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.
3-Alkylpyridinium polymers (poly-APS), composed of 29 or 99 N-butyl-3-butyl pyridinium units, were isolated from the marine sponge Reniera sarai. They act as potent cholinesterase inhibitors. The inhibition kinetics pattern reveals several successive phases ending in irreversible inhibition of the enzyme. To provide more information on mechanism of inhibition, interaction of poly-APS and N-butyl-3-butyl pyridinium iodide (NBPI) with soluble dimeric and monomeric insect acetylcholinesterase (AChE) was studied by using enzyme intrinsic fluorescence and light scattering, conformational probes ANS and trypsin, and SDS–PAGE. Poly-APS quenched tryptophan fluorescence emission of AChE more extensively than NBPI. Both inhibitors exhibited a pseudo-Lehrer type of quenching. Interaction of poly-APS with dimeric AChE did not induce significant changes of the enzyme conformation as assayed by using the hydrophobic probe ANS and trypsin digestion. In contrast to NBPI, titration of both monomeric and dimeric AChE with poly-APS resulted in the appearance of large complexes detected by measuring light scattering. An excess of poly-APS produced AChE precipitation as proved on SDS–PAGE. None of the effects were observed with trypsin as a control. It was concluded that AChE aggregation and precipitation rather than the enzyme conformational changes accounted for the observed irreversible component of poly-APS inhibition.  相似文献   

2.
Lipocalin 11 (Lcn11), a recently identified member of the lipocalin family, potentially plays crucial physiological roles in male reproduction. In this present work, we cloned, expressed and purified the rat Lcn11 (rLcn11) protein in Escherichia coli. A C59A/C156A substitution was introduced to ameliorate the misfolding and aggregation problem associated with the wild-type protein. From circular dichroism and non-reducing SDS–PAGE, we characterized the conformational properties of rLcn11 as a typical lipocalin scaffold with the conserved disulfide bridge. The results obtained from size-exclusion chromatography, cross-linking experiment and dynamic light scattering analysis indicate that the recombinant rLcn11 protein forms dimer in neutral solution. By using fluorescent probe 8-anilino-1-naphtahlene sulfonic acid (ANS), we found rLcn11 might contain multiple hydrophobic binding sites for ligand binding. Similarly to the odorant-binding protein, rLcn11 processes a moderate affinity for binding 1-aminoanthracene (AMA), implying that Lcn11 might work as a dimeric chemoreception protein in male reproductive system.  相似文献   

3.
It is assumed that amino acid sequence differences in highly homologous enzymes would be found at the peripheral level, subtle changes that would not necessarily affect catalysis. Here, we demonstrate that, using the same set of mutations at the level of the interface loop 3, the activity of a triosephosphate isomerase monomeric enzyme is ten times higher than that of a homologous enzyme with 74% identity and 86% similarity, whereas the activity of the native, dimeric enzymes is essentially the same. This is an example of how the dimeric biological unit evolved to compensate for the intrinsic differences found at the monomeric species level. Biophysical techniques of size exclusion chromatography, dynamic light scattering, X-ray crystallography, fluorescence and circular dichroism, as well as denaturation/renaturation assays with guanidinium hydrochloride and ANS binding, allowed us to fully characterize the properties of the new monomer.  相似文献   

4.
A trypsin fraction was isolated from the pyloric ceca of New Zealand farmed chinook salmon (Oncorhynchus tshawytscha) by ammonium sulfate fractionation, acetone precipitation and affinity chromatography. The chinook salmon enzyme hydrolyzed the trypsin-specific synthetic substrate benzoyl-dl-arginine-p-nitroanilide (dl-BAPNA), and was inhibited by the general serine protease inhibitor phenyl methyl sulfonyl fluoride (PMSF), and also by the specific trypsin inhibitors — soybean trypsin inhibitor (SBTI) and benzamidine. The enzyme was active over a broad pH range (from 7.5 to at least pH 10.0) at 25 °C and was stable from pH 4.0 to pH 10.0 when incubated at 20 °C, with a maximum at pH 8.0. The optimum temperature for the hydrolysis of dl-BAPNA by the chinook salmon enzyme was 60 °C, however, the enzyme was unstable at temperatures above 40 °C. The molecular mass of the chinook salmon trypsin was estimated as 28 kDa by SDS–PAGE.  相似文献   

5.
O-Ethyl S-[2-(diisopropylamino)ethyl] methylphosphonothioate (MPT) is an active site directed inhibitor of acetylcholinesterase (AChE). Inhibition of the Electrophorus electricus (G4) enzyme follows classical second-order kinetics. However, inhibition of total mouse skeletal muscle AChE and inhibition of the individual molecular forms from muscle, including the monomeric species, do not proceed as simple irreversible bimolecular reactions. Similarly, complex inhibition kinetics are observed for the purified enzyme from Torpedo californica. AChE can be cross-linked with glutaraldehyde into a semisolid matrix. Under these conditions the abnormal concentration dependence for MPT inhibition is accentuated, and a range of MPT concentrations can be found where inhibition of polymerized AChE is far less than that observed at lower concentrations. Inhibition in certain concentration ranges is partially reversible after removal of all unbound ligand. Thus, there are two different modes of organophosphorus inhibition by MPT: the classical irreversible phosphorylation of the active site and a reversible interaction at a site peripheral to the active center. Propidium, a well-studied peripheral site ligand, can prevent the later interaction. Hence, the second site of MPT interaction with AChE may overlap or be linked to the peripheral anionic site of AChE characterized by the binding of propidium and other peripheral site inhibitors.  相似文献   

6.
Previous studies on the conformation of the monomeric acetylcholinesterase (AChE) from the krait (Bungarus fasciatus) venom showed that the protein possesses a large permanent dipole moment. These studies predicted that thermal irreversible denaturation must occur via partially unfolded states. The thermal stability of Bungarus AChE was determined using capillary electrophoresis (CE) with optimized conditions. Runs performed at convenient temperature scanning rates provided evidence for an irreversible denaturation process according to the Lumry and Eyring model. The mid-transition temperature, T(m), and the effective enthalpy change, DeltaH(m) were determined at different pH. The temperature dependence of the free energy, DeltaG, of Bungarus AChE unfolding was drawn using values of T(m), DeltaH(m) and DeltaC(p) determined by CE. The thermodynamic parameters for the thermal denaturation of the monomeric snake enzyme were compared with those of different dimeric and tetrameric ChEs. It was shown that the changes in the ratio of DeltaH(cal/)DeltaH(vH) and DeltaC(p) reflect the oligomerization state of these proteins. All these results indicate that wild-type monomeric Bungarus AChE is a stable enzyme under standard conditions. However, designed mutants of this enzyme capable of degrading organophosphates have to be engineered to enhance their thermostability.  相似文献   

7.
Three structurally distinct forms of phosphoglycerate mutase from the trypanosomatid parasite Leishmania mexicana were isolated by standard procedures of bacterial expression and purification. Analytical size-exclusion chromatography coupled to a multi-angle scattering detector detected two monomeric forms of differing hydrodynamic radii, as well as a dimeric form. Structural comparisons of holoenzyme and apoenzyme trypanosomatid cofactor-independent phosphoglycerate mutase (iPGAM) X-ray crystal structures show a large conformational change between the open (apoenzyme) and closed (holoenzyme) forms accounting for the different monomer hydrodynamic radii. Until now iPGAM from trypanosomatids was considered to be only monomeric, but results presented here show the appearance of a dimeric form. Taken together, these observations are important for the choice of screening strategies to identify inhibitors of iPGAM for parasite chemotherapy and highlight the need to select the most biologically or functionally relevant form of the purified enzyme.  相似文献   

8.
1-anilino-8-naphtalenesulfonate (ANS) is a hydrophobic dipole previously used to demonstrate that the proton for potassium exchange by the gastric HK-ATPase is electroneutral. In this paper, we demonstrate that ANS binds to gastric membranes and probes conformational changes of the HK-ATPase independently of any active H for K exchange. Conformational changes require the presence of potassium-valinomycin and are not triggered by sodium. Potassium effect is enhanced by ATP, in the presence and in the absence of magnesium and, by ADP, in the presence of magnesium. Labeling of the pig HK-ATPase K518 by fluorescein-5-isothiocyanate inhibits the enzyme activity and knocks out the ATP effect on ANS fluorescence. Scherring 28080 and the monoclonal antibody 95-111, two competitive inhibitors of K-activated ATPase dephosphorylation, do not modify K-effect on ANS fluorescence but inhibit ATP effects. This supports that ANS does not probe K-site between the H1–H2 loop. Treatment of gastric membranes with trypsin does not inhibit the ANS response to potassium but does inhibit the response to ATP. This suggests that the ATP site inducing the ANS response is cytoplasmic and the potassium site is intramembranous. Titration reveals that one mole of ANS interacts with one mole of ATPase. We suggest that ANS probes a hydrophobic potassium site of gastric ATPase and that addition of ATP and ADP-Mg embed that site. Received: 16 July 1997/Revised: 10 June 1998  相似文献   

9.
Chloroplast coupling factor 1 (CF1) contains a high-affinity binding site for 8-anilino-1-napthalene sulphonate (ANS,Kd = 5-6 microM). The binding of ANS to the enzyme is associated with a fluorescence enhancement and a blue-shift in the emission spectrum. ANS only slightly inhibits ATP hydrolysis by CF1. Adenine nucleotides and inorganic phosphate induce a fast ANS fluorescence quenching of about 50% which is due to a decrease in the affinity of the enzyme for ANS (Kd increases from 6 microM to 22 microM) and in the fluorescence quantum yield of the bound probe (by 33%) but not in the number of ANS sites (n = 1). Conversely, Mg and Ca ions induce a fluorescence enhancement of bound ANS. Inactivation of the enzyme enhances ANS fluorescence, eliminates the response to adenine nucleotides and inorganic phosphate but increases the response to divalent metals. The affinity of latent CF1 for ADP (Kd = 12 microM) is considerably higher than for ATP (Kd = 95 microM) in buffer containing EDTA. The Kd for inorganic phosphate is 140 microM. Mg increases the apparent affinity for ATP (Kd = 28 microM) but not for ADP or Pi. Binding of ATP to the tight-sites does not inhibit the ADP or Pi-induced fluorescence quenching but decreases the affinity for ADP (Kd = 34 microM) and for inorganic phosphate (Kd = 320 microM). These results suggest that the ADP and phosphate binding sites are different but not independent from the tight sites. Activation of a Mg-specific ATPase in CF1 by octyl glucoside decreases the affinity for ADP and inorganic phosphate by about threefold but increases the affinity for ATP. ATPase activation of CF1 also increases the Ki for ADP inhibition of ATP hydrolysis. ATPase activation also influences the ANS responses to Ca and Mg. Ca-ATPase activation increases the fluorescence enhancement and the apparent affinity for Ca whereas Mg-ATPase activation specifically increases the Mg-induced fluorescence enhancement. The fluorescence of CF1-bound ANS is enhanced by Dio-9 and quenched by phloridzin, quercetin, Nbf-Cl and FITC. Nbf-Cl and FITC completely inhibit the ADP-induced fluorescence quenching whereas Dio-9 inhibits the Mg-induced fluorescence enhancement. ANS does not relieve the quercetin or phloridzin inhibition of ATP hydrolysis indicating that these inhibitors do not compete with ANS for a common binding site. ANS may be used, therefore, as a sensitive probe to detect conformational changes in CF1 in response to activation or inactivation and to binding of substrates and of inhibitors.  相似文献   

10.
Binding of 8-anilinonaphthalene sulfonate (ANS) to glutamate dehydrogenase results in enzyme inhibition and a marked increase in the fluorescence of ANS. Perphenazine and GTP increase the fluorescence of ANS-glutamate dehydrogenase secondary to their known ability to alter the conformation of this enzyme. Aspartate aminotransferases, which form enzyme-enzyme complexes with glutamate dehydrogenase, produce a slight decrease in the fluorescence of ANS-glutamate dehydrogenase.While ANS and perphenazine are allosteric inhibitors of reactions catalyzed by free glutamate dehydrogenase, they do not inhibit reactions catalyzed by aminotransferaseglutamate dehydrogenase complexes. This is in spite of the fact that the aminotransferase does not prevent either ANS or perphenazine from being bound to glutamate dehydrogenase. Therefore, reactions catalyzed by the enzyme-enzyme complex are apparently not inhibited by ANS or perphenazine because binding of the aminotransferase to glutamate dehydrogenase prevents these ligands from altering the conformation of glutamate dehydrogenase. This is consistent with the fact that the aminotransferase also prevents perphenazine from enhancing the fluorescence of ANS-glutamate dehydrogenase.Reactions catalyzed by the enzyme-enzyme complex are inhibited by GTP and the aminotransferase does not prevent GTP from enhancing the fluorescence of ANS-glutamate dehydrogenase. Therefore, binding of the aminotransferase to glutamate dehydrogenase does not prevent GTP from altering the conformation of glutamate dehydrogenase.The fact that the aminotransferase completely prevents perphenazine from increasing the fluorescence of ANS-glutamate dehydrogenase suggests that in the enzymeenzyme complex each glutamate dehydrogenase polypeptide chain can be bound to an aminotransferase polypeptide chain. This would mean that three aminotransferase molecules can be bound to each monomeric unit (Mr 3 × 105) of glutamate dehydrogenase.  相似文献   

11.
The pressure denaturation of trypsin from bovine pancreas was investigated by fluorescence spectroscopy in the pressure range 0. 1-700 MPa and by FTIR spectroscopy up to 1000 MPa. The tryptophan fluorescence measurements indicated that at pH 3.0 and 0 degrees C the pressure denaturation of trypsin is reversible but with a large hysteresis in the renaturation profile. The standard volume changes upon denaturation and renaturation are -78 mL.mol-1 and +73 mL.mol-1, respectively. However, the free energy calculated from the data in the compression and decompression directions are quite different in absolute values with + 36.6 kJ.mol-1 for the denaturation and -5 kJ. mol-1 for the renaturation. For the pressure denaturation at pH 7.3 the tryptophan fluorescence measurement and enzymatic activity assays indicated that the pressure denaturation of trypsin is irreversible. Interestingly, the study on 8-anilinonaphthalene-1-sulfonate (ANS) binding to trypsin under pressure leads to the opposite conclusion that the denaturation is reversible. FTIR spectroscopy was used to follow the changes in secondary structure. The pressure stability data found by fluorescence measurements are confirmed but the denaturation was irreversible at low and high pH in the FTIR investigation. These findings confirm that the trypsin molecule has two domains: one is related to the enzyme active site and the tryptophan residues; the other is related to the ANS binding. This is in agreement with the study on urea unfolding of trypsin and the knowledge of the molecular structure of trypsin.  相似文献   

12.
The type 1 human immunodeficiency virus presents a conical capsid formed by several hundred units of the capsid protein, CA. Homodimerization of CA occurs via its C-terminal domain, CA-C. This self-association process, which is thought to be pH-dependent, seems to constitute a key step in virus assembly. CA-C isolated in solution is able to dimerize. An extensive thermodynamic characterization of the dimeric and monomeric species of CA-C at different pHs has been carried out by using fluorescence, circular dichroism (CD), absorbance, nuclear magnetic resonance (NMR), Fourier transform infrared (FTIR), and size-exclusion chromatography (SEC). Thermal and chemical denaturation allowed the determination of the thermodynamic parameters describing the unfolding of both CA-C species. Three reversible thermal transitions were observed, depending on the technique employed. The first one was protein concentration-dependent; it was observed by FTIR and NMR, and consisted of a broad transition occurring between 290 and 315 K; this transition involves dimer dissociation. The second transition (Tm approximately 325 K) was observed by ANS-binding experiments, fluorescence anisotropy, and near-UV CD; it involves partial unfolding of the monomeric species. Finally, absorbance, far-UV CD, and NMR revealed a third transition occurring at Tm approximately 333 K, which involves global unfolding of the monomeric species. Thus, dimer dissociation and monomer unfolding were not coupled. At low pH, CA-C underwent a conformational transition, leading to a species displaying ANS binding, a low CD signal, a red-shifted fluorescence spectrum, and a change in compactness. These features are characteristic of molten globule-like conformations, and they resemble the properties of the second species observed in thermal unfolding.  相似文献   

13.
South African cassava mosaic virus (SACMV) is a circular ssDNA bipartite begomovirus, whose genome comprises DNA-A (encodes six genes) and DNA-B (encodes BC1 cell-to-cell movement and BV1 nuclear shuttle proteins) components. A few secondary and tertiary structural and physicochemical characteristics of partial but not full-length begomovirus proteins have been elucidated to date. The full-length codon-optimised SACMV BC1 gene was cloned into a pET-28a (+) expression vector and transformed into expression host cells E. coli BL21 (DE3). The optimal expression of the full-length BC1-encoded movement protein (MP) was obtained via induction with 0.25 mM IPTG at an OD600 of ~0.45 at 37 °C for four hours. Denatured protein fractions (dialysed in 4 M urea), passed through an IMAC column, successfully bound to the nickel resin, and eluted using 250 mM imidazole. The protein was refolded using stepwise dialysis. The molecular weight of MP was confirmed to be 35 kDa using SDS–PAGE. The secondary structure of SACMV MP presented as predominantly β-strands. An ANS (1-anilino-8-naphthalene sulphonate)-binding assay confirmed that MP possesses hydrophobic pockets with the ability to bind ligands such as ANS (8-anilino-1-naphthalenesulphonic acid). A 2′ (3′)-N-methylanthraniloyl-ATP (mant-ATP) assay showed binding of mant-ATP to MP and indicated that, while hydrophobic pockets are present, MP also exhibits hydrophilic regions. Intrinsic tryptophan fluorescence indicated a significant conformational change in the denatured form of BC1 in the presence of ATP. In addition, a phosphatase assay showed that MP possessed ATPase activity.  相似文献   

14.
Methionyl-tRNA synthetase from Bacillus stearothermophilus, a dimer of molecular weight 2 X 85K, is converted by limited subtilisin digestion into a fully active monomeric fragment of molecular weight 64K. The reversible methionine activation reaction of these enzymes was followed through the variation of the intensity of their trypotophan fluorescence. Equilibrium and stopped-flow experiments show that the rate and mechanism for adenylate formation supported by the monomeric derivative are undistinguishable from those of each adenylating site of the native dimeric enzyme. In contrast, the rate of tRNA aminoacylation is improved upon limited proteolysis of the native enzyme. This behavior can be related to the anticooperativity of the binding of tRNA molecules to native dimeric enzyme. Accordingly, at 25 degrees C, the dimer might behave as a half-of-the-sites enzyme with only one active tRNA site at a time, compared to two after limited proteolysis with consequent irreversible disociation into two 64K fragments. Another modified form of the enzyme is obtained through limited tryptic digestion. This derivative is completely devoid of activity although its molecular weight under nondenaturating conditions remains undistinguishable from that of the 64K fragment generated by subtilisin. Denaturation reveals that this tryptic derivative is composed of two subfragments with molecular weights of 33K and 29K, respectively. The same fragments may also be directly obtained through limited tryptic digestion of the subtilsic fragment. Interestingly, although trypsin treatment has abolished the activity of the enzyme, fluorescence studies demonstrate that the ATP and methionine binding sites have remained intact. It is shown that the effect of the internal cut made by trypsin into the active 64K fragment has been to considerably depress the "coupling" between the methionine and nucleotide binding sites. Finally, the rate of inactivation of the enzyme by trypsin is observed to be substantially decreased by in situ synthetized methionyl adenylate but not by tRNA. These properties and others are discussed in relation to the problem of its significance of repeating sequences and structural "domains" within the class of aminoacyl-tRNA synthetases.  相似文献   

15.
Banana lectin (BL) is a homodimeric protein categorized among jacalin-related family of lectins. The effect of acidic pH was examined on conformational stability of BL by using circular dichroism, intrinsic fluorescence, 1-anilino-8-napthalene sulfonate (ANS) binding, size exclusion chromatography (SEC) and dynamic light scattering (DLS). During acid denaturation of BL, the monomerization of native dimeric protein was found at pH 2.0. The elution profile from SEC showed two different peaks (59.65 ml & 87.98 ml) at pH 2.0 while single peak (61.45 ml) at pH 7.4. The hydrodynamic radii (R h) of native BL was 2.9 nm while at pH 2.0 two species were found with R h of 1.7 and 3.7 nm. Furthermore at, pH 2.0 the secondary structures of BL remained unaltered while tertiary structure was significantly disrupted with the exposure of hydrophobic clusters confirming the existence of molten globule like state. The unfolding of BL with different subunit status was further evaluated by urea and temperature mediated denaturation to check their stability. As inferred from high Cm and ΔG values, the monomeric form of BL offers more resistance towards chemical denaturation than the native dimeric form. Besides, dimeric BL exhibited a Tm of 77°C while no loss in secondary structures was observed in monomers even up to 95°C. To the best of our knowledge, this is the first report on monomeric subunit of lectins showing more stability against denaturants than its native dimeric state.  相似文献   

16.
Spontaneous mutations at numerous sites distant from the active site of human immunodeficiency virus type 1 protease enable resistance to inhibitors while retaining enzymatic activity. As a benchmark for probing the effects of these mutations on the conformational adaptability of this dimeric β-barrel protein, the folding free-energy surface of a pseudo-wild-type variant, HIV-PR?, was determined by a combination of equilibrium and kinetic experiments on the urea-induced unfolding/refolding reactions. The equilibrium unfolding reaction was well described by a two-state model involving only the native dimeric form and the unfolded monomer. The global analysis of the kinetic folding mechanism reveals the presence of a fully folded monomeric intermediate that associates to form the native dimeric structure. Independent analysis of a stable monomeric version of the protease demonstrated that a small-amplitude fluorescence phase in refolding and unfolding, not included in the global analysis of the dimeric protein, reflects the presence of a transient intermediate in the monomer folding reaction. The partially folded and fully folded monomers are only marginally stable with respect to the unfolded state, and the dimerization reaction provides a modest driving force at micromolar concentrations of protein. The thermodynamic properties of this system are such that mutations can readily shift the equilibrium from the dimeric native state towards weakly folded states that have a lower affinity for inhibitors but that could be induced to bind to their target proteolytic sites. Presumably, subsequent secondary mutations increase the stability of the native dimeric state in these variants and, thereby, optimize the catalytic properties of the resistant human immunodeficiency virus type 1 protease.  相似文献   

17.
Several bacteria use trimethylamine N-oxyde (TMAO) as an exogenous electron acceptor for anaerobic respiration. This metabolic pathway involves expression of the tor operon that codes for a periplasmic molybdopterin-containing reductase of the DMSO/TMAO family, a pentahemic c-type cytochrome, and the TorD cytoplasmic chaperone, possibly required for acquisition of the molybdenum cofactor and translocation of the reductase by the twin-arginine translocation system. In this report, we show that the TorD chaperone from Shewanella massilia forms multiple and stable oligomeric species. The monomeric, dimeric, and trimeric forms were purified to homogeneity and characterized by analytical ultracentrifugation. Small-angle X-ray scattering (SAXS) and preliminary diffraction data indicated that the TorD dimer is made of identical protein modules of similar size to the monomeric species. Interconversion of the native oligomeric forms occurred at acidic pH value. In this condition, ANS fluorescence indicates a non-native conformation of the polypeptide chain in which, according to the circular dichroism spectra, the alpha-helical content is similar to that of the native species. Surface plasmon resonance showed that both the monomeric and dimeric species bind the mature TorA enzyme, but that the dimer binds its target protein more efficiently. The possible biologic significance of these oligomers is discussed in relation to the chaperone activity of TorD, and to the ability of another member of the TorD family to bind the Twin Arginine leader sequences of the precursor of DMSO/TMAO reductases.  相似文献   

18.
Proteins for therapeutic use may contain small amounts of partially misfolded monomeric precursors to postproduction aggregation. To detect these misfolded proteins in the presence of an excess of properly folded protein, fluorescent probes such as 8-anilino-1-naphthalene sulfonate (ANS) are commonly used. We investigated the possibility of using isothermal titration calorimetry (ITC) to improve the detection of this type of conformational change using hydrophobic probes. As a case study, conformational changes in human polyclonal immunoglobulin G (IgG) were monitored by measuring the enthalpies of binding of ANS using ITC. Results were compared with those using fluorescence spectroscopy. IgG heated at 63 °C was used as a model system for “damaged” IgG. Heat-treated IgG can be detected already at levels below 5% with both ITC and fluorescence. However, ITC allows a much wider molar probe-to-protein ratio to be sampled. In particular, using reverse titration experiments (allowing high probe-to-protein ratios not available to fluorescence spectroscopy), an increase in the number of binding sites with a Kd > 10 mM was observed for heat-treated IgG, reflecting subtle changes in structure. Both ITC and fluorescence spectroscopy showed low background signals for native IgG. The nature of the background signals was not clear from the fluorescence measurements. However, further analysis of the ITC background signals shows that a fraction (8%) binds ANS with a dissociation constant of approximately 0.2 mM. Measurements were also carried out at pH 4.5. Precipitation of IgG was induced by ANS at concentrations above 0.5 mM, interfering with the ITC measurements. Instead, with the nonfluorescent probes 4-amino-1-naphthalene sulfonate and 1-naphthalene sulfonate, no precipitation is observed. These probes yield differences in the enthalpies of binding to heated and nonheated IgG similar to ANS. The data illustrate that ITC with low-molecular-weight probes is a versatile tool to monitor conformational changes in proteins with a wider application potential than fluorescence measurements.  相似文献   

19.
pH-dependent conformational changes are known to occur in dimeric procaspase-3, and they have been shown to affect the rate of automaturation. We studied the equilibrium unfolding of procaspase-3(C163S) as a function of pH (between pH 8.5 and pH 4) in order to examine these changes in the context of folding and stability. The data show that the procaspase dimer undergoes a pH-dependent dissociation below pH 5, so that the protein is mostly monomeric at pH 4. Consistent with this, the dimer unfolds via a four-state process between pH 8.5 and pH 4.75, in which the native dimer isomerizes to a dimeric intermediate, and the dimeric intermediate dissociates to a monomer, which then unfolds. In contrast, a small protein concentration dependence was observed by circular dichroism, but not by fluorescence emission, at pH 4.5 and pH 4.2. There was no protein-concentration dependence to the data collected at pH 4. Overall, the results are consistent with the redistribution of the population of native dimer (N(2)) to dimeric intermediate (I(2)) to monomeric intermediate (I), as the pH is lowered so that at pH 4, the "native" ensemble resembles the monomeric intermediate (I) observed during unfolding at higher pH. An emerging picture of the monomeric procaspase is discussed. Procaspase-3 is most stable at pH approximately 7 (24-26 kcal/mol), and while the stability decreased with pH, it was observed that dimerization contributes the majority (>70%) of the conformational free energy.  相似文献   

20.
R Tellam  J de Jersey  D J Winzor 《Biochemistry》1979,18(24):5316-5321
The binding of N-acetyl-tryptophan to the monomeric and dimeric forms of alpha-chymotrypsin in I = 0.2 acetate-chloride buffer, pH 3.86, has been studied quantitatively. Equilibrium sedimentation studies in the absence of inhibitor yielded a dimerization constant of 3.5 L/g. This value was confirmed by frontal gel chromatography of the enzyme on Bio-Gel P-30, which was also used to establish that N-acetyl-L-tryptophan binds preferentially to monomeric enzyme. From kinetic studies of competitive inhibition with N-acetyl-L-tryptophan ethyl ester as substrate, an equilibrium constant of 1300 M-1 was determined for the binding of N-acetyl-L-tryptophan to monomeric alpha-chymotrypsin. An intrinsic binding constant of 250 M-1 for the corresponding interaction with dimeric enzyme was calculated on the basis of these results and binding data obtained with concentrated (18.5 g/L) alpha-chymotrypsin. The present results refute earlier claims for exclusive binding of competitive inhibitors to monomer and also those for equivalence of inhibitor binding to monomeric and dimeric forms of alpha-chymotrypsin.  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