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.
A new caged proton, 1-(2-nitrophenyl)ethyl sulfate (caged sulfate), is characterized by infrared spectroscopy and compared with a known caged, proton 2-hydroxyphenyl 1-(2-nitrophenyl)ethyl phosphate (caged HPP). In contrast to caged HPP, caged sulfate can induce large pH jumps and protonate groups that have pK values as low as 2.2. The photolysis mechanism of caged sulfate is analogous to that of P(3)-[1-(2-nitrophenyl)ethyl] ATP (caged ATP), and the photolysis efficiency is similar. The utility of this new caged compound for biological studies was demonstrated by its ability to drive the acid-induced conformational change of metmyoglobin. This transition from the native conformation to a partially unfolded form takes place near pH 4 and was monitored by near-UV absorption spectroscopy.  相似文献   

2.
A new caged proton, 2-methoxy-5-nitrophenyl sulfate, was synthesized and used in time-resolved pH jump experiments to study proton binding in the sarcoplasmic reticulum Ca-ATPase. The major advantage of this compound is that it does not produce significant artifacts in experiments in which the fluorescent styryl dye 2BITC is used to monitor ion movements in the Ca pump. Two rate-limiting processes were resolved and their dependence on pH, Ca(2+) concentration, and temperature investigated. The faster process showed a relaxation time between 4 and 8 ms independent on pH and Ca(2+) concentration, and the time constant of the slower process varied between 31 ms (0 Ca(2+)) and 100 ms (100 microM Ca(2+)). A consistent mechanism to explain the results was derived in agreement with previous studies and the generally accepted Post-Albers scheme of the pump cycle. This mechanism requires that under physiological conditions the ion-binding sites are always occupied and two protons and a Ca(2+) ion replace each other. In the absence of ATP at low pH a nonphysiological state can be induced in which up to four protons bind to the Ca pump in the E(1) conformation. So far it could not be verified whether these additional protons bind to amino acid side chains or are coordinated as hydronium ions.  相似文献   

3.
Fibroblast growth factor (FGF) 1 is known to be released in response to cellular stress conditions through formation of a multi-protein complex with synaptotagmin 1 and S100A13. In this study, we characterized the denaturant-induced unfolding of synaptotagmin 1, C2A domain in a residue-specific manner by NMR spectroscopy. The amide protons of 30 residues distributed throughout the 3D structure of the whole protein could be followed in a series of 1H-15N HSQC spectra recorded from 0 to 8 M urea under equilibrium conditions. The midpoint for the urea-induced unfolding obtained from NMR coincides with those obtained from steady state fluorescence and CD spectroscopy, revealing that the protein unfolds via a two-state mechanism without accumulating stable intermediates. The thermodynamic parameter obtained from the denaturation curve illustrates the cooperative unfolding of the C2A domain. The implications of C2A domain folding in relation to the release of FGF-1 from the multi-protein complex were discussed.  相似文献   

4.
Experiments were carried out to measure the effect of concentrations of glycerol on H-exchange (HX) rates by using myoglobin as a test protein. Concentrated glycerol has only a small slowing effect on the HX kinetics of freely exposed amides, studied in a small molecule model (acetamide). Larger effects occur in structured proteins. The effect of solvent glycerol on different parts of the HX curve of myoglobin was studied by use of a selective "kinetic labeling" approach. Concentrated glycerol exerts an apparently reverse effect on protein H exchange; the faster exchanging "surface" protons are least affected, while the slower and slower amide NH is further slowed by larger and larger factors. These results seem inconsistent with solvent penetration models which generally visualize slower and slower protons as being placed, and undergoing exchange, farther and farther from the solvent-protein interface. On the other hand, the results are as expected for the local unfolding model for protein H exchange since concentrated glycerol is known to stabilize proteins against unfolding. In the local unfolding model, slower exchanging protons are released by way of higher energy and therefore generally larger, unfolding reactions. Larger unfoldings must be more inhibited by the glycerol effect.  相似文献   

5.
1-(2-Nitrophenyl)ethyl esters of D-myo-inositol 1,4,5-trisphosphate (InsP3) have been synthesized and shown to have suitable properties for use as photolabile precursors of InsP3. Synthesis was accomplished by treatment of InsP3 with 1-(2-nitrophenyl)diazoethane in a CHCl3/water mixture. This resulted in esterification of each of the three phosphate residues in InsP3, the 1-phosphate being more reactive than the 4- or 5-phosphate. Singly esterified P-1, P-4, and P-5 esters, termed P-1, P-4, and P-5 caged InsP3, were isolated from the reaction mixture by anion-exchange HPLC and characterized by 500-MHz 1H NMR spectroscopy. Each of these caged InsP3 esters exists as a pair of diastereoisomers and was identified by examining the effects of pH and nitrophenyl ring current shielding on the chemical shifts of nonexchangeable inositol protons. 1H NMR spectra of InsP3 were analyzed for comparison. On photolysis the compounds released InsP3 with rate constants of 175 (P-1), 225 (P-4), and 280 s-1 (P-5) as determined by monitoring the aci-nitro decay reaction at pH 7.1, 0.2 M ionic strength, 21 degrees C. Quantum yields determined by steady-state near-UV photolysis were 0.65 +/- 0.08 for each compound. P-4 and P-5 caged InsP3 were the most promising biologically inactive InsP3 precursors since at concentrations up to 50 microM they did not release Ca2+ from smooth muscle sarcoplasmic reticulum (SR) and were not metabolized by vascular smooth muscle InsP3 5-phosphatase or bovine brain InsP3 3-kinase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Photolysable caged ligands are used to investigate protein function and activity. Here, we investigate the binding properties of caged nucleotides and their photo released products to well established but evolutionary and structurally unrelated nucleotide-binding proteins, rabbit muscle creatine kinase (RMCK) and human annexin A6 (hAnxA6), using saturation transfer difference NMR spectroscopy. We detect the binding of the caged nucleotides and discuss the general implications on interpreting data collected with photolysable caged ligands using different techniques. Strategies to avoid non-specific binding of caged compound to certain proteins are also suggested.  相似文献   

7.
Aggregation of the Alzheimer's disease-related Aβ1-28 peptide was induced by a rapid, sub-millisecond pH jump and monitored by time-resolved infrared spectroscopy on the millisecond to second time-scale. The release of protons was induced by the photolysis of a caged compound, 1-(2-nitrophenyl)ethyl sulfate (NPE-sulfate). The pH jump generated in our experimental setup is used to model the Aβ peptide structural conversions that may occur in the acidic endosomal/lysosomal cell compartment system. The aggregation of the Aβ1-28 peptide induced by the pH jump from 8.5 to < 6 yields an antiparallel β-sheet structure. The kinetics of the structural transition is biphasic, showing an initial rapid phase with a transition from random coil to an oligomeric β-sheet form with a time constant of 3.6 s. This phase is followed by a second slower transition, which yields larger aggregates during 48.0 s.  相似文献   

8.
The synthesis and characterization of a new photolabile precursor of glycine (caged glycine) is described. The alpha-carboxyl group of glycine is covalently coupled to the alpha-carboxy-2-nitrobenzyl (alphaCNB) protecting group. Photolysis of the caged glycine with UV light produces free glycine. At 308 nm, the compound photolyzes with a quantum yield of 0.38. The absorption spectrum and the pH dependence of a transient absorption produced after laser-flash illumination are typical for aci-nitro intermediates of alphaCNB-protected compounds. The time constant for the major component of the aci-nitro intermediate decay ( approximately 84% of the total aci-nitro absorbance) was determined to be 7 micros at physiological pH. A minor component ( approximately 16%) decays with a rate constant of 170 micros. The compound does not activate or inhibit the alpha(1)-homomeric glycine receptor transiently expressed in HEK293 cells. After photolysis with a 10 ns pulse of 325 nm laser light, the glycine released from the caged compound activates glycine-mediated whole-cell currents in the same cells. The rise of these currents can be measured in a time-resolved fashion and occurs on a millisecond to sub-millisecond time scale. It can be described with a single-exponential function over >85% of the total current. The rate constant of the current rise is about 2 orders of magnitude slower than the rate constant of caged glycine photolysis. Thermal hydrolysis of the alphaCNB-caged glycine takes place with a half-life of 15.6 h at physiological pH. The new caged glycine is the first in a series of photoprotected glycine derivatives that has the required properties for use with chemical kinetic methods for investigation of glycine-activated cell surface receptors. Photolysis is rapid and efficient with respect to the receptor reactions to be studied; hydrolysis in aqueous solution is sufficiently slow, and the compound is biologically inert. It will, therefore, be a useful tool for investigation of the processes leading to channel opening of glycine receptor channels and the effects of mutations of the glycine receptor and of inhibitors on these processes.  相似文献   

9.
10.
Juneja J  Udgaonkar JB 《Biochemistry》2002,41(8):2641-2654
The unfolding of ribonuclease A was studied in 5.2 M guanidine hydrochloride at pH 8 and 10 degrees C using multiple optical probes, native-state hydrogen exchange (HX), and pulse labeling by hydrogen exchange. First, native-state HX studies were used to demonstrate that the protein exists in two slowly interconverting forms under equilibrium native conditions: a predominant exchange-incompetent N form and an alternative ensemble of conformations, N(I), in which some amide hydrogens are fully exposed to exchange. Pulsed HX studies indicated that, during unfolding, the rates of exposure to exchange with solvent protons were similar for all backbone NH probe protons. It is shown that two parallel routes of unfolding are available to the predominant N conformation as soon as it encounters strong unfolding conditions. A fraction of molecules appears to rapidly form N(I) on one route. On the other route an exchange-incompetent intermediate state ensemble, I(U)(2), is formed. The kinetics of unfolding measured by far-UV circular dichroism (CD) were faster than those measured by near-UV CD and intrinsic tyrosine fluorescence of the protein. The logarithms of the rate constants of the unfolding reaction measured by all three optical probes also showed a nonlinear dependence on GdnHCl concentration. All of the data suggest that N(I) and I(U)(2) are nativelike in their secondary and tertiary structures. While N(I) unfolds directly to the fully exchange-competent unfolded state (U), I(U)(2) forms another intermediate I(U)(3) which then unfolds to U. I(U)(3) is devoid of all native alpha-helical secondary structure and has only 30% of the tertiary interactions still intact. Since the rates of global unfolding measured by near-UV CD and fluorescence agree well with the rates of exposure determined for all of the backbone NH probe protons, it appears that the rate-limiting step for the unfolding of RNase A is the dissolution of the entire native tertiary structure and penetration of water into the hydrophobic core.  相似文献   

11.
The conversion of substrate, heptenitol, to product, beta-1-C-methyl, alpha-D-glucose-1-phosphate (heptulose-2-P), in crystals of glycogen phosphorylase b has been studied by Laue and monochromatic diffraction methods. The phosphorolysis reaction in the crystal was started following liberation of phosphate from a caged phosphate compound, 3,5-dinitrophenyl phosphate (DNPP). The photolysis of DNPP, stimulated by flashes from a xenon flash lamp, was monitored in the crystal with a diode array spectrophotometer. In the Laue diffraction experiments, data to 2.8 A resolution were collected and the first time shot was obtained at 3 min from the start of reaction, and data collection comprised three 800-ms exposures. Careful data processing of Laue photographs for the large enzyme resulted in electron density maps of almost comparable quality to those produced by monochromatic methods. The difference maps obtained from the Laue measurements showed that very little catalysis had occurred 3 min and 1 h after release of phosphate, and a distinct peak consistent with the position expected for phosphate, in the attacking position was observed. Data collection times with monochromatic crystallographic methods on a home source took 16 h for data to 2.3 A resolution. Sufficient phosphate was released from the caged phosphate in the crystal from 5 flashes with a xenon flashlamp within 1 min for the reaction to go to completion within the time scale of the monochromatic data collection procedures. The heptulose-2-P product complex has been refined and the model agrees with that obtained previously with the major difference that the interchange of an aspartic acid (Asp 283) by an arginine (Arg 569) was not observed at the catalytic site. This change is part of the activation process of glycogen phosphorylase and may not have taken place in the current experiments because the caged compound binds weakly at the inhibitor site, restricting conformational change, and because activators of the enzymic reaction were not present in the crystal. In experiments with monochromatic radiation in which low phosphate concentrations were generated either by fewer photons or by diffusion of known phosphate concentrations, mixtures of substrate and product were observed. It was not possible through crystallographic refinement at 2.3 A resolution to establish the fractional occupancies of the enzyme-substrate and enzyme-product complexes, but the results did indicate that the reaction was proceeding slowly, consistent with approximate calculations for the likely rate of the reaction in the crystal.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Proton and sulfate inhibition of the obligatory chloride-chloride exchange of human erythrocytes was measured at 0 degrees C to determine their mechanism of reaction with the anion transporter. The proton and sulfate that are co-transported by this mechanism at higher temperatures behaved as nontransported inhibitors at 0 degrees C. We analyzed the data in terms of four molecular mechanisms: (1) HSO4- addition to the transporter; (2) ordered addition with the proton first; (3) ordered addition with the sulfate first; (4) random addition to the transporter. The Dixon plots of 1/MCl vs. [SO4] at different proton concentrations were not parallel. Thus protons and sulfate ions were not mutually exclusive inhibitors. The slope of these Dixon plots was independent of pH above 7.0, which indicates that sulfate could bind to the unprotonated carrier and excludes the first two mechanisms. Protons were inhibitors of chloride flux in the absence of sulfate, which indicates that protons could bind to the unloaded carrier and excludes mechanism 3. The KI for sulfate was 4.35 +/0 0.36 mM. The pK for the protonatable group was 5.03 +/- 0.02. The binding of either a proton or sulfate to the carrier decreased the KI of the other by ninefold. The only simple mechanism consistent with the data is a random-ordered mechanism with more transporters loaded with a sulfate than loaded with a proton at the pH and sulfate concentrations of plasma.  相似文献   

13.
Conformational changes occurring in the catalytic cycle of the H+/K+-ATPase were monitored by Fourier transform infrared spectroscopy (FTIR). Caged compounds were used to release ATP, in the presence of Ca2+, to induce the transition between the E1 and E1-P conformation of the H+/K+-ATPase. In addition to bands associated with the photolysis of the caged compounds, some peaks of the difference infrared spectra were associated with changes in secondary structure and modifications of the ionization in the side chains of amino-acid residues (Glu or Asp). These changes were specific to the reaction between the ligand and the enzyme. We estimated that 39 amino acids changed their secondary structure during the reaction and four amino-acid residues were deprotonated. Similar spectral changes appeared when ADP was released from its precursor. The release of Pi from the same caged molecule did not induce similar changes. Changes in tertiary structure occurring during the binding of adenosine and phosphorylation of the enzyme were demonstrated by recording hydrogen/deuterium exchange kinetics by attenuated total reflectance FTIR spectroscopy (ATR-FTIR). At least 129 amide protons were involved in a tertiary structure change induced by ATP. This suggested that secondary structure change transduced a much larger tertiary structure modification.  相似文献   

14.
In order to solve the problem of an apparent discrepancy between the pH variance of oxygen equilibrium curve and the linear relation between the number of released Bohr protons and the degree of ligation, precise oxygen equilibrium curves of human hemoglobin were determined at a number of pH values from 6.5 to 8.8. From the equilibrium data individual steps (Adair constants), ki (i equals 1, 2, 3, 4), were obtained and the number of Bohr protons (deltaHi+) released on the ith stage of oxygenation was estimated. The pH dependence of k4 was very small, while the other ks strongly depended on pH over the pH range examined. As a consequence, the contribution of each step of oxygen binding to the alkaline Bohr effect nonuniform: deltaH4 was very small compared with deltaH1+, deltaH2+, and deltaH3+. In spite of this, calcuation has shown that the fractional number of released protons is essentially proportional to fractional oxygen saturation because of cooperative effects in hemoglobin. Thus, the present study indicates that the linear relationship between the fractional number of released protons and the degree of ligation, as obtained from titration experiments, is not necessarily incompatible with the pH variance of the shape of the oxygen equilibrium curve. The nonuniform pH depencence of the Adair constants implies that the two-state allosteric model of Monod, J., Wyman, J., and Changeus, J.P. (1965) J. Mol. Biol. 12, 88-118 is not adequate to describe the heterotropic effect caused by protons.  相似文献   

15.
We employed X-irradiation to activate a caged amino acid with a 2-oxoalkyl group. We designed and synthesized tyrosine derivative caged by a 2-oxoalkyl group (Tyr(Oxo)) to evaluate its radiolytic one-electron reduction characteristics in aqueous solution. Upon hypoxic X-irradiation, Tyr(Oxo) released a 2-oxopropyl group to form the corresponding uncaged tyrosine. In addition, radiolysis of dipeptides containing Tyr(Oxo) revealed that the efficiency of radiolytic removal of 2-oxopropyl group increased significantly by the presence of neighboring aromatic amino acids.  相似文献   

16.
We have reported the discovery of gambogic acid (GA) as a potent apoptosis inducer and the identification of transferrin receptor as its molecular target. In order to understand the basic pharmacophore of GA for inducing apoptosis and to discover novel and simplified derivatives as potential anti-cancer agents, we explored the synthesis of caged 2,3,3a,7a-tetrahydro-3,6-methanobenzofuran-7(6H)-ones (4-oxatricyclo[4.3.1.0]decan-2-ones). Three types of 2,3,3a,7a-tetrahydro-3,6-methanobenzofuran-7(6H)-ones based on xanthone, 2-phenylchromene-4-one and benzophenone, were synthesized using a Claisen/Diels-Alder reaction cascade. All the reactions produced the targeted caged compound as well as its neo-isomer. The caged compounds based on xanthone and 2-phenylchromene-4-one were found to maintain the apoptosis inducing and cell growth inhibiting activity of GA, although with less potency. The caged compounds based on benzophenone were found to be inactive. Our study determined the minimum structure of GA for its apoptosis inducing activity, which could lead to the development of simple derivatives as potential anti-cancer drugs.  相似文献   

17.
Renthal R 《Biochemistry》2006,45(49):14559-14566
Reversible unfolding of helical transmembrane proteins could provide valuable information about the free energy of interaction between transmembrane helices. Thermal unfolding experiments suggest that this process for integral membrane proteins is irreversible. Chemical unfolding has been accomplished with organic acids, but the unfolding or refolding pathways involve irreversible steps. Sodium dodecyl sulfate (SDS) has been used as a perturbant to study reversible unfolding and refolding kinetics. However, the interpretation of these experiments is not straightforward. It is shown that the results could be explained by SDS binding without substantial unfolding. Furthermore, the SDS-perturbed state is unlikely to include all of the entropy terms involved in an unfolding process. Alternative directions for future research are suggested: fluorinated alcohols in homogeneous solvent systems, inverse micelles, and fragment association studies.  相似文献   

18.
It has recently been shown that adenosine-5'-triphosphate (ATP) is released together with glutamate from sensory axons in the olfactory bulb, where it stimulates calcium signaling in glial cells, while responses in identified neurons to ATP have not been recorded in the olfactory bulb yet. We used photolysis of caged ATP to elicit a rapid rise in ATP and measured whole-cell current responses in mitral cells, the output neurons of the olfactory bulb, in acute mouse brain slices. Wide-field photolysis of caged ATP evoked an increase in synaptic inputs in mitral cells, indicating an ATP-dependent increase in network activity. The increase in synaptic activity was accompanied by calcium transients in the dendritic tuft of the mitral cell, as measured by confocal calcium imaging. The stimulating effect of ATP on the network activity could be mimicked by photo release of caged adenosine 5'-diphosphate, and was inhibited by the P2Y(1) receptor antagonist MRS 2179. Local photolysis of caged ATP in the glomerulus innervated by the dendritic tuft of the recorded mitral cell elicited currents similar to those evoked by wide-field illumination. The results indicate that activation of P2Y(1) receptors in the glomerulus can stimulate network activity in the olfactory bulb.  相似文献   

19.
We develop a statistical mechanical theory for the mechanism of hydrogen exchange in globular proteins. Using the HP lattice model, we explore how the solvent accessibilities of chain monomers vary as proteins fluctuate from their stable native conformations. The model explains why hydrogen exchange appears to involve two mechanisms under different conditions of protein stability; (1) a “global unfolding” mechanism by which all protons exchange at a similar rate, approaching that of the denatured protein, and (2) a “stable-state” mechanism by which protons exchange at rates that can differ by many orders of magnitude. There has been some controversy about the stable-state mechanism: does exchange take place inside the protein by solvent penetration, or outside the protein by the local unfolding of a subregion? The present model indicates that the stable-state mechanism of exchange occurs through an ensemble of conformations, some of which may bear very little resemblance to the native structure. Although most fluctuations are small-amplitude motions involving solvent penetration or local unfolding, other fluctuations (the conformational distant relatives) can involve much larger transient excursions to completely different chain folds.  相似文献   

20.
W K?ller  P E Kolattukudy 《Biochemistry》1982,21(13):3083-3090
Cutinase from Fusarium solani f. sp. pisi was inhibited by diisopropyl fluorophosphate and phenylboronic acid, indicating the involvement of an active serine residue in enzyme catalysis. Quantitation of the number of phosphorylated serines showed that modification of one residue resulted in complete loss of enzyme activity. One essential histidine residue was modified with diethyl pyrocarbonate. This residue was buried in native cutinase and became accessible to chemical modification only after unfolding of the enzyme by sodium dodecyl sulfate. The modification of carboxyl groups with 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide in the absence of sodium dodecyl sulfate did not result in inactivation of the enzyme; however, such modifications in the presence of sodium dodecyl sulfate resulted in complete loss of enzyme activity. The number of residues modified was determined by incorporation of [14C]glycine ethyl ester. Modification of cutinase in the absence of sodium dodecyl sulfate and subsequent unfolding of the enzyme with detergent in the presence of radioactive glycine ester showed that one buried carboxyl group per molecule of cutinase resulted in complete inactivation of the enzyme. Three additional peripheral carboxyl groups were modified in the presence of sodium dodecyl sulfate. Carbethoxylation of the essential histidine and subsequent incubation with the esterase substrate p-nitrophenyl [1-14C]acetate revealed that carbethoxycutinase was about 10(5) times less active than the untreated enzyme. The acyl-enzyme intermediate was stabilized under these conditions and was isolated by gel permeation chromatography. The results of the present chemical modification study indicate that catalysis by cutinase involves the catalytic triad and an acyl-enzyme intermediate, both characteristic for serine proteases.  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