首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Elongation factor Tu (EF-Tu) promotes the binding of aminoacyl-tRNA (aa-tRNA) to the acceptor site of the ribosome. During the elongation cycle, EF-Tu interacts with guanine nucleotides, aa-tRNA and its nucleotide exchange factor (EF-Ts). Quantitative determination of the equilibrium dissociation constants that govern the interactions of mammalian mitochondrial EF-Tu (EF-Tu(mt)) with guanine nucleotides was the focus of the work reported here. Equilibrium dialysis with [3H]GDP was used to measure the equilibrium dissociation constant of the EF-Tu(mt) x GDP complex (K(GDP) = 1.0 +/- 0.1 microM). Competition of GTP with a fluorescent derivative of GDP (mantGDP) for binding to EF-Tu(mt) was used to measure the dissociation constant of the EF-Tu(mt) x GTP complex (K(GTP) = 18 +/- 9 microM). The analysis of these data required information on the dissociation constant of the EF-Tu(mt) x mantGDP complex (K(mGDP) = 2.0 +/- 0.5 microM), which was measured by equilibrium dialysis. Both K(GDP) and K(GTP) for EF-Tu(mt) are quite different (about two orders of magnitude higher) than the dissociation constants of the corresponding complexes formed by Escherichia coli EF-Tu. The forward and reverse rate constants for the association and dissociation of the EF-Tu(mt) x GDP complex were determined using the change in the fluorescence of mantGDP upon interaction with EF-Tu(mt). These values are in agreement with a simple equilibrium binding interaction between EF-Tu(mt) and GDP. The results obtained are discussed in terms of the recently described crystal structure of the EF-Tu(mt) x GDP complex.  相似文献   

2.
Immunogenic peptides have been shown to bind detergent-solubilized class II (Ia) molecules from mice. In this investigation, we report that highly purified HLA-DR (DR) molecules in detergent solution are capable of binding a synthetic peptide (HAp) derived from the influenza hemagglutinin sequence. Although the presentation of this peptide has been demonstrated only to DR1-restricted Th cells, the association rate constants for the formation of HAp-DR1, -DR5, and -DR8 complexes were essentially identical (ka = 1.1 x 10(2) to 1.6 x 10(2) M-1 s-1). By contrast, the value of the rate constants for the dissociation of preformed HAp-DR1, -DR5, and -DR8 complexes varied nearly threefold (kd = 1.6 x 10(6) to 4.4 x 10(-6) s-1). The value of the equilibrium dissociation constants (KD) derived from these rate constants were 13 nM, 24 nM, and 28 nM, for HAp-DR1, -DR5, and -DR8 complexes, respectively. Scatchard analysis demonstrated that the KD obtained from the rate constants for the HAp-DR1 reaction was in excellent agreement with that obtained under equilibrium conditions. SDS-PAGE confirmed that the HAp-DR complexes were remarkably stable, as HAp remained associated with the DR alpha beta heterodimer after treatment of the complexes with SDS and beta-mercaptoethanol. Steady-state binding studies demonstrated that 18% of all DR1 molecules had bound HAp at equilibrium, whereas only 3.8% of all DR8 molecules had bound HAp under identical conditions. The slight differences in the KD for HAp-DR complexes suggest that differences in the affinity of a peptide for DR alleles alone may not always explain the process of MHC restriction.  相似文献   

3.
Using surface plasmon resonance (SPR)-based biosensor analysis and fluorescence spectroscopy, the apparent kinetic constants, k(on) and k(off), and equilibrium dissociation constant, K(d), have been determined for the binding interaction between rabbit skeletal troponin C (TnC) and rabbit skeletal troponin I (TnI) regulatory region peptides: TnI(96-115), TnI(96-131) and TnI(96-139). To carry out SPR analysis, a new peptide delivery/capture system was utilized in which the TnI peptides were conjugated to the E-coil strand of a de novo designed heterodimeric coiled-coil domain. The TnI peptide conjugates were then captured via dimerization to the opposite strand (K-coil), which was immobilized on the biosensor surface. TnC was then injected over the biosensor surface for quantitative binding analysis. For fluorescence spectroscopy analysis, the environmentally sensitive fluoroprobe 5-((((2-iodoacetyl)amino)ethyl)amino) naphthalene-1-sulfonic acid (1,5-IAEDANS) was covalently linked to Cys98 of TnC and free TnI peptides were added. SPR analysis yielded equilibrium dissociation constants for TnC (plus Ca(2+)) binding to the C-terminal TnI regulatory peptides TnI(96-131) and TnI(96-139) of 89nM and 58nM, respectively. The apparent association and dissociation rate constants for each interaction were k(on)=2.3x10(5)M(-1)s(-1), 2.0x10(5)M(-1)s(-1) and k(off)=2.0x10(-2)s(-1), 1.2x10(-2)s(-1) for TnI(96-131) and TnI(96-139) peptides, respectively. These results were consistent with those obtained by fluorescence spectroscopy analysis: K(d) being equal to 130nM and 56nM for TnC-TnI(96-131) and TnC-TnI(96-139), respectively. Interestingly, although the inhibitory region peptide (TnI(96-115)) was observed to bind with an affinity similar to that of TnI(96-131) by fluorescence analysis (K(d)=380nM), its binding was not detected by SPR. Subsequent investigations examining salt effects suggested that the binding mechanism for the inhibitory region peptide is best characterized by an electrostatically driven fast on-rate ( approximately 1x10(8) to 1x10(9)M(-1)s(-1)) and a fast off-rate ( approximately 1x10(2)s(-1)). Taken together, the determination of these kinetic rate constants permits a clearer view of the interactions between the TnC and TnI proteins of the troponin complex.  相似文献   

4.
Characterizing how chemical compounds bind to human serum albumin (HSA) is essential in evaluating drug candidates. Using warfarin as a test system, we validate the application of BIACORE SPR biosensors to reliably determine binding constants for drug/HSA interactions. The binding responses for warfarin over HSA surfaces were extremely reproducible even though warfarin is small compared to the size of the immobilized protein. At high concentrations, warfarin bound at more than one site on HSA, which is consistent with its known binding properties. The affinity we determined for the high-affinity site (K(25 degrees C)(d) = 3.7 +/- 1.2 microM), as well as the dissociation rate constant (k(25 degrees C)(d) = 1.2 s(-1)), are also consistent with binding constants determined previously. These results validate the biosensor technology and illustrate how BIACORE can be used to study drug/HSA interactions in a high-resolution mode. Using a set of 10 test compounds, we present a protocol for determining equilibrium dissociation constants for HSA in a high-throughput mode. Our method involves working at low compound concentrations and fitting the equilibrium data for all compounds simultaneously. We show that the % bound values determined by SPR correlate with the values determined by solution-based methods. The ability to examine directly the binding of small molecules (130-800 Da), coupled with minimal sample requirements and automated instrumentation, makes BIACORE technology applicable for evaluating drug/HSA interactions.  相似文献   

5.
We determined the number and the dissociation rate constants of different complexes formed from arsenite and two peptides containing either one (RVCAVGNDYASGYHYGV for peptide 20) or three cysteines (LECAWQGK CVEGTEHLYSMKCK for peptide 10) via radioactive 73As-labeled arsenite and vacuum filtration methodology. Nonlinear regression analysis of the dissociation of both arsenite-peptide complexes showed that triphasic fits gave excellent r2 values (0.9859 for peptide 20 and 0.9890 for peptide 10). The first phase of arsenite-peptide dissociation had the largest span (decrease in binding), and the rate was too fast to be measured using vacuum filtration methods. The dissociation rate constants of arsenite-peptide complexes for the second phase were 0.35 and 0.54 min(-1) and for the third phase were 0.0071 and 0.0045 min(-1) for peptides 20 and 10, respectively. For peptide 20, the three spans of triphasic decay were 85%, 9%, and 7% of the total binding of 16.1 nmol/mg protein. For peptide 10, which can bind in both an intermolecular and intramolecular manner, the three spans of triphasic decay were 59%, 16%, and 25% of the total binding of 43.7 nmol/mg protein. Binding of trivalent arsenicals to peptides and proteins can alter their structure and function and contribute to adverse health outcomes such as toxicity and carcinogenicity.  相似文献   

6.
The binding of substrates and the herbicide N-(phosphonomethyl)glycine (glyphosate) to enolpyruvoylshikimate-3-phosphate (EPSP) synthase was evaluated by stopped-flow and equilibrium fluorescence measurements. Changes in protein fluorescence were observed upon the binding of EPSP and upon the formation of the enzyme-shikimate 3-phosphate-glyphosate ternary complex; no change was seen with either shikimate 3-phosphate (S3P) or glyphosate alone. By fluorescence titrations, the dissociation constants were determined for the formation of the enzyme binary complexes with S3P (Kd,S = 7 +/- 1.2 microM) and EPSP (Kd,EPSP = 1 +/- 0.01 microM). The dissociation constant for S3P was determined by competition with EPSP or by measurements in the presence of a low glyphosate concentration. At saturating concentrations of S3P, glyphosate bound to the enzyme--S3P binary complex with a dissociation constant of 0.16 +/- 0.02 microM. Glyphosate did not bind significantly to free enzyme, so the binding is ordered with S3P binding first: (formula; see text) where S refers to S3P, G refers to glyphosate, and E.S.G. represents the complex with altered fluorescence. The kinetics of binding were measured by stopped-flow fluorescence methods. The rate of glyphosate binding to the enzyme--S3P complex was k2 = (7.8 +/- 0.2) X 10(5) M-1 s-1, from which we calculated the dissociation rate k-2 = 0.12 +/- 0.02 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
H E Swaisgood  I M Chaiken 《Biochemistry》1986,25(14):4148-4155
Bovine neurophysin II (BNP II) was covalently immobilized on both nonporous and porous (200-nm pore diameter) glass beads and incorporated in a high-performance liquid chromatograph to evaluate analytical high-performance affinity chromatography as a microscale method for characterizing biomolecular interactions. By extension of the theoretical treatment of analytical affinity chromatography, both the self-association of neurophysin and its binding of the peptide hormone vasopressin were characterized by using a single chromatographic column containing immobilized neurophysin predominantly in the monomer form. Both [3H] [Arg8]vasopressin (AVP) and 125I-BNP II were rapidly eluted (less than 25 min). The relatively symmetrical elution peaks obtained allowed calculation of both equilibrium dissociation constants and kinetic dissociation rate constants. The dissociation constant measured chromatographically for the AVP-immobilized neurophysin complex, KM/L = 11 microM with porous glass beads and 75 microM with nonporous glass (NPG) beads, was in reasonable agreement with those previously obtained by curve fitting of Scatchard plots (16-20 microM) and from binding to [BNP II]Sepharose (50 microM). The values obtained are larger than that for dissociation of AVP from BNP II dimer, by a factor consistent with the intended nature of immobilized BNP II as monomers. Chromatography of BNP II on the [BNP II]NPG gave a dimer dissociation constant of 166 microM, a value in excellent agreement with that derived from equilibrium sedimentation studies (172 microM). In contrast to the agreement of chromatographic equilibrium binding constants with those measured in solution, the dissociation rate, k-3, determined from the variance of the affinity chromatographic elution profile with nonporous beads, was several orders of magnitude smaller than the solution counterpart.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Nitrophorins 1-4 (NP1-4) are ferriheme proteins from the blood-sucking insect Rhodnius prolixus that transport nitric oxide (NO) to the victim, sequester histamine, and inhibit blood coagulation. Here, we report kinetic and thermodynamic analyses for ligand binding by all four proteins and their reduction potentials. All four undergo biphasic association and dissociation reactions with NO. The initial association is fast (1.5-33 microM(-)(1) s(-)(1)) and similar to that of elephant metmyoglobin. However, unlike in metmyoglobin, a slower second phase follows ( approximately 50 s(-)(1)), and the stabilized final complexes are resistant to autoreduction (E degrees = +3 to +154 mV vs normal hydrogen electrode). NO dissociation begins with a slow, pH-dependent step (0.02-1.4 s(-)(1)), followed by a faster phase that is again similar to that of metmyoglobin (3-52 s(-)(1)). The equilibrium dissociation constants are quite small (1-850 nM). NP1 and NP4 display larger release rate constants and smaller association rate constants than NP2 and NP3, leading to values for K(d) that are about 10-fold greater. The results are discussed in light of the recent crystal structures of NP1, NP2, and NP4, which display open, polar distal pockets, and of NP4-NO, which displays an NO-induced conformational change that leads to expulsion of solvent and complete burial of the NO ligand in a now nonpolar distal pocket. Taken together, the results suggest that tighter NO binding in the nitrophorins is due to the trapping of the molecule in a nonpolar distal pocket rather than through formation of particularly strong Fe-NO or hydrogen bonds.  相似文献   

9.
The pharmacological specificity of a nicotinic acetylcholine receptor (nAChR) optical biosensor was investigated using three fluorescein isothiocyanate (FITC)-tagged neurotoxic peptides that vary in the reversibility of their receptor inhibition: alpha-bungarotoxin (alpha-BGT), alpha-Naja toxin (alpha-NT), and alpha-conotoxin (GI) (alpha-CNTX). Kinetic analysis of the time course of binding of FITC-neurotoxins to the nAChR-coated fiber gave association rate constants (k+1) of 8.4 x 10(6) M-1 min-1 for FITC-alpha-BGT, 6.0 x 10(6) M-1 min-1 for FITC-alpha-NT and 1.4 x 10(6) M-1 min-1 for FITC-alpha-CNTX. The dissociation rate constants (k-1) for the three neurotoxins were 7.9 x 10(-3) min-1. 4.8 x 10(-2) min-1 and 8.0 x 10(-1) min-1 for FITC-alpha-BGT. FITC-alpha-NT and FITC-alpha-CNTX, respectively. The equilibrium dissociation constant (Kd) values for the three toxins. calculated from these rare constants, were similar to published values obtained from tissue responses or ligand binding assays. The optical signal generated by FITC-alpha-NT binding to the nAChR-coated fiber was effectively quenched by agonists and antagonists of the nAChR but not by most of the tested agonists and antagonists of muscarinic cholinergic, adrenergic, glutamatergic, serotonergic, dopaminergic or GABAergic receptors. Interestingly, 5-hydroxy-tryptamine, haloperidol and (+)cis-methyldioxolane gave significant inhibition of FITC-alpha-NT binding to the immobilized receptor. Equilibrium constants of inhibition (Ki) for d-tubocurarine (d-TC) and carbamylcholine (carb) were determined from competition studies using FITC-alpha-CNTX. FITC-alpha-NT or FITC-alpha-BGT as probes for receptor occupancy. When the more reversible probe FITC-alpha-CNTX was used, the Ki value for d-TC was an order of magnitude lower than those determined using the less reversible probes. Ki values for carb however, were independent of the FITC-toxin probe used.  相似文献   

10.
The connection between experimentally measured values of ED50 (concentration of added peptide required to bind half of the protein), which characterize peptide-protein binding and the equilibrium dissociation constant of peptide-protein complex Kd (affinity) is considered. It is shown and confirmed by experimental studies that in certain cases, as a result of the absence of equilibrium in the system, the value of Kd could be much less than the experimental value of ED50, but not equal to that as commonly assumed. This is especially applicable to the formation of peptide-MHC complexes with low dissociation rates (strong binding), which may require longer time-intervals to reach equilibrium. Thus the search of the good binding peptides based on finding ones with the smallest measured values' of ED50 may result in missing the best binders with the lowest values of dissociation constant (highest affinity). To analyze the problem we considered the formal chemical kinetics of peptide-protein binding. Experimental studies of peptide binding was performed to obtain the parameters of the kinetic model. According to the predictions of the model, it was confirmed that peptide binding occurs through the preceding step, which is either a release of an endogenous peptide or some conformational change of the molecule. The half decay time for this process was determined to be approximately 3 h. Based on the model developed, a new effective method for determination of the dissociation rates of peptide-MHC complexes and the equilibrium dissociation constants Kd was proposed, which implies the comparison of binding levels (ED50) at different instants of time. This method works especially well for the peptide-MHC complexes with relatively slow dissociation rates (stable complexes), for which the direct off-rate measurements as well as obtaining equilibrium binding data to determine Kd are highly time consuming and not very reliable.  相似文献   

11.
Ryanodine is a specific ligand for the calcium release channel which mediates calcium release in excitation-contraction coupling in muscle. In this study, ryanodine binding in sarcoplasmic reticulum from heart muscle and skeletal muscle is further compared and correlated with function. The new findings include the following: (1) Two types of binding, high affinity (KD1 approximately 5-10 nM) and low affinity (KD2 approximately 3 microM), can now be discerned for the skeletal muscle receptor. KD1 is approximately the same as and KD2 of similar magnitude to that previously reported for heart. (2) The dissociation rates for the high-affinity binding have been directly measured for both heart and skeletal muscle (t1/2 approximately 30-40 min). These rates are more rapid than previously reported (t1/2 approximately 14 h). (3) KD1's obtained from the ratio of the dissociation and association rate constants agree with the dissociation constant measured by equilibrium binding Scatchard analysis. (4) Ryanodine binding to the low-affinity site can be correlated with a decrease in the dissociation rate constant (k-1) of the high-affinity site, and thereby in the apparent dissociation constant (KD1). The inhibition constant (KI) for inhibiting the high-affinity off rate obtained from a double-reciprocal plot of the change in off rate vs [ryanodine] is practically the same in heart (0.66 microM) and skeletal muscle (0.64 microM) and in the range of the KD2. The binding of cold ryanodine to the low-affinity site appears to lock the bound [3H]ryanodine onto the high-affinity site rather than to exchange with it. Thus, in this sense, the ryanodine receptor exhibits "positive cooperativity".(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Stopped-flow fluorescence anisotropy was used to determine the kinetic parameters that define acetylation-dependent bromodomain-histone interactions. Bromodomains are acetyllysine binding motifs found in many chromatin associated proteins. Individual bromodomains were derived from the polybromo-1 protein, which is a subunit of the PBAF chromatin-remodeling complex that has six tandem bromodomains in the amino-terminal region. The average k(on) and k(off) values for the formation of high-affinity complexes are 275 M(-1) s(-1) and 0.41 x 10(-3) s(-1), respectively. The average k(on) and k(off) values for the formation of low-affinity complexes are 119 M(-1) s(-1) and 1.42 x 10(-3) s(-1), respectively. Analysis of the on- and off-rates yields acetylation site-dependent equilibrium dissociation constants averaging 1.4 and 12.9 microM for high- and low-affinity complexes, respectively. This work represents the first examination of kinetic mechanisms of acetylation-dependent bromodomain-histone interactions.  相似文献   

13.
The binding of the monoiodinated alpha-neurotoxin I from Naja mossambica mossambica to the membrane-bound acetylcholine receptor from Torpedo marmorata was investigated using a new picomolar-sensitive microtitration assay. From equilibrium binding studies a non-linear Scatchard plot demonstrated two populations of binding sites characterized by the two dissociation constants Kd1 = 7 +/- 4 pM and Kd2 = 51 +/- 16 pM and having equal binding capacities. These two populations differed in their rate of dissociation (k-1.1 = 25 x 10(-6) s-1 and k-1.2 = 623 x 10(-6) s-1 respectively), but not in their rate of formation of the toxin-receptor complex (k + 1 = 11.7 x 10(6) M-1 s-1). From these rate constants the same two values of dissociation constant were deduced (Kd1 = 2 pM and Kd2 = 53 pM). All the specific binding was prevented by the cholinergic antagonists alpha-bungarotoxin and d-tubocurarine. In addition, a biphasic competition phenomenon allowed us to differentiate between two d-tubocurarine sites (Kda = 103 nM and Kdb = 13.7 microM respectively). Evidence is provided indicating that these two sites are shared by d-tubocurarine and alpha-neurotoxin I, with inverse affinities. Fairly conclusive agreement between our equilibrium, kinetic and competition data demonstrates that the two high-affinity binding sites for this short alpha-neurotoxin are selectively distinguishable.  相似文献   

14.
Angiotensin converting enzyme interacts with the chelator, 1,10-phenanthroline (OP) to form an OP-Zn-ACE ternary complex, which subsequently dissociates to OP-Zn and apoenzyme. The association and dissociation rate constants for the reaction OP + Zn-ACE in equilibrium OP-Zn-ACE have been determined and compared with those of known OP-metal complexes. Such constants were also used to calculate the rate constant for formation of the OP-Zn complex from OP-Zn-ACE. The rate of dissociation of zinc from ACE has been measured in the presence of EDTA (which acts only as a metal scavenger) as a function of chelator concentration, at different pH values, and with different buffers. The stability constant for the binding of zinc to apoACE log Kc = 8.2, determined by equilibrium dialysis using atomic absorption spectroscopy to assess metal concentration, is much smaller than that for Zn-carboxypeptidase A. Zn-thermolysin, or Zn-carbonic anhydrase. This weak binding is attributable to the zinc dissociation rate constant of ACE, 7.5 X 10(-3) sec-1 at pH 7.0, which is much greater than that of the other zinc metalloenzymes. These results lead to inferences regarding the metal binding site of ACE.  相似文献   

15.
J Ellis  C R Bagshaw  W V Shaw 《Biochemistry》1991,30(44):10806-10813
Chloramphenicol acetyltransferase (CAT) catalyzes the acetyl-CoA-dependent acetylation of chloramphenicol by a ternary complex mechanism with a rapid equilibrium and essentially random order of addition of substrates. Such a kinetic mechanism for a two-substrate reaction provides an opportunity to compare the affinity of enzyme for each substrate in the binary complexes (1/Kd) with corresponding values (1/Km) for affinities in the ternary complex where any effect of the other substrate should be manifest. The pursuit of such information for CAT involved the use of four independent methods to determine the dissociation constant (Kd) for chloramphenicol in the binary complex, techniques which included stopped-flow measurements of on and off rates, and a novel fluorometric titration method. The binary complex dissociation constant (Kd) for acetyl-CoA was measured by fluorescence enhancement and steady-state kinetic analysis. The ternary complex dissociation constant (Km) for each substrate (in the presence of the other) was determined by kinetic and fluorometric methods, using CoA or ethyl-CoA to form nonproductive ternary complexes. The results demonstrate an unequivocal decrease in affinity of CAT for each of its substrates on progression from the binary to the ternary complex, a phenomenon most economically described as negative cooperativity. The binary complex dissociation constants (Kd) for chloramphenicol and acetyl-CoA are 4 microM and 30 microM whereas the corresponding dissociation constants in the ternary complex (Km) are 12 microM and 90 microM, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Total internal reflection fluorescence microscopy (TIRFM) has been used to directly measure the weak dissociation constants of IgG with a mouse IgG receptor (moFc gamma RII) that has been purified and reconstituted into substrate-supported planar membranes. Dissociation constants were measured for three different mouse monoclonal anti-dinitrophenyl (DNP) IgG1 antibodies and for polyclonal mouse IgG, in the absence and presence of saturating amounts of hapten (DNP-glycine). The dissociation constant for polyclonal mouse IgG was 3 microM, which agrees well with previous results. The dissociation constants for the three monoclonal antibodies with moFc gamma RII ranged from 2 microM to 3 microM and were not statistically different, suggesting that changes in moFc gamma RII dissociation constants which may exist within the IgG1 subclass are less than the error of the TIRFM measurements (approximately 20%). The measured IgG1-moFc gamma RII dissociation constants were not different for individual monoclonal antibodies in the absence or presence of saturating concentrations of DNP-glycine, directly showing that possible allosteric changes which might occur upon hapten binding and affect the equilibrium characteristics of Fc receptor binding are small. This work demonstrates a new approach for quantitatively examining the effects of solution components on weak receptor-ligand interactions.  相似文献   

17.
D A Malencik  S R Anderson 《Biochemistry》1984,23(11):2420-2428
Calmodulin and troponin C exhibit calcium-dependent binding of 1 mol/mol of dynorphin. The dissociation constants of the complexes, determined in 0.20 N KC1-1.0 mM CaCI2, pH 7.3, are 0.6 microM for calmodulin, 2.4 microM for rabbit fast skeletal muscle troponin C, and 9 microM for bovine heart troponin C. Experiments with deletion peptides of dynorphin show that peptide chain length and especially charge affect the binding of the peptides by calmodulin. Dynorphin, but not mastoparan or melittin, inhibits adenosinetriphosphatase activity in a reconstituted rabbit skeletal muscle actomyosin assay. The inhibition is partially reversed by the addition of calmodulin or troponin C in the presence of calcium. Calmodulin also exhibits calcium-dependent binding of a synthetic peptide corresponding to positions 104-115 of rabbit fast skeletal muscle troponin I. Mastoparan is a tetradecapeptide from the vespid wasp having exceptional affinity for calmodulin, with Kd approximately 0.3 nM [Malencik, D.A., & Anderson, S.R. (1983) Biochem. Biophys. Res. Commun. 114, 50]. The addition of 1 mol/mol of mastoparan to the complex of calmodulin with dynorphin results in complete dissociation of dynorphin. Similar titrations of the skeletal muscle troponin C-dynorphin complex produce a gradual dissociation consistent with a dissociation constant of 0.2 microM for the troponin C-mastoparan complex. Fluorescence anisotropy measurements using the intrinsic tryptophan fluorescence of mastoparan X show strongly calcium-dependent binding by proteolytic fragments of calmodulin. binding by proteolytic fragments of calmodulin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
19.
The Escherichia coli DEAD-box protein A (DbpA) is an RNA helicase that utilizes the energy from ATP binding and hydrolysis to facilitate structural rearrangements of rRNA. We have used the fluorescent nucleotide analogues, mantADP and mantATP, to measure the equilibrium binding affinity and kinetic mechanism of nucleotide binding to DbpA in the absence of RNA. Binding generates an enhancement in mant-nucleotide fluorescence and a corresponding reduction in intrinsic DbpA fluorescence, consistent with fluorescence resonance energy transfer (FRET) from DbpA tryptophan(s) to bound nucleotides. Fluorescent modification does not significantly interfere with the affinities and kinetics of nucleotide binding. Different energy transfer efficiencies between DbpA-mantATP and DbpA-mantADP complexes suggest that DbpA adopts nucleotide-dependent conformations. ADP binds (K(d) approximately 50 microM at 22 degrees C) 4-7 times more tightly than ATP (K(d) approximately 400 microM at 22 degrees C). Both nucleotides bind with relatively temperature-independent association rate constants (approximately 1-3 microM(-1) s(-1)) that are much lower than predicted for a diffusion-limited reaction. Differences in the binding affinities are dictated primarily by the dissociation rate constants. ADP binding occurs with a positive change in the heat capacity, presumably reflecting a nucleotide-induced conformational rearrangement of DbpA. At low temperatures (<22 degrees C), the binding free energies are dominated by favorable enthalpic and unfavorable entropic contributions. At physiological temperatures (>22 degrees C), ADP binding occurs with positive entropy changes. We favor a mechanism in which ADP binding increases the conformational flexibility and dynamics of DbpA.  相似文献   

20.
The apoenzyme of wild-type (WT) dihydrofolate reductase (DHRF) from Escherichia coli exists in two conformational states, Et and Ew, which differ in affinity for NADPH and in kinetic competence. Dissociation constants for the binary complex of NADPH with the two conformers differ by over 100-fold (KDt = 0.17 microM, KDw = 22 microM). Rate constants governing the interconversion of conformers are small (t1/2 for Ew----Et = 71 s), and since Ew is not catalytically competent, this conversion is accompanied by an increase in catalytic velocity. The equilibrium proportion of Et in the absence of ligands is 63%, but binding of NADPH greatly increases this proportion, and t1/2 for conversion of Ew.NADPH to Et.NADPH is 30 s. This conformational equilibrium has also been examined in mutant enzyme in which aspartate 27 is replaced by asparagine (D27N E. coli DHFR). Although ASp27 is an active site residue, it does not interact directly with bound NADPH, and in the mutant the rate constant for NADPH binding to Et is unchanged as are the dissociation constants for NADPH complexes with Et or Ew. However, for mutant apoenzyme, the proportion of Et is decreased to 18% in the absence of ligands so that the overall KD for NADPH is increased (0.15 microM for WT E. coli DHFR, 0.68 microM for D27N E. coli DHFR). The lower proportion of Et is due to a decreased rate for Ew----Et (t1/2 = 221 s) and an increased rate for Et----Ew (t1/2 = 50 s versus 120 s for WT E. coli DHFR).  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