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.
Reaction of α-chymotrypsin with o-[35S]hydroxy-α-toluenesulfonic acid sultone (I) at pH 7 produces an intermediate which substantially decomposes over a period of 24 hr at pH 3 as determined by gel chromatography. Isolation of the intermediate at pH 3 followed by denaturation in urea, reduction of disulfide bonds, and SDS-urea gel electrophoresis resulted in separation of the three peptides which compose the enzyme. Radioactivity (35S) was associated with the peptide which contains serine 195 and not with the peptide containing histidine 57. These results indicate that a covalent linkage is established between sultone I and α-chymotrypsin, presumably at serine 195.  相似文献   

2.
M Rüegg  V Metzger  H Susi 《Biopolymers》1975,14(7):1465-1471
Infrared spectra of myoglobin, ribonuclease, lysozyme, α-chymotrypsin, α-lactalbumin, and β-lactoglobulin A were obtained in deuterium oxide solution in units of absorbance versus wavenumber from 1340 to 1750 cm?1. The spectra were resolved into Gaussian components by means of an iterative computer program. Resolved characteristic absorption peaks for the two infrared active amide I′ components of antiparallel chain-pleated sheets (β-structure) were obtained. The characteristic amide I′ peaks of α-helical regions and apparently unordered regions overlap in D2O solution. Absorptivity values for the resolved β-structure peak around 1630 cm?1 were estimated on the basis of the known structure of ribonuclease, lysozyme, and β-chymotrypsin. The β-structure content of β-lactoglobulin was estimated to be ca. 48% of α-lactalbumin ca. 18%, and of αs-casein close to zero. The results are in general agreement with conclusions drawn from circular dichroism and optical rotatory dispersion studies.  相似文献   

3.
Two new inhibitors in which the terminal α-carboxyl groups of Z-Ala-Ala-Phe-COOH and Z-Ala-Pro-Phe-COOH have been replaced with a proton to give Z-Ala-Ala-Phe-H and Z-Ala-Pro-Phe-H, respectively, have been synthesized. Using these inhibitors, we estimate that for α-chymotrypsin and subtilisin Carlsberg the terminal carboxylate group decreases the level of inhibitor binding 3-4-fold while a glyoxal group increases the level of binding by 500-2000-fold. We show that at pH 7.2 the effective molarities of the catalytic hydroxyl group of the active site serine are 41000-229000 and 101000-159000 for α-chymotrypsin and subtilisin Carlsberg, respectively. It is estimated that oxyanion stabilization and the increased effective molarity of the catalytic serine hydroxyl group can account for the catalytic efficiency of the reaction. We argue that substrate binding induces the formation of a strong hydrogen bond or low-barrier hydrogen bond between histidine-57 and aspartate-102 that increases the pK(a) of the active site histidine, allowing it to be an effective general base catalyst for the formation of the tetrahedral intermediate and increasing the effective molarity of the catalytic hydroxyl group of serine-195. A catalytic mechanism for acyl intermediate formation in the serine proteases is proposed.  相似文献   

4.
The effect of Gd3+ on the nuclear magnetic resonance (nmr) relaxation rates, T1m?1 and T2m?1, of inhibitor protons in metal-inhibitor-α-chymotrypsin ternary complexes has been measured. The Solomon-Bloembergen equations were used to calculate the distance from the methyl protons of p-toluamidine (a competitive inhibitor) to the Gd3+ binding site which is 9.2 ± 0.5 Å. Calcium ion and gadolinium ion compete for the same binding site on α-chymotrypsin. Distances from the specificity pocket of α-chymotrypsin to the metal binding site have been measured by fluorescence energy transfer experiments. By observing energy transfer between proflavine and Nd3+, Pr3+, or Ho3+, we have been able to calculate a distance of approximately 10 Å between the two chromophores. This agrees well with the data obtained by nmr techniques and also gives nearly identical values to those obtained for trypsin (Darnall, D., Abbott, F., Gomez, J. E., and Birnbaum, E. R., Biochemistry15, 5017, 1976). This is consistent with the calcium ion binding sites being composed of the same residues in both trypsin and α-chymotrypsin.  相似文献   

5.
Single-channel and [3H]ryanodine binding experiments were carried out to examine the effects of imperatoxin activator (IpTxa), a 33 amino acid peptide isolated from the venom of the African scorpion Pandinus imperator, on rabbit skeletal and canine cardiac muscle Ca2+ release channels (CRCs). Single channel currents from purified CRCs incorporated into planar lipid bilayers were recorded in 250 mM KCl media. Addition of IpTxa in nanomolar concentration to the cytosolic (cis) side, but not to the lumenal (trans) side, induced substates in both ryanodine receptor isoforms. The substates displayed a slightly rectifying current–voltage relationship. The chord conductance at −40 mV was ∼43% of the full conductance, whereas it was ∼28% at a holding potential of +40 mV. The substate formation by IpTxa was voltage and concentration dependent. Analysis of voltage and concentration dependence and kinetics of substate formation suggested that IpTxa reversibly binds to the CRC at a single site in the voltage drop across the channel. The rate constant for IpTxa binding to the skeletal muscle CRC increased e-fold per +53 mV and the rate constant of dissociation decreased e-fold per +25 mV applied holding potential. The effective valence of the reaction leading to the substate was ∼1.5. The IpTxa binding site was calculated to be located at ∼23% of the voltage drop from the cytosolic side. IpTxa induced substates in the ryanodine-modified skeletal CRC and increased or reduced [3H]ryanodine binding to sarcoplasmic reticulum vesicles depending on the level of channel activation. These results suggest that IpTxa induces subconductance states in skeletal and cardiac muscle Ca2+ release channels by binding to a single, cytosolically accessible site different from the ryanodine binding site.  相似文献   

6.
In general, biological macromolecules require significant dynamical freedom to carry out their different functions, including signal transduction, metabolism, catalysis and gene regulation. Effectors (ligands, DNA and external milieu, etc) are considered to function in a purely dynamical manner by selectively stabilizing a specific dynamical state, thereby regulating biological function. In particular, proteins in presence of these effectors can exist in several dynamical states with distinct binding or enzymatic activity. Here, we have reviewed the efficacy of ultrafast fluorescence spectroscopy to monitor the dynamical flexibility of various proteins in presence of different effectors leading to their biological activity. Recent studies demonstrate the potency of a combined approach involving picosecond-resolved Förster resonance energy transfer, polarisation-gated fluorescence and time-dependent stokes shift for the exploration of ultrafast dynamics in biomolecular recognition of various protein molecules. The allosteric protein–protein recognition following differential protein–DNA interaction is shown to be a consequence of some ultrafast segmental motions at the C-terminal of Gal repressor protein dimer with DNA operator sequences OE and OI. Differential ultrafast dynamics at the C-terminal of λ-repressor protein with two different operator DNA sequences for the protein–protein interaction with different strengths is also reviewed. We have also systemically briefed the study on the role of ultrafast dynamics of water molecules on the functionality of enzyme proteins α-chymotrypsin and deoxyribonuclease I. The studies on the essential ultrafast dynamics at the active site of the enzyme α-chymotrypsin by using an anthraniloyl fluorescent extrinsic probe covalently attached to the serine-195 residue for the enzymatic activity at homeothermic condition has also been reviewed. Finally, we have highlighted the evidence that a photoinduced dynamical event dictates the molecular recognition of a photochromic ligand, dihydroindolizine with the serine protease α-chymotrypsin and with a liposome (L-α-phosphatidylcholine).  相似文献   

7.
Phenomena occurring in the heme pocket after photolysis of carbonmonoxymyoglobin (MbCO) below about 100 K are investigated using temperature-derivative spectroscopy of the infrared absorption bands of CO. MbCO exists in three conformations (A substrates) that are distinguished by the stretch bands of the bound CO. We establish connections among the A substates and the substates of the photoproduct (B substates) using Fourier-transform infrared spectroscopy together with kinetic experiments on MbCO solution samples at different pH and on orthorhombic crystals. There is no one-to-one mapping between the A and B substates; in some cases, more than one B substate corresponds to a particular A substate. Rebinding is not simply a reversal of dissociation; transitions between B substates occur before rebinding. We measure the nonequilibrium populations of the B substates after photolysis below 25 K and determine the kinetics of B substate transitions leading to equilibrium. Transitions between B substates occur even at 4 K, whereas those between A substates have only been observed above about 160 K. The transitions between the B substates are nonexponential in time, providing evidence for a distribution of substates. The temperature dependence of the B substate transitions implies that they occur mainly by quantum-mechanical tunneling below 10 K. Taken together, the observations suggest that the transitions between the B substates within the same A substate reflect motions of the CO in the heme pocket and not conformational changes. Geminate rebinding of CO to Mb, monitored in the Soret band, depends on pH. Observation of geminate rebinding to the A substates in the infrared indicates that the pH dependence results from a population shift among the substates and not from a change of the rebinding to an individual A substate.  相似文献   

8.
A plastein was synthesized with α-chymotrypsin from a dialyzable fraction of a peptic hydrolysate of soybean protein.

The plastein was obtainable also by use of an insoluble preparation of α-chymotrypsin. This may rule out the possibility that the plastein is a product resulting from some chemical peptide-protein (enzyme) aggregation.

No appreciable amount of the plastein was produced when chymotrypsinogen was used instead of α-chymotrypsin.

The plastein synthetic, as well as the protein hydrolytic, activity of α-chymotrypsin was inhibited more or less by a hydrophobic inhibitor (n-hexane), a competitive inhibitor (benzolyl-d,l-phenylalanine), and divalent cations (Zn2+, Hg2+ and Cu2+); the degree of inhibition in each case was approximately similar against both the synthetic and the hydrolytic activities.

Either diisopropylphosphorylation of the β-O of Ser-195 or methylation of the 3-N of His-57 imidazole of α-chymotrypsin repressed the synthetic, as well as the hydrolytic, activity.

Based on these results a possible mechanism was discussed of the plastein synthesis by α-chymotrypsin, especially in relevance to its acylation and deacylation.  相似文献   

9.
The large enhancement of the green luminescence of terbium ion which occurs on binding to porcine and bovine trypsins and to bovine α-chymotrypsin has been used to study the calcium binding sites of these enzymes. Excitation spectra, taken at low protein concentrations to minimize absorption effects, demonstrate that in each case, energy transfer occurs between the side chain of a tryptophan residue and bound Tb3+. Association constants for the binding of Tb3+ to the single binding site on each of the three proteins have been measured at 25 °C and pH 6.6. Ca2+ ions compete with Tb3+ for the single binding site, and association constants for Ca2+ were determined by Tb3+ displacement. The ratio of binding strengths of Ca2+ to α-chymotrypsin, bovine trypsin, porcine trypsin, and elastase is 1:12:24:23. Addition of Tb3+ to the homologous bacterial enzyme α-lytic protease caused no luminescence enhancement.  相似文献   

10.
MutT distinguishes substrate 8-oxo-dGTP from dGTP and also 8-oxo-dGMP from dGMP despite small differences of chemical structures between them. In this paper we show by the method of molecular dynamics simulation that the transition between conformational substates of MutT is a key mechanism for a high-resolution molecular recognition of the differences between the very similar chemical compounds. (1) The native state MutT has two conformational substates with similar free energies, each characterized by either open or closed of two loops surrounding the substrate binding active site. Between the two substates, the open substate is more stable in free MutT and in dGMP-MutT complex, and the closed substate is more stable in 8-oxo-dGMP-MutT complex. (2) Conformational fluctuation of the open substate is much larger than that of the closed substate. An estimate of associated entropy difference was found to be consistent with the experimentally found difference of entropy contribution to the binding free energies of the two molecules. (3) A hydrogen bond between H7 atom of 8-oxo-dGMP and the sidechain of Asn119 plays a crucial role for maintaining the closed substate in 8-oxo-dGMP-MutT complex. When this hydrogen bond is absent in the H7-deficient dGMP-MutT complex, the closed substate is no more maintained and transition to the more entropically-favored open substate is induced. (4) Thus, this mechanism of the hydrogen bond controlling the relative stabilities of the drastically different two conformational substates enhances the resolution to recognize the small difference of the chemical structures between the two molecules, dGMP and 8-oxo-dGMP.  相似文献   

11.
The crystal structure of γ-chymotrypsin, the monomeric form of chymotrypsin, has been determined and refined to a crystallographic R-factor of 0.18 at 1.9 Å resolution. The details of the catalytic triad involving Asp102, His57 and Ser195 agree well with the results found for trypsin (Chambers & Stroud, 1979) and Streptomyces griseus protease A (Sielecki et al., 1979). As in many of the other serine proteases, the Oγ of Ser195 does not appear to be hydrogen-bonded to His57.The three-dimensional structures of γ- and α-chymotrypsin (Birktoft & Blow, 1972) are closely similar. The largest backbone differences occur in the “calcium binding loop” (residues 75 to 78) and in the “autolysis loop” (residues 146, 149 and 150). Ala149 and Asn150 are disordered in γ-chymotrypsin, whereas they are stabilized by intermolecular interactions in α-chymotrypsin. The conformation of Ser218 is also different, presumably the indirect result of the dimeric interactions of α-chymotrypsin. These results are discussed in terms of the slow, pH-dependent interconversion of α- and γ-chymotrypsin.  相似文献   

12.
The interpretation of the CD spectra of proteins to date requires additional secondary structural information of the proteins to be analyzed, such as x-ray or nmr data. Therefore, these methods are inappropriate for a CD data base whose secondary structures are unknown, as in the case of the membrane proteins. The Convex Constraint Analysis algorithm [A. Perczel, M. Hollósi, G. Tusnády, and G. D. Fasman (1991) Protein Engineering, Vol. 4, 669–679], on the other hand, operates only on a collection of spectral data to extract the common spectral components with their spectral weights. The linear combinations of these derived “pure” CD curves can reconstruct the original data set with great accuracy. For a membrane protein data set, the five-component spectra so obtained from the deconvolution consisted of two different types of α-helices (the α-helix in the soluble domain and the αT-helix, for the transmembrane α-helix), a β-pleated sheet, a class C-like spectrum related to β-turns, and a spectrum correlated with the unordered conformation. The deconvoluted CD spectrum for the αT-helix was characterized by a positive red-shifted band in the range 195–200 nm (+95,000 deg cm2 dmol?l), with the intensity of the negative band at 208 nm being slightly less negative than that of the 222 nm band (?50,000 and ?60,000 deg cm2 dmol?1, respectively) in comparison with the regular α-helix, with a positive band at 190 nm and two negative bands at 208 and 222 nm with magnitudes of + 70,000, ?30,000, and ?30,000 deg cm2 dmol?1, respectively. © 1994 John Wiley & Sons, Inc.  相似文献   

13.
Insertion of nickel ions into the empty catalytic site of horse liver alcohol dehydrogenase yields an active enzyme with 65% metal substitution and about 12% intrinsic activity. The electronic absorption spectrum is characterized by bands at 357 nm (2900 M?1 cm?1, 407 nm (3500 M?1 cm?1), 505 nm (300 M?1 cm?1), 570 nm (?130 M?1 cm?1), and 680 nm (?80 M?1 cm?1). The absorption and CD spectra are similar to those of nickel(II) azurin and nickel(II) aspartate transcarbamoylase and prove coordination of the nickel(II) ions to sulfur in a distorted tetrahedral coordination geometry. Changes of the spectra upon ligand binding at the metal or conformation changes of the protein induced by coenzyme, or both, indicate alterations of the metal geometry.The chromophoric substrate trans-4-(N, N-dimethylamino)-cinnamaldehyde forms a ternary complex with Ni(II) liver alcohol dehydrogenase and the coenzyme analogue 1,4,5,6-tetrahydronicotinamide-adenine-dinucleotide, stable between pH 6 and 10. The corresponding ternary complex with NADH is only stable at pH > 9.0. The spectral redshifts induced in the substrate are 11 nm larger than those found in the zinc enzyme. We suggest direct coordination of the substrate to the catalytic metal ion which acts as a Lewis acid in both substrate coordination and catalysis.  相似文献   

14.
The interaction of the inhibitor VJ (InhVJ), isolated from sea anemone R. macrodactylus, with different proteases was investigated using the method of biosensor analysis. The following enzymes were tested: serine proteases (trypsin, α-chymotrypsin, plasmin, thrombin, kallikrein), cysteina protease (papain) and aspartic protease (pepsin). In the rage of the concentrations studied (10–400 nM) inhibitor VJ interacted only with trypsin and α-chymotrypsin. The intermolecular complexes formation between inhibitor VJ and each of these enzymes was characterized by the following kinetic and thermodynamics parameters: KD = 7.38 × 10?8 M and 9.93 × 10?7 M for pairs InhVJ/trypsin and InhVJ/α-chymotrypsin, respectively.  相似文献   

15.
The active center of human leukocyte cathepsin G, human pancreatic chymotrypsin II, and bovine α-chymotrypsin has been investigated with a series of substrates of general formula succinyl-(l-alanine)n-phenylalanine-p-nitroanilide (n = 0 to 3). The three proteinases have an extended substrate binding site which includes at least six subsites. Secondary interactions are very important for their catalytic power since the longest substrate is hydrolyzed 600 to 1100 times faster than the shortest one. The regulatory subsite is S4 for bovine α-chymotrypsin and human cathepsin G whereas it is S5 for human chymotrypsin II. Cathepsin G is a poor catalyst compared to the two other enzymes.  相似文献   

16.
An acid-stable and heat-labile proteinous protease inhibitor which was found in spinach leaves but not in seeds was isolated by sequential chromatography and preparative isoelectric focusing. The isoelectric point of this inhibitor was 4.5. The inhibitor had a Mr of ca 18 000 and was rich in aspartic acid and glycine; it had 4 half-cystine, 2 tryptophan and no methionine residues. Its extinction coefficient (E|cm%) was 13.7 at 280 nm. The inhibition was competitive and the dissociation constant was 3.32 × 10?13 M. The inhibitor was specific to serine proteases and strongly inhibited trypsin and weakly inhibited α-chymotrypsin and kallikrein.  相似文献   

17.
The motions of water molecules, the acyl moiety, the catalytic triad, and the oxyanion binding site of acyl-chymotrypsin were studied by means of a stochastic boundary molecular dynamics simulation. A water molecule that could provide the nucleophilic OH? for the deacylation stage of the catalysis was found to be trapped between the imidazole ring of His-57 and the carbonyl carbon of the acyl group. It makes a hydrogen bond with the Nε2 of His-57 and is heldin place through a network of hydrogen-bonded water molecules in theactive site. The water molecule was found as close as 2.8 Å to the carbonyl carbon. This appears to be due to the constraints imposed by nonbonded interaction in the active site. Configurations were found in which one hydrogen of the trapped water shared a bifurcated hydrogen bond with His-57-Nε2 and Ser-195-0γ with the water oxygen very close to the carbonyl carbon. The existence of such a water molecule suggests that large movement of the His-57 imidazole ring between positions suitable for providing general-base catalyzed assistance and for providing general-acid catalyzed assistance may notbe required during the reaction. The simulation indicates that the side chains of residues involved in catalysis (i.e., His-57, Ser-195, and Asp-102) are significantly less flexible than other side chains in the protein. The 40% reduction in rms fluctuations is consistent with a comparable reduction calculated from the temperature factors obtained in the X-ray crystal-lographic data of γ-chymotrypsin. The greater rigidity of active site residues seems to result from interconnected hydrogen bonding networks among the residues and between the residues and the solvent water in the active site. © Wiley-Liss, Inc.  相似文献   

18.
The low pH conformational transitions of a series of modified chymotrypsins have been examined and compared to the association properties of these same proteins in order to determine if the effect of the modification was directly upon the association or indirectly through conformational changes. The modifications of α-chymotrypsin that have been studied are serine-195 modified with acetyl or with phenylmethane sulfonyl groups, His-57 modified by tosylamidophenylethyl chloromethyl ketone or with a methyl group, and Met-192 modified with a nitrophenacyl group. Chymotrypsinogen and native and modified delta-chymotrypsin have also been studied. Ultraviolet difference spectra and optical rotatory dispersion measurements show that the various proteins may be divided into three groups depending upon their response in the low pH transition. These groupings based upon conformational transitions are not divided in the same manner as were those based upon the effect of the modifications on the association process itself. The latter process had earlier been found to correlate with the size of the modification in the active site. From these findings it has been concluded that the effect of the modifications on the association was directly upon the interacting subunit interface and that the observed conformational change is only incidental to this effect. Therefore, the amino acids involved in the association in solution are the same as those involved in the dimerization in the crystal, namely, the active site region of Ser-195, His-57, and Met-192. This is the most direct demonstration that the mode of association of a protein in solution is the same as the mode of association in the crystal.  相似文献   

19.
Addition of KCN to Helix pomatia β-hemocyanin fully saturated with either O2 or CO results in a decrease of the spectroscopic properties of the protein (absorbance at 340 nm and luminescence at 550 nm) due to the displacement of the gaseous ligands (O2 or CO) from the active site. The anionic form of cyanide (CN?) is supposed to bind to the active site; its intrinsic affinity for the protein, as calculated from independent O2 and CO displacement experiments, is between 2 and 6 × 106M?1. The replacement of O2 or CO shows some differences which may be correlated with the different modes of binding at the active site. Thus, while displacement of oxygen by cyanide is hyperbolic, addition of cyanide to carbonylated hemocyanin shows a lag phase. This finding suggests the formation of a mixed liganded complex at the active site. The simultaneous presence of CO and CN? at the active site of hemocyanin is also supported by the experiment in which addition of small amounts of KCN to hemocyanin partially saturated with O2 and CO gives rise to an increase of emission intensity and a concomitant decrease of the O2 absorption band. The mixed-liganded species displays luminescence properties similar to those of CO-saturated hemocyanin, and the formation of the complex is reversible on dialysis or oxygenation.  相似文献   

20.
A. F. Bunkin 《Biophysics》2012,57(6):709-715
The novel method of nonlinear laser spectroscopy — low frequency spectroscopy of four-photon scattering of laser radiation was applied to detect a considerable growth of ortho-H2O spin isomer and also H2O2 molecule concentration in a hydrate layer at the interface between water and DNA, denatured DNA molecules and α-chymotrypsin. Spectra of rotational resonances of ortho/para-H2O spin isomers were observed in aqueous solutions of different biopolymers and also in distilled water in the range from zero to 100 cm?1 with the spectral resolution of 0.05–0.1 cm?1. The fitting of four-wave mixing spectra shows notable growth of the H2O2 concentration and rotational line’s amplitude by a factor of ~3 in DNA solutions due to denaturizing. Besides, we studied the four-photon scattering spectra of α-chymotrypsin aqueous solutions at protein concentrations between 0 and 20 mg/cm3 in the range of ±7 cm?1. We found that the velocity of sound in the protein aqueous solution measured by the shift of the Mandelstam-Brillouin scattering spectrum components was a cubic dependence on the protein concentration and reached the value of about 3000 m/s at 20 mg/cm3.  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