate and bioassays. Receptor deletion mutants did not respond to photorelease of the caged ligands. Population responses occurred without measurable latency. Response times increased with decreased stimulus strength. Physiological or genetic perturbation of motor rotation bias leading to increased tumbling reduced response sensitivity but did not affect response times. Exceptions were found. A CheR-CheB mutant strain had normal motility, but reduced response. A CheZ mutant had tumbly motility, reduced sensitivity, and increased response time to attractant, but a normal repellent response. These observations are consistent with current ideas that motor interactions with a single parameter, namely phosphorylated CheY protein, dictate motor response to both attractant and repellent stimuli. Inverse motility motor mutants with extreme rotation bias exhibited the greatest reduction in response sensitivity but, nevertheless, had normal attractant response times. This implies that control of CheY phosphate concentration rather than motor reactions limits responses to attractants.  相似文献   

19.
When bacterial cells are tethered to glass by their flagella, many of them spin. On the basis of experiments with tethered cells it has generally been thought that the motor which drives the flagellum is a two-state device, existing in either a counterclockwise or a clockwise state. Here we show that a third state of the motor is that of pausing, the duration and frequency of which are affected by chemotactic stimuli. We have recorded on video tape the rotation of tethered Escherichia coli and Salmonella typhimurium cells and analyzed the recordings frame by frame and in slow motion. Most wild-type cells paused intermittently. The addition of repellents caused an increase in the frequency and duration of the pauses. The addition of attractants sharply reduced the number of pauses. A chemotaxis mutant which lacks a large part of the chemotaxis machinery owing to a deletion of the genes from cheA to cheZ did not pause at all and did not respond to repellents by pausing. A tumbly mutant of S. typhimurium responded to repellents by smooth swimming and to attractants by tumbling. When tethered, these cells exhibited a normal rotational response but an inverse pausing response to chemotactic stimuli: the frequency of pauses decreased in response to repellents and increased in response to attractants. It is suggested that (i) pausing is an integral part of bacterial motility and chemotaxis, (ii) pausing is independent of the direction of flagellar rotation, and (iii) pausing may be one of the causes of tumbling.  相似文献   

20.
The oxidation-reduction midpoint potentials, Em, of the FAD and active site disulfide couples of Escherichia coli thioredoxin reductase have been determined from pH 5.5 to 8.5. The FAD and disulfide couples have similar Em values and thus a linked equilibrium of four microscopic enzyme oxidation-reduction states exists. The binding of phenylmercuric acetate to one enzyme form could be monitored which allowed solving the four microscopic Em values. The Em values at pH 7.0 and 12 degrees C of the four couples of thioredoxin reductase are: (S)2-enzyme-FAD/FADH2 = -0.243 V, (SH)2-enzyme-FAD/FADH2 = -0.260 V, (FAD)-enzyme-(S)2/(SH)2 = -0.254 V, and (FADH2)-enzyme-(S)2/(SH)2 = -0.271 V. Thus, at pH 7.0, the FAD and disulfide moieties have a 0.017-V negative interaction and Em values which are different by 0.011 V. The delta Em/delta pH of the FAD couples E2m and E3m are about 0.060 V/pH throughout the pH range studied, showing an approximately 2-proton stoichiometry of reduction of the enzyme FAD. The delta Em/delta pH of the disulfide couples E1m and E4m are about 0.052 V/pH from pH 5.5 to 8.5, showing an apparently nonintegral proton stoichiometry of reduction of 1.8 in this pH range. This proton stoichiometry suggests the presence of a base with an ionization behavior that is linked to the oxidation-reduction state of the disulfide. A novel method is presented for determining the pK values on oxidized and reduced enzyme which agrees with the less accurate classical method. The proton stoichiometry results are consistent with the presence of a thiol-base ion pair in which the pK of the base is elevated from 7.6 in disulfide containing enzyme to greater than 8.5 upon forming an ion pair with a thiol anion of pK 7.0 generated upon reduction of the disulfide. The fluorescence of the FAD in thioredoxin reductase decreases as the pH is lowered with a pK of 7.0, direct evidence for a base near the FAD probably distinct from the base interacting with the dithiol.  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