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 31 毫秒
1.
The blue light receptor photoactive yellow protein (PYP) displays rhodopsin-like photochemistry based on the trans to cis photoisomerization of its p-coumaric acid chromophore. Here, we report that protein refolding from the acid-denatured state of PYP mimics the last photocycle transition in PYP. This implies a direct link between transient protein unfolding and photosensory signal transduction. We utilize this link to study general issues in protein folding. Chromophore trans to cis photoisomerization in the acid-denatured state strongly decelerates refolding, and converts the pH dependence of the barrier for refolding from linear to nonlinear. We propose transition state movement to explain this phenomenon. The cis chromophore significantly stabilizes the acid-denatured state, but acidification of PYP results in the accumulation of the acid-denatured state containing a trans chromophore. This provides a clear example of kinetic control in a protein unfolding reaction. These results demonstrate the power of PYP as a light-triggered model system to study protein folding.  相似文献   

2.
Kawamura M  Kiyotake R  Kudo K 《Chirality》2002,14(9):724-726
Novel photoresponsive axially chiral monophosphine ligands containing azobenzene moiety were prepared and applied to a palladium-catalyzed allylic alkylation. The reaction of rac-1,3-diphenyl-2-propenyl acetate gave the alkylated products with up to 90% enantiomeric excess. The ligand exhibited a trans to cis photoisomerization upon irradiation with UV light.  相似文献   

3.
Horseradish peroxidase mutants containing L-p-phenylazophenylalanine (azoAla) at various positions were synthesized by using an Escherichia coli in vitro translation system. Among the 15 mutants examined, four mutants containing a single azoAla unit at the 6th, 68th, 142nd, and 179th positions, respectively, retained the peroxidase activity. The activity of the Phe68azoAla mutant was higher when the azobenzene group was in the cis form than in the trans form. On the contrary, the activity of the Phe179azoAla mutant disappeared when the azobenzene group was photoisomerized to the cis form, but recovered in the trans form. In the latter mutant, therefore, an on/off photoswitching of the peroxidase activity was attained.  相似文献   

4.
An azobenzene group containing beta-amino acid N-Fmoc-4-aminomethyl phenylazobenzoic acid was synthesized and with the exception of the C-terminal amino acid residue was substituted by solid-phase peptide synthesis into all positions of the FLAG sequence (DYKDDDDK), an octapeptide capable of specific interaction with the monoclonal antibody 4E11. The trans state of the beta-amino acid was thermodynamically more stable than the cis state. However, the molecule could be switched into the cis conformation by illumination at 340 nm. Peptides containing the artificial amino acid also became photoresponsive. In the absence of light, the spontaneous back-isomerization into the trans conformation of the photoresponsive was extremely slow (>8 h no significant increase in trans content). When illuminated with visible light (440 nm), the back-isomerization from the cis to the trans state was accelerated and occurred with a half-life of approximately 10 min. The cis form of the photopeptides was more hydrophilic than the trans form, as evidenced by differences in the retention time of the two isomeric forms in reversed-phase chromatography. Photopeptides that contained the intact sequences responsible for binding of the FLAG tag to the antibody, namely, the DYK motive at the N-terminus, showed binding to the antibody in both a dot blot immunoassay and in Biacore binding studies, albeit with lower affinity than the unmodified FLAG sequence. Peptides with a substitution in positions 4-6 showed differences in binding strength between the trans and the cis form in the Biacore studies, no such difference could be observed for the peptide with a substitution in position 7.  相似文献   

5.
The substitution of trans- for half of the cis-monounsaturated fatty acids in the diet of Macaca fasicularis monkeys resulted in alterations in erythrocyte fatty acid composition and insulin receptor properties but not in membrane fluidity. Both cis and trans diets contained 10% fat and similar fatty acid compositions, except that approximately 50% of the cis-octadecenoate (c-18:1) in the cis diet was replaced with trans-octadecenoate isomers (t-18:1) in the trans diet. Compared with the cis diet, the trans diet resulted in the incorporation of approximately 11% t-18:1, an approximately 50% decrease in c-18:1, an approximately 16% decrease in total saturated fatty acids, and an approximately 20% increase in 18:2(n-6) in erythrocyte membrane lipids. The increase in 18:2(n-6) may reflect on homeostatic mechanisms designed to maintain overall membrane fluidity, as no diet-related changes in fluidity were observed with diphenylhexatriene steady state fluorescence polarization. Values observed for insulin binding and insulin receptor number were higher and binding affinity was lower in monkeys fed the cis diet. In the absence of an effect on overall membrane fluidity, altered receptor activity suggests that insulin receptor activity is dynamic, requiring specific fluid membrane subdomains or highly specific fatty acid-protein interactions.  相似文献   

6.
Apoptotic pathways in breast cancer cells are frequently altered, reducing the efficiency of radiotherapy. Conjugated linoleic acid (CLA), known to trigger apoptosis, was tested as radiosensitizer in breast cancer cells MCF-7 and MDA-MB-231. The CLA-mix, made up of the isomers CLA-9cis 11trans and CLA-10trans 12cis, was compared to three purified isomers, i.e., the CLA-9cis 11cis, CLA-9cis 11trans, and CLA-10trans 12cis. Using the apoptotic marker YO-PRO-1, the CLA-9cis 11cis at 50 micro mol/L turned out to be the best apoptotic inducer leading to a 10-fold increase in MCF-7 cells and a 2,5-fold increase in MDA-MB-231 cells, comparatively to the CLA-mix. Contrary to previous studies on colorectal and prostate cancer cells, CLA-10trans 12cis does not lead to an apoptotic response on breast cancer cell lines MCF-7 and MDA-MB-231. Our results also suggest that the main components of the CLA-mix (CLA-9cis 11trans and CLA-10trans 12cis) are not involved in the induction of apoptosis in the breast cancer cells studied. A dose of 5 Gy did not induce apoptosis in MCF-7 and MDA-MB-231 cells. The addition of CLA-9cis 11cis or CLA-mix has allowed us to observe a radiation-induced apoptosis, with the CLA-9cis 11cis being about 8-fold better than the CLA-mix. CLA-9cis 11cis turned out to be the best radiosensitizer, although the isomers CLA-9cis 11trans and CLA-10trans 12cis have also reduced the cell survival following irradiation, but using a mechanism not related to apoptosis. In conclusion, the radiosensitizing property of CLA-9cis 11cis supports its potential as an agent to improve radiotherapy against breast carcinoma.  相似文献   

7.
The aim of this work was to elucidate the roles of individual residues within the flexible second binding loop of human cystatin A in the inhibition of cysteine proteases. Four recombinant variants of the inhibitor, each with a single mutation, L73G, P74G, Q76G or N77G, in the most exposed part of this loop were generated by PCR-based site-directed mutagenesis. The binding of these variants to papain, cathepsin L, and cathepsin B was characterized by equilibrium and kinetic methods. Mutation of Leu73 decreased the affinity for papain, cathepsin L and cathepsin B by approximately 300-fold, >10-fold and approximately 4000-fold, respectively. Mutation of Pro74 decreased the affinity for cathepsin B by approximately 10-fold but minimally affected the affinity for the other two enzymes. Mutation of Gln76 and Asn77 did not alter the affinity of cystatin A for any of the proteases studied. The decreased affinities were caused exclusively by increased dissociation rate constants. These results show that the second binding loop of cystatin A plays a major role in stabilizing the complexes with proteases by retarding their dissociation. In contrast with cystatin B, only one amino-acid residue of the loop, Leu73, is of principal importance for this effect, Pro74 assisting to a minor extent only in the case of cathepsin B binding. The contribution of the second binding loop of cystatin A to protease binding varies with the protease, being largest, approximately 45% of the total binding energy, for inhibition of cathepsin B.  相似文献   

8.
Thermally unfolded staphylococcal nuclease has been rapidly quenched to temperatures near 0 degree C and the refolding behavior examined using an NMR kinetic experiment. Unfolded protein, exhibiting random coil chemical shifts, persists following the quench and refolds in two distinct kinetic phases. A protein folding intermediate with a trans Lys 116-Pro 117 peptide bond is transiently overpopulated and relaxes to the predominantly cis native cis-trans equilibrium. The rate of trans-->cis isomerization in the native-like nuclease intermediate is approximately 100-fold faster than that observed in a Lys-Pro model peptide. The activation enthalpy of 20 kcal/mol observed for the nuclease Lys 116-Pro 117 peptide bond is comparable to that observed for other X-Pro isomerizations.  相似文献   

9.
Y Zhang  E R Kantrowitz 《Biochemistry》1989,28(18):7313-7318
Lysine-60 in the regulatory chain of aspartate transcarbamoylase has been changed to an alanine by site-specific mutagenesis. The resulting enzyme exhibits activity and homotropic cooperativity identical with those of the wild-type enzyme. The substrate concentration at half the maximal observed specific activity decreases from 13.3 mM for the wild-type enzyme to 9.6 mM for the mutant enzyme. ATP activates the mutant enzyme to the same extent that it does the wild-type enzyme, but the concentration of ATP required to reach half of the maximal activation is reduced approximately 5-fold for the mutant enzyme. CTP at a concentration of 10 mM does not inhibit the mutant enzyme, while under the same conditions CTP at concentrations less than 1 mM will inhibit the wild-type enzyme to the maximal extent. Higher concentrations of CTP result in some inhibition of the mutant enzyme that may be due either to hetertropic effects at the regulatory site or to competitive binding at the active site. UTP alone or in the presence of CTP has no effect on the mutant enzyme. Kinetic competition experiments indicate that CTP is still able to displace ATP from the regulatory sites of the mutant enzyme. Binding measurements by equilibrium dialysis were used to estimate a lower limit on the dissociation constant for CTP binding to the mutant enzyme (greater than 1 x 10(-3) M). Equilibrium competition binding experiments between ATP and CTP verified that CTP still can bind to the regulatory site of the enzyme. For the mutant enzyme, CTP affinity is reduced approximately 100-fold, while ATP affinity is increased by 5-fold.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Fluorine-19 magnetization transfer experiments have been used to determine the rates of cis/trans isomerization about the X-Pro7 peptide bond in [p-fluoro-Phe8]bradykinin (cis/trans ratio approximately 0.1) and its Gly6 analogue (cis/trans ratio approximately 0.4). The measurements were carried out both prior to and after the addition of cyclophilin, which has recently been shown to have peptidyl-proline cis/trans isomerase activity and is the apparent target enzyme of the immunosuppressive agent cyclosporin A. Magnetization transfer measurements over the temperature range 40-75 degrees C in the absence of enzyme give activation energies of 22.8 and 23.0 kcal/mol for [p-fluoro-Phe8]bradykinin and its Gly6 analogue, respectively. The values for the uncatalyzed cis----trans rate constant, kc, are determined by extrapolation to be 4.8 x 10(-2) and 2.1 x 10(-2) s-1 for the two peptides at 25 degrees C. The enzyme-catalyzed enhancement of the cis/trans interconversion rate was proportional to added cyclophilin concentration and was strongly sequence specific, with bradykinin a much better substrate than [Gly6]bradykinin. At a peptide concentration of 2.2 mM, the catalytic activity expressed as kc per micromolar cyclophilin was determined to be 1.2 s-1/microM for [p-fluoro-Phe8]bradykinin and 0.13 s-1/microM for the Gly6 analogue. The increased cis----trans interconversion rates were strongly inhibited by cyclosporin A and the 6-(methylalanine) derivative, which bind to cyclophilin, but not by the 1-(tetrahydrofurfuryl) derivative of cyclosporin that binds weakly.  相似文献   

11.
Two conformationally restricted analogues of (-)-indolactam-V (1) (cis and trans amides) were examined for their binding selectivity to the synthetic C1 peptides of all protein kinase C (PKC) isozymes. Although the binding constants of the cis amide-restricted analogue (2) were equal to those of 1, the trans amide-restricted analogue (3) bound significantly only to the novel PKC (delta, epsilon, eta, theta) isozymes.  相似文献   

12.
The kinetics and mechanism of the autocatalytic assembly of holo-Cph1 phytochrome (from Synechocystis) from the apoprotein and the bilin chromophores phycocyanobilin (PCB) and phycoerythrobilin (PEB) were investigated by stopped flow and circular dichroism. At 1:1 stoichiometry, pH 7.9, and 10 degrees C, SVD analysis of the kinetic data for PCB revealed three spectral components involving three transitions with time constants tau(1) approximately 150 ms, tau(2) approximately 2.5 s, and tau(3) approximately 50 s. Tau(1) was associated with a major red shift and transfer of oscillator strength from the Soret region to the 680 nm region. When the sulfhydryl group of cysteine 259 was blocked with iodoacetamide, preventing the formation of a covalent adduct, a noncovalent red-shifted complex (680 nm) was formed with a time constant of 200 ms. Tau(1) could thus be assigned to the formation of a noncovalent complex. The absorption changes during tau(1) are due to the formation of the extended conformation of the linear tetrapyrrole and to its protonation in the binding pocket. From the concentration and pH dependence of the kinetics we obtained a value of 1.5 microM for the K(D) of this noncovalent complex and a value of 8.4 for the pK(a) of the proton donor. The tau(2) component was associated with a blue shift of about 25 nm and was attributed to the formation of the covalent bond (P(r)), accompanied with the loss of the 3-3' double bond to ring A. Tau(3) was due to photoconversion to P(fr). For PEB, which is not photochromic, the formation of the noncovalent complex is faster (tau(1) = 70 ms), but the covalent bond formation is about 80 times slower (tau(2) = 200 s) than with the natural chromophore PCB. The CD spectra of the PCB adduct in the 250-800 nm range show that the chromophore geometries in P(r) and P(fr) are similar to those in plant phytochrome. The opposite rotational strengths of P(r) and P(fr) in the longest wavelength band suggest that the photoisomerization induces a reversal of the chirality. The Cph1 complex with noncovalently bound PCB was still photochromic when cysteine 259 was blocked with IAA or with the bulkier IAF. The covalent linkage to cysteine 259 is thus not required for photoconversion. The CD spectra of the noncovalently bound PCB in P(r)- and P(fr)-like states are qualitatively similar to those of the covalent adducts, suggesting analogous structures in the binding pocket. The noncovalent interactions with the binding pocket are apparently sufficient to hold the chromophore in the appropriate geometry for photoisomerization.  相似文献   

13.
The angiotensin I-converting enzyme (peptidyl-dipeptide hydrolase, EC 3.4.15.1) inhibitor, ramiprilat (2-[N-[(S)-1-ethoxycarbonyl-3-phenylpropyl]-L-Ala]-(1S,3S,5S)-2- azabicyclo[3.3.0]octane-3-carboxylic acid), is shown to exist in tow conformational isomers, cis and trans, which interconvert around the amide bond. The two conformers were separated by reversed-phase high-performance liquid chromatography. The conformers were identified by nuclear Overhauser effect measurements. From line shape analysis the isomerization rate constants were determined to be kcis----trans = 15 s-1 and ktrans----cis = 5 s-1 at 368 K in [2H]phosphate buffer (p2H 7.5). By enzyme kinetic studies using 3-(2-furylacryloyl)-L-Phe-Gly-Gly as substrate, the trans conformer was found to be the most potent enzyme inhibitor, whereas the cis conformer had a very low inhibitory effect. A new inhibition mechanism is presented for this type of slow, tight-binding inhibitors that contain an amide bond. This mechanism involves an equilibrium between the two conformers and the enzyme-bound inhibitor complex.  相似文献   

14.
Four 1-ns molecular dynamics computer simulations of tuftsin, Thr-Lys-Pro-Arg, are analyzed: (1) cis tuftsin in water, (2) trans tuftsin in water, (3) cis tuftsin in 1 M NaCl, and (4) trans tuftsin in 1 M NaCl. Independently of the salt concentration, the trans conformer has a higher dielectric constant than the cis conformer because the former exhibits a more widely distributed charge distribution in space. Independently of the peptide conformation, the presence of salt reduces the dielectric constants of both the peptide and the solvating water molecules because ions, on binding, restrict the motion of other atoms. In contrast to the dielectric constants, neither the peptide conformation nor the salt concentration shows a significant influence on the dielectric relaxation time of water molecules.  相似文献   

15.
Preparation and characterization of YADH-bound magnetic nanoparticles   总被引:17,自引:0,他引:17  
The covalently binding of yeast alcohol dehydrogenase (YADH) to magnetic nanoparticles via carbodiimide activation was studied. The magnetic nanoparticles Fe3O4 with a mean diameter of 10.6 nm were prepared by co-precipitating Fe2+ and Fe3+ ions in an ammonia solution and treating under hydrothermal conditions. Transmission electron microscopy (TEM) micrographs showed that the magnetic nanoparticles remained discrete and had no significant change in size after binding YADH. X-ray diffraction (XRD) patterns indicated both the magnetic nanoparticles before and after binding YADH were pure Fe3O4. Magnetic measurement revealed the resultant magnetic nanoparticles were superparamagnetic characteristics, and their saturation magnetization was reduced only slightly after enzyme binding. The analysis of Fourier transform infrared (FTIR) spectroscopy confirmed the binding of YADH to magnetic nanoparticles and suggested a possible binding mechanism. In addition, the measurement of protein content revealed that the maximum weight ratio of YADH bound to magnetic nanoparticles was 0.125, below which the binding efficiency of YADH was almost 100%. The kinetic measurements indicated the bound YADH retained 62% of its original activity and exhibited a 10-fold improved stability than did the free enzyme. The maximum specific activities and Michaelis constants were also determined.  相似文献   

16.
The kinetics of the hydrodynamic volume change accompanying the reversible unfolding of staphylococcal nuclease have been observed by size-exclusion chromatography at 4 degrees C and pH 7.0 using the denaturant guanidine hydrochloride. The observed chromatographic profiles have been simulated by a six-component unfolding/refolding mechanism using a consistent set of equilibrium and kinetic parameters. The native protein is an equilibrium mixture of the cis and trans isomers of the peptide bond preceding proline-117. The native conformation containing the cis isomer dominates the equilibrium mixture, is more stable, and unfolds more slowly at its transition midpoint. The denatured protein is an equilibrium mixture of at least four components, the cis/trans isomers of proline-117 and one of the five remaining prolines. The dominant refolding pathway is initiated from the denatured component containing the trans isomer of proline-117. The six-component mechanism is consistent with tryptophan fluorescence kinetic measurements of the wild-type protein and with chromatographic measurements of a mutant P117G protein.  相似文献   

17.
The turnover numbers and other kinetic constants for human alcohol dehydrogenase (ADH) 4 ("stomach" isoenzyme) are substantially larger (10-100-fold) than those for human class I and horse liver alcohol dehydrogenases. Comparison of the primary amino acid sequences (69% identity) and tertiary structures of these enzymes led to the suggestion that residue 317, which makes a hydrogen bond with the nicotinamide amide nitrogen of the coenzyme, may account for these differences. Ala-317 in the class I enzymes is substituted with Cys in human ADH4, and locally different conformations of the peptide backbones could affect coenzyme binding. This hypothesis was tested by making the A317C substitution in horse liver ADH1E and comparisons to the wild-type ADH1E. The steady-state kinetic constants for the oxidation of benzyl alcohol and the reduction of benzaldehyde catalyzed by the A317C enzyme were very similar (up to about 2-fold differences) to those for the wild-type enzyme. Transient kinetics showed that the rate constants for binding of NAD(+) and NADH were also similar. Transient reaction data were fitted to the full Ordered Bi Bi mechanism and showed that the rate constants for hydride transfer decreased by about 2.8-fold with the A317C substitution. The structure of A317C ADH1E complexed with NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol at 1.2 ? resolution is essentially identical to the structure of the wild-type enzyme, except near residue 317 where the additional sulfhydryl group displaces a water molecule that is present in the wild-type enzyme. ADH is adaptable and can tolerate internal substitutions, but the protein dynamics apparently are affected, as reflected in rates of hydride transfer. The A317C substitution is not solely responsible for the larger kinetic constants in human ADH4; thus, the differences in catalytic activity must arise from one or more of the other hundred substitutions in the enzyme.  相似文献   

18.
The cloned genes encoding carboxylesterase E3 in the blowfly Lucilia cuprina and its orthologue in Drosophila melanogaster were expressed in Sf9 cells transfected with recombinant baculovirus. Resistance of L. cuprina to organophosphorus insecticides is due to mutations in the E3 gene that enhance the enzyme's ability to hydrolyse insecticides. Previous in vitro mutagenesis and expression of these modifications (G137D, in the oxyanion hole and W251L, in the acyl pocket) have confirmed their functional significance. We have systematically substituted these and nearby amino acids by others expected to affect the hydrolysis of pyrethroid insecticides. Most mutations of G137 markedly decreased pyrethroid hydrolysis. W251L was the most effective of five substitutions at this position. It increased activity with trans permethrin 10-fold, and the more insecticidal cis permethrin >130-fold, thereby decreasing the trans:cis hydrolysis ratio to only 2, compared with >25 in the wild-type enzyme. Other mutations near the bottom of the catalytic cleft generally enhanced pyrethroid hydrolysis, the most effective being F309L, also in the presumptive acyl binding pocket, which enhanced trans permethrin hydrolysis even more than W251L. In these assays with racemic 1RS cis and 1RS trans permethrin, two phases were apparent, one being much faster suggesting preferential hydrolysis of one enantiomer in each pair as found previously with other esterases. Complementary assays with individual enantiomers of deltamethrin and the dibromo analogue of cis permethrin showed that the wild type and most mutants showed a marked preference for the least insecticidal 1S configuration, but this was reversed by the F309L substitution. The W251L/F309L double mutant was best overall in hydrolysing the most insecticidal 1R cis isomers. The results are discussed in relation to likely steric effects on enzyme-substrate interactions, cross-resistance between pyrethroids and malathion, and the potential for bioremediation of pyrethroid residues.  相似文献   

19.
The separate interaction of the substrate fructose 1,6-bisphosphate and a metal ion cofactor Mn2+ with neutral hexosebisphosphatase has been studied under equilibrium conditions at pH 7.5 with gel filtration and electron paramagnetic resonance measurements, respectively. Binding data for both ligands to the enzyme yielded nonlinear Scatchard plots that analyze in terms of four negatively cooperative binding sites per enzyme tetramer. Graphical estimates of the binding constants were refined by a computer searching procedure and nonlinear least squares analysis. These results are qualitatively similar to those obtained from binding studies involving teh alkaline enzyme, a modified form of hexosebisphosphatase whose pH optimum is in the alkaline pH region. Both forms of the enzyme enhance the proton relaxation rate of water protons by a factor of approximately 7 to 8 at 24 MHz, demonstrating similar metal ion environments. Teh activator Co(III)-EDTA did not affect Mn2+ binding to the neutral enzyme. In the presence of (alpha + beta)methyl-D-fructofuranoside 1,6-bisphosphate, however, two sets--each containing four Mn2+ binding sites--were observed per enzyme tetramer with loss of the negatively cooperative interaction. These results are viewed in terms of four noncatalytic and four catalytic Mn2+ binding sites. Parallel kinetic investigations were conducted on the neutral enzyme to determine specific activity as a function of Mn2+ and fructose 1,6-bisphosphate concentration. A pro-equilibrium sequential pathway model involving Mn2+-enzyme and the Mn2+-fructose 1,6-bisphosphate complex both as substrate and as an allosteric inhibitor satisfactorily fit the kinetic observations. All possible enzyme species were computed from the determined binding constants and grouped according to the number of moles of Mn2+-fructose 1,6-bisphosphate complex bound to the Mn2+-enzyme, and individual rate constants were calculated. The testing of other models and their failure to describe the kinetic observations are discussed.  相似文献   

20.
The phosphorylation-specific peptidyl-prolyl isomerase Pin1 catalyzes the isomerization of the peptide bond preceding a proline residue between cis and trans isomers. To best understand the mechanisms of Pin1 regulation, rigorous enzymatic assays of isomerization are required. However, most measures of isomerase activity require significant constraints on substrate sequence and only yield rate constants for the cis isomer, [Formula: see text] and apparent Michaelis constants, [Formula: see text]. By contrast, NMR lineshape analysis is a powerful tool for determining microscopic rates and populations of each state in a complex binding scheme. The isolated catalytic domain of Pin1 was employed as a first step towards elucidating the reaction scheme of the full-length enzyme. A 24-residue phosphopeptide derived from the amyloid precurser protein intracellular domain (AICD) phosphorylated at Thr668 served as a biologically-relevant Pin1 substrate. Specific (13)C labeling at the Pin1-targeted proline residue provided multiple reporters sensitive to individual isomer binding and on-enzyme catalysis. We have performed titration experiments and employed lineshape analysis of phosphopeptide (13)C-(1)H constant time HSQC spectra to determine [Formula: see text], [Formula: see text], [Formula: see text], and [Formula: see text] for the catalytic domain of Pin1 acting on this AICD substrate. The on-enzyme equilibrium value of [E·trans]/[E·cis]?=?3.9 suggests that the catalytic domain of Pin1 is optimized to operate on this substrate near equilibrium in the cellular context. This highlights the power of lineshape analysis for determining the microscopic parameters of enzyme catalysis, and demonstrates the feasibility of future studies of Pin1-PPIase mutants to gain insights on the catalytic mechanism of this important enzyme.  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