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.
Using 1-(2-nitrophenyl)ethyl sulfate (caged sulfate) as a photoactivatable caged proton, we could induce complete acid unfolding of myoglobin with a single nanosecond laser pulse. This was possible because of the high ( approximately mM) concentration of protons released by the photolabile compound. The ability of the compound to produce a large pH jump arises because the other photoproducts (2-nitrosoacetophenone and sulfate ion) do not buffer the released protons. The complete time course of the unfolding kinetics, spanning a range from milliseconds to several seconds, could be accurately reproduced by monitoring absorbance changes in the visible spectrum at 633 nm.  相似文献   

2.
A fast perturbation in proton concentration can be induced in aqueous solution using a pulsed ultraviolet laser and suitable photolabile compounds which, upon photoexcitation, irreversibly release protons. The volume change and the rate constant for the reaction of the photodetached protons with proton-accepting groups in solution can be monitored using time resolved photoacoustics. A typical proton concentration jump of 1 microM can be obtained with a 200-microJ laser pulse at 308 nm. Reaction dynamics from 20 ns to 5 micros can be easily followed. The methodology we establish represents a direct, time-resolved measurement of the reaction volume in proton transfer processes and an extension to the nanosecond-microsecond range of traditional relaxation techniques, such as stopped-flow. We report example applications to reactions involving simple molecules and polypeptides.  相似文献   

3.
Proton flux measurements from tissues in buffered solution   总被引:5,自引:1,他引:4  
Proton movement across plant cell membranes is part of many important physiological processes. The net proton flux to or from tissues can be determined non-invasively by measuring the proton electrochemical potential gradient in the adjacent solution. In buffered solution, some of the protons crossing the tissue boundary diffuse as proto-nated buffer whose flux is not included in the flux calculated from the proton (hydrogen ion) electrochemical gradient. In this theoretical paper, it is shown how experimenters can calculate the protonated buffer flux from the measured proton flux in solution. The ratio of these two components of total proton flux depends on the pH of the solution and on the concentration and pK of the buffer. For a given concentration of a buffer which has a single pK, the flux ratio rises with pH when the solution pH is lower than the buffer pK. The slope is about 2 on a log10 scale. As the pH increases above the pK, the flux ratio levels off to approach its maximum. With mixed buffers, or one having two or more pK values, the flux ratios are additive: each buffer acts independently based on its concentration and its pK value. Unbuffered solutions always have the buffering effects of water itself and also of carbonates due to carbon dioxide dissolved from the atmosphere. In unbuffered solutions at pH 6, the flux carried by water and carbonate is about 1 % of the measured proton flux. This validates measurements of proton flux from tissues, made by a number of workers, in unbuffered solutions below pH 6.  相似文献   

4.
The problem of predicting the kinetics of proton efflux and the decay of the internal proton concentration for vesicles containing one or more buffers for which the internal proton concentration is initially higher than that of the surrounding medium is examined. An analytical solution is derived that describes the time course of the proton efflux from vesicles and the decay of the internal proton concentration under conditions of zero transmembrane electric potential. The effect of the internal buffers is to increase the time required for the proton concentration gradient to equilibrate across the membrane. To simplify the analysis we assume that the equilibration of the internal and external proton activity is due primarily to proton diffusion through the membrane, and not to hydroxyl ion flux. For a vesicle containing a single buffer the solution requires six independent physical parameters: the initial internal proton concentration, the external proton concentration, the ratio of the vesicle surface area to the internal volume, the permeability coefficient of the membrane for protons, the total concentration of the internal buffer, and the equilibrium constant for the dissociation of the internal buffer. Determination of these physical values is sufficient to predict the time dependence of the internal proton concentration and of the proton efflux. Over a pH range that is below or near the pK of the internal buffer the solution is complex. However, if the initial pH is one unit or more higher than the pK of the internal buffer the kinetics of the internal proton concentration and proton efflux can be described by a pseudo first order reaction. In this case the apparent rate constant depends linearly on the permeability coefficient and is dominated by the total internal buffer concentration and its pK. For example, increasing the internal buffer concentration inside a vesicle by 10-fold results in an approximately 10-fold increase in the half-time of the proton efflux kinetics. The theoretical analysis is applied to thylakiod vesicles using experimentally determined values for the physical parameters. The predictions of the analysis are compared to experimentally observed kinetics.  相似文献   

5.
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   

6.
Arg(82) is one of the four buried charged residues in the retinal binding pocket of bacteriorhodopsin (bR). Previous studies show that Arg(82) controls the pK(a)s of Asp(85) and the proton release group and is essential for fast light-induced proton release. To further investigate the role of Arg(82) in light-induced proton pumping, we replaced Arg(82) with histidine and studied the resulting pigment and its photochemical properties. The main pK(a) of the purple-to-blue transition (pK(a) of Asp(85)) is unusually low in R82H: 1.0 versus 2.6 in wild type (WT). At pH 3, the pigment is purple and shows light and dark adaptation, but almost no light-induced Schiff base deprotonation (formation of the M intermediate) is observed. As the pH is increased from 3 to 7 the M yield increases with pK(a) 4.5 to a value approximately 40% of that in the WT. A transition with a similar pK(a) is observed in the pH dependence of the rate constant of dark adaptation, k(da). These data can be explained, assuming that some group deprotonates with pK(a) 4.5, causing an increase in the pK(a) of Asp(85) and thus affecting k(da) and the yield of M. As the pH is increased from 7 to 10.5 there is a further 2.5-fold increase in the yield of M and a decrease in its rise time from 200 &mgr;s to 75 &mgr;s with pK(a) 9. 4. The chromophore absorption band undergoes a 4-nm red shift with a similar pK(a). We assume that at high pH, the proton release group deprotonates in the unphotolyzed pigment, causing a transformation of the pigment into a red-shifted "alkaline" form which has a faster rate of light-induced Schiff base deprotonation. The pH dependence of proton release shows that coupling between Asp(85) and the proton release group is weakened in R82H. The pK(a) of the proton release group in M is 7.2 (versus 5.8 in the WT). At pH < 7, most of the proton release occurs during O --> bR transition with tau approximately 45 ms. This transition is slowed in R82H, indicating that Arg(82) is important for the proton transfer from Asp(85) to the proton release group. A model describing the interaction of Asp(85) with two ionizable residues is proposed to describe the pH dependence of light-induced Schiff base deprotonation and proton release.  相似文献   

7.
The synthesis of a new, pK 8.05 acrylamido weak base for isoelectric focusing in immobilized pH gradients (IPG) is here reported. This compound N,N-bis(2-hydroxyethyl)-N'-acryloyl-1,3-diaminopropane is strongly hydrophilic, and thus inhibits any potential hydrophobic interaction among proteins and the grafted basic groups in an IPG matrix. In addition, this novel buffer represents a step ahead towards the goal of closing the 'gap' between the commercially available Immobilines, pK 7.0 and 8.5. Owing to the large distance between these two neighboring pK values, it is difficult to arrange for linear narrow pH gradients in this region. IPG compositions obtained with this new buffer give highly linear pH gradients and protein profiles identical to those obtained with commercial Immobilines.  相似文献   

8.
J H Bradbury  J A Carver 《Biochemistry》1984,23(21):4905-4913
In paramagnetic metmyoglobin, cyanomyoglobin (CNMb), and deoxymyoglobin, His-36 has a high pK (approximately 8), and the NMR titration behavior of the H-2 resonance is perturbed, due to the presence at low pH of a hydrogen bond with Glu-38, which is broken at high pH. The His-36 H-4 resonance shows no shift with pK approximately 8 because of two opposing chemical shift effects but monitors the titration of nearby Glu-36 (pK = 5.6). In diamagnetic derivatives [(carbon monoxy)myoglobin (COMb) and oxymyoglobin (oxyMb)], the titration behavior of His-36 H-2 and H-4 resonances is normalized (pK approximately 6.8). The very slight alkaline Bohr effect in sperm whale myoglobin (Mb) is interpreted in terms of the pK change of His-36 from deoxyMb to oxyMb and compensating pK changes in the opposite direction of other unspecified groups. In sperm whale COMb at 40 degrees C, the distal histidine (His-64) and His-97 have pK values of 5.0 and 5.9. The meso proton resonances remote from these groups do not show a titration shift, but the nearby gamma-meso proton (pK = 5.3) responds to titration of both histidines, and the upfield Val-68 methyl at -2.3 ppm (pK = 4.7) witnesses the titration of nearby His-64. At 20 degrees C, the latter resonance is reduced in size, and a second resonance occurs at -2.8 ppm, which is insensitive to pH and, hence, more remote from His-64. Both resonances arise from two conformations of Val-68 in slow equilibrium. In oxyMb at 20 degrees C, only the latter resonance is observed, presumably because of the steric restrictions imposed by the hydrogen bond between ligand and His-64 in oxyMb, which is not present in COMb. In oxyMb the pK of His-97 (5.6) is similar to that of the meso proton resonances (5.5) and to the pK of other pH-dependent processes, including the very small acid Bohr effect. It is likely that these processes are controlled by the titration of His-97.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Intestinal epithelial membrane transport of L-lactic acid was characterized using rabbit jejunal brush-border membrane vesicles (BBMVs). The uptake of L-[(14)C]lactic acid by BBMVs showed an overshoot phenomenon in the presence of outward-directed bicarbonate and/or inward-directed proton gradients. Kinetic analysis of L-[(14)C]lactic acid uptake revealed the involvement of two saturable processes in the presence of both proton and bicarbonate gradients. An arginyl residue-modifying agent, phenylglyoxal, inhibited L-[(14)C]lactic acid transport by the proton cotransporter, but not by the anion antiporter. The initial uptakes of L-[(14)C]lactic acid which are driven by bicarbonate ion and proton gradients were inhibited commonly by monocarboxylic acids and selectively by anion exchange inhibitor 4, 4'-diisothiocyanostilbene-2,2'-disulfonic acid and protonophore carbonylcyanide p-trifluoromethoxyphenylhydrazone, respectively. These observations demonstrate that L-lactic acid is transported across the intestinal brush-border membrane by multiple mechanisms, including an anion antiporter and a previously known proton cotransporter.  相似文献   

10.
The activity of purified DNA photolyase from Baker's yeast is enhanced by a compound (Activator (III)) obtained from yeast by chloroform extraction ion exchange chromatography and gel filtration. Thin layer chromatography and spectral data indicate that the compound is homogeneous. Activator III emits at 350 and 440 nm when excited at 290 nm, and emits at 440 nm when excited at 358 nm. After acid hydrolysis, emission at 440 nm is produced only by excitation at 358 nm, indicating that activator (III) contains two separate chromophoric moieties. The chromophore excited by 358 nm light has a pK of 9-11, while the other chromophore has a pK of 4-5, and possibly of 9-11. The enhancement of photolytic activity by activator (III) at a concentration equimolar with that of the enzyme and the similarity of the fluorescent spectra of the activator with that of heat-denatured photolyase, suggests that the activator may be the chromophore associated with the enzyme.  相似文献   

11.
The ion permeability properties of dark adapted bovine rod outer segment disk membranes were studied using light scattering to monitor osmotic responses of disks to various salts and ionophores. A preparation procedure is presented which provides very fresh rod outer segment material with mostly intact stacked disks, but with perforated plasma membrane. It is shown that in this preparation the disks (or rod sacs) are the only osmotically responding compartments and that these responses can be readily monitored by means of light-scattering techniques. The disk membrane is found under the conditions tested, to possess no measurable permeability to cations Na+, Ca2+, Mg2+ nor the the anions Cl-, Br-, NO3-, SO4(2-), H2PO4- and HPO4(2-). There is a considerable K+ permeability, which can be completely abolished by millimolar amounts of divalent cations. The proton permeability of the disk membrane is found to depend dramatically upon the preparation procedure and duration. The fresher the material used the lower is the proton permeability measured. In our freshest preparations, even after freeze-thawing in liquid nitrogen, the disks exhibit an H+ permeability which is so low that it cannot be measured with the techniques used in this study. Even in mitochondrial or chloroplast membmranes, in which proton gradients and therefore a low proton conductance play an essential role, such low proton permeabilities have not been found. This would suggest that proton gradients across the disk membrane could play an important role in the physiological function of the photoreceptor cell. In summary it can be said that the disk membrane, apparently more than any other natural membrane system studied so far, is capable of retaining ion gradients for extended periods of time.  相似文献   

12.
Tyrosine-83, a residue which is conserved in all halobacterial retinal proteins, is located at the extracellular side in helix C of bacteriorhodopsin. Structural studies indicate that its hydroxyl group is hydrogen bonded to Trp189 and possibly to Glu194, a residue which is part of the proton release complex (PRC) in bacteriorhodopsin. To elucidate the role of Tyr83 in proton transport, we studied the Y83F and Y83N mutants. The Y83F mutation causes an 11 nm blue shift of the absorption spectrum and decreases the size of the absorption changes seen upon dark adaptation. The light-induced fast proton release, which accompanies formation of the M intermediate, is observed only at pH above 7 in Y83F. The pK(a) of the PRC in M is elevated in Y83F to about 7.3 (compared to 5.8 in WT). The rate of the recovery of the initial state (the rate of the O --> BR transition) and light-induced proton release at pH below 7 is very slow in Y83F (ca. 30 ms at pH 6). The amount of the O intermediate is decreased in Y83F despite the longer lifetime of O. The Y83N mutant shows a similar phenotype in respect to proton release. As in Y83F, the recovery of the initial state is slowed several fold in Y83N. The O intermediate is not seen in this mutant. The data indicate that the PRC is functional in Y83F and Y83N but its pK(a) in M is increased by about 1.5 pK units compared to the WT. This suggests that Tyr83 is not the main source for the proton released upon M formation in the WT; however, Tyr83 is involved in the proton release affecting the pK(a) of the PRC in M and the rate of proton transport from Asp85 to PRC during the O --> bR transition. Both the Y83F and the Y83N mutations lead to a greatly decreased functionality of the pigment at high pH because most of the pigment is converted into the inactive P480 species, with a pK(a) 8-9.  相似文献   

13.
Uptake of glutamate from the synaptic cleft is mediated by high affinity transporters and is driven by Na(+), K(+), and H(+) concentration gradients across the membrane. Here, we characterize the molecular mechanism of the intracellular pH change associated with glutamate transport by combining current recordings from excitatory amino acid carrier 1 (EAAC1)-expressing HEK293 cells with a rapid kinetic technique with a 100-micros time resolution. Under conditions of steady state transport, the affinity of EAAC1 for glutamate in both the forward and reverse modes is strongly dependent on the pH on the cis-side of the membrane, whereas the currents at saturating glutamate concentrations are hardly affected by the pH. Consistent with this, the kinetics of the pre-steady state currents, measured after saturating glutamate concentration jumps, are not a function of the pH. In addition, we determined the deuterium isotope effect on EAAC1 kinetics, which is in agreement with proton cotransport but not OH(-) countertransport. The results can be quantitatively explained with an ordered binding model that includes a rapid proton binding step to the empty transporter followed by glutamate binding and translocation of the proton-glutamate-transporter complex. The apparent pK of the extracellular proton binding site is approximately 8. This value is shifted to approximately 6.5 when the substrate binding site is exposed to the cytoplasm.  相似文献   

14.
Light absorbed by bacteriorhodopsin (bR) leads to a proton being released at the extracellular surface of the purple membrane. Structural studies as well as studies of mutants of bR indicate that several groups form a pathway for proton transfer from the Schiff base to the extracellular surface. These groups include D85, R82, E204, E194, and water molecules. Other residues may be important in tuning the initial state pK(a) values of these groups and in mediating light-induced changes of the pK(a) values. A potentially important residue is R134: it is located close to E194 and might interact electrostatically to affect the pK(a) of E194 and light-induced proton release. In this study we investigated effects of the substitution of R134 with a histidine on light-induced proton release and on the photocycle transitions associated with proton transfer. By measuring the light-induced absorption changes versus pH, we found that the R134H mutation results in an increase in the pK(a) of the proton release group in both the M (0.6 pK unit) and O (0.7 pK unit) intermediate states. This indicates the importance of R134 in tuning the pK(a) of the group that, at neutral and high pH, releases the proton upon M formation (fast proton release) and that, at low pH, releases the proton simultaneously with O decay (slow proton release). The higher pK(a) of the proton release group found in R134H correlates with the slowing of the rate of the O --> bR transition at low pH and probably is the cause of this slowing. The pH dependence of the fraction of the O intermediate is altered in R134H compared to the WT but is similar to that in the E194D mutant: a very small amount of O is present at neutral pH, but the fraction of O increases greatly upon decreasing the pH. These results provide further support for the hypothesis that the O --> bR transition is controlled by the rate of deprotonation of the proton release group. These data also provide further evidence for the importance of the R134-E194 interaction in modulating proton release from D85 after light has led to its being protonated.  相似文献   

15.
Conductance routes for protons across membrane barriers   总被引:1,自引:0,他引:1  
J Bramhall 《Biochemistry》1987,26(10):2848-2855
Simple phospholipid bilayers show a high level of permeability to protons; in spite of this fact, large proton gradients existing across such bilayers may decay very slowly. In sealed systems, the free movement of protons across a membrane barrier is severely restricted by the coincident development of a proton diffusion potential. Using the fluorescent weak acid N-[5-(dimethylamino)naphth-1-ylsulfonyl]glycine strongly buffered systems movement of the small number of protons giving rise to this electrical potential is insufficient to perturb the proton concentration gradient; significant flux of protons (and hence significant collapse of the concentration gradient) can only occur if protons traverse the membrane as part of an electroneutral complex or if there is a balancing flow of appropriate counterions. In both instances, proton flux is obligatorily coupled to the translocation of species other than protons. In weakly buffered systems, the small initial uncoupled electrogenic flux of protons may significantly alter the concentration gradient. This initial rapid gradient collapse caused by uncoupled electrogenic proton movements is then superimposed upon the residual collapse attributable to tightly coupled proton flux. The initial uncoupled electrogenic proton flux shows a temperature dependence very similar to that demonstrated for water permeation across simple lipid bilayers; upon cooling, there is a sharp decrease in flux at the temperature coinciding with the main gel-liquid-crystalline phase transition of the lipid. The coupled proton flux shows a markedly different temperature dependence with no dramatic change in rate at the phase transition temperature and strong similarity to the behavior previously seen with solutes known to be permeating as electrically neutral compounds.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Pyrene-1-carboxylic acid has a pK of 4.0 in the ground state and 8.1 in the singlet electronic excited state. In the pH range of physiological interest (pH approximately 5-8), the ground state compound is largely ionized as pyrene-1-carboxylate, but protonation of the excited state molecule occurs when a proton donor reacts with the carboxylate during the excited state lifetime of the fluorophore. Both forms of the pyrene derivatives are fluorescent, and in this work the protonation reaction was measured by monitoring steady-state and time-resolved fluorescence. The rate of protonation of pyrene-COO(-) by acetic, chloroacetic, lactic, and cacodylic acids is a function of DeltapK, as predicted by Marcus theory. The rate of proton transfer from these acids saturates at high concentration, as expected for the existence of an encounter complex. Trihydrogen-phosphate is a much better proton donor than dihydrogen- and monohydrogen-phosphate, as can be seen by the pH dependence. The proton-donating ability of phosphate does not saturate at high concentrations, but increases with increasing phosphate concentration. We suggest that enhanced rate of proton transfer at high phosphate concentrations may be due to the dual proton donating and accepting nature of phosphate, in analogy to the Grotthuss mechanism for proton transfer in water. It is suggested that in molecular structures containing multiple phosphates, such as membrane surfaces and DNA, proton transfer rates will be enhanced by this mechanism.  相似文献   

17.
A new acrylamido buffer has been synthesized, for use in isoelectric focusing in immobilized pH gradients. This compound (2-acrylamido glycolic acid) has a pK = 3.1 (at 25 degrees C, 20 mM concentration during titration) and is used, by titration with the pK 9.3 Immobiline, to produce a linear pH gradient in the pH 2.5-3.5 interval. Pepsin (from pig stomach) focused in this acidic pH gradient is resolved into four components, two major (with pI values 2.76 and 2.78) and two minor (having pI values 2.89 and 2.90). This is the first time that such strongly acidic proteins could be focused in an immobilized pH gradient. Even in conventional isoelectric focusing in amphoteric buffers it has been impossible to focus reproducibly very-low-pI macromolecules.  相似文献   

18.
The existence of a proton-selective pathway through a protein is a common feature of voltage-gated proton channels and a number of molecules that play pivotal roles in bioenergetics. Although the functions and structures of these molecules are quite diverse, the proton conducting pathways share a number of fundamental properties. Conceptual parallels include the translocation by hydrogen-bonded chain mechanisms, problems of supply and demand, equivalence of chemical and electrical proton gradients, proton wells, alternating access sites, pK(a) changes induced by protein conformational change, and heavy metal participation in proton transfer processes. An archetypal mechanism involves input and output proton pathways (hydrogen-bonded chains) joined by a regulatory site that switches the accessibility of the bound proton from one 'channel' to the other, by means of a pK(a) change, molecular movement, or both. Although little is known about the structure of voltage-gated proton channels, they appear to share many of these features. Evidently, nature has devised a limited number of mechanisms to accomplish various design strategies, and these fundamental mechanisms are repeated with minor variation in many superficially disparate molecules.  相似文献   

19.
A dioleoylphosphatidylcholine unilamellar vesicle model system was used to determine proton permeability. The fluorescence of the pH reporter group, pyranine, trapped within vesicles with a difference in pH across the bilayer, was digitized and analyzed with numerical integration. When H+ flux was initiated by the acidification of the external buffer (acid jump), the apparent H+ permeability was found to be a linear function of the reciprocal of the internal H+ concentration with the slope inversely proportional to the initial size of the H+ gradient. When flux was initiated by the alkalinization of the external buffer (base jump), the apparent permeability coefficient was constant for each external H+ concentration. However, the value of the apparent permeability was linearly dependent on the reciprocal of the external H+. The possibility that carbonates (carbon dioxide, carbonic acid, bicarbonate and carbonate) could be acting as proton carriers was tested by adding millimolar concentrations of bicarbonate to solutions greatly reduced in carbonates. The slopes of the graphs of apparent permeability coefficient vs. reciprocal H+ were linear functions of added bicarbonate concentration for both acid and base jump conditions. These observations were interpreted in terms of a model suggesting that carbonic acid or carbon dioxide together with bicarbonate was an efficient proton carrier across phospholipid bilayers.  相似文献   

20.
Enthalpy changes of alpha-chymotrypsin acylation by 3-(2-furyl)acryloylimidazole (FAI) were calorimetrically determined as a function of pH. By observing the functional dependence of acylation enthalpies on buffer ionization heats, a complex pH profile was obtained describing proton release accompanying formation of acyl-enzyme. A pKa of 4.0 for FAI ionization and apparent pKa values of 6.8, 7.55 and 8.8 on the enzyme were used to account for the proton release data. A model which accounts for the proton release behavior was used to fit the acylation enthalpy data and values for the apparent dissociation enthalpies of the groups involved were obtained along with a pH-independent intrinsic enthalpy of acylation. This model suggests a group with an apparent pK = 6.8 and delta Hion = 8.7 kcal/mol which is perturbed to a pK of 7.55 and delta Hion = 7.6 kcal/mol on attachment of the acyl moiety to the enzyme. The apparent ionization enthalpy change for the active-inactive transition (pK3 = 8.8; delta H = 3.0 kcal/mol) corresponds with that calculated from the data of Fersht (J. Mol. Biol. 64 (1972) 497). The pH-independent intrinsic enthalpy of acylation (delta H = -7.9 kcal/mol) is corrected for group ionizations linked to the acylation process. Consequently, it more closely reflects molecular processes of interest such as substrate binding, covalent bond rearrangement, and product release.  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