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 718 毫秒
1.
The thermodynamics of binding of both the substrate glutathione (GSH) and the competitive inhibitor S-hexylglutathione to the mutant Y49F of human glutathione S-transferase (hGST P1-1), a key residue at the dimer interface, has been investigated by isothermal titration calorimetry and fluorescence spectroscopy. Calorimetric measurements indicated that the binding of these ligands to both the Y49F mutant and wild-type enzyme is enthalpically favorable and entropically unfavorable over the temperature range studied. The affinity of these ligands for the Y49F mutant is lower than those for the wild-type enzyme due mainly to an entropy change. Therefore, the thermodynamic effect of this mutation is to decrease the entropy loss due to binding. Calorimetric titrations in several buffers with different ionization heat amounts indicate a release of protons when the mutant binds GSH, whereas protons are taken up in binding S-hexylglutathione at pH 6.5. This suggests that the thiol group of GSH releases protons to buffer media during binding and a group with low pKa (such as Asp98) is responsible for the uptake of protons. The temperature dependence of the free energy of binding, DeltaG0, is weak because of the enthalpy-entropy compensation caused by a large heat capacity change. The heat capacity change is -199.5 +/- 26.9 cal K-1 mol-1 for GSH binding and -333.6 +/- 28.8 cal K-1 mol-1 for S-hexylglutathione binding. The thermodynamic parameters are consistent with the mutation Tyr49 --> Phe, producing a slight conformational change in the active site.  相似文献   

2.
Competition experiments using 9-anthroylcholine, a fluorescent dye that undergoes calmodulin-dependent binding by smooth muscle myosin light chain kinase [Malencik, D. A., Anderson, S. R., Bohnert, J. L., & Shalitin, Y. S. (1982) Biochemistry 21, 4031], demonstrate a strongly stabilizing interaction between the adenosine 5'-triphosphate and myosin light chain binding sites operating within the enzyme-calmodulin complex but probably not in the free enzyme. The interactions in the latter case may be even slightly destabilizing. The fluorescence enhancement in solutions containing 5.0 microM each of the enzyme and calmodulin is directly proportional to the maximum possible concentration of bound calcium on the basis of four calcium binding sites. Evidently, all four calcium binding sites of calmodulin contribute about equally to the enhanced binding of 9-anthroylcholine by the enzyme. Fluorescence titrations on solutions containing 1.0 microM enzyme plus calmodulin yield a Hill coefficient of 1.2 and K = 0.35 +/- 0.08 microM calcium. Three proteolytic fragments of smooth muscle myosin light chain kinase, apparent products of endogenous proteolysis, were isolated and characterized. All three possess calmodulin-dependent catalytic activity. Their interactions with 9-anthroylcholine, in both the presence and absence of calmodulin, are similar to those of the native enzyme. However, the stabilities of their complexes with calmodulin vary. The corresponding dissociation constants range from 2.8 nM for the native enzyme and 8.5 nM for the 96K fragment to approximately 15 nM for the 68K and 90K fragments [0.20 N KCl, 50 mM 3-(N-morpholino)propanesulfonic acid, and 1 mM CaCl2, pH 7.3, 25 degrees C]. A coupled fluorometric assay, modified from a spectrophotometric assay for adenosine cyclic 3',5'-phosphate dependent protein kinase [Cook, P. F., Neville, M. E., Vrana, K. E., Hartl, F. T., & Roskoski, R. (1982) Biochemistry 21, 5794], has provided the first continuous recordings of myosin light chain kinase phosphotransferase activity. The results show that smooth muscle myosin light chain kinase is a responsive enzyme, whose activity adjusts rapidly to changes in solution conditions.  相似文献   

3.
The glutathione S-transferase enzymes (GSTs) have a tyrosine or serine residue at their active site that hydrogen bonds to and stabilizes the thiolate anion of glutathione, GS(-). The importance of this hydrogen bond is obvious, in light of the enhanced nucleophilicity of GS(-) versus the protonated thiol. Several A-class GSTs contain a C-terminal segment that undergoes a ligand-dependent local folding reaction. Here, we demonstrate the effects of the Y9F substitution on binding affinity for glutathione conjugates and on rates of the order-disorder transition of the C terminus in rat GST A1-1. The equilibrium binding affinity of the glutathione conjugate, GS-NBD (NBD-Cl, 7-chloro-4-nitrobenzo-2-oxa-1, 3-diazole), was decreased from 4.09 microm to 0.641 microm upon substitution of Tyr-9 with Phe. This result was supported by isothermal titration calorimetry, with K(d) values of 1.51 microm and 0.391 microm for wild type and Y9F, respectively. The increase in binding affinity for the mutant is associated with dramatic decreases in rates for the C-terminal order-disorder transition, based on a stopped-flow kinetic analysis. The same effects were observed, qualitatively, for a second GSH conjugate, GS-ethacrynic acid. Apparently, the phenolic hydroxyl group of Tyr-9 is critical for orchestrating C-terminal dynamics and efficient product release, in addition to its role in lowering the pK(a) of GSH.  相似文献   

4.
Rev is an essential HIV-1 regulatory protein that binds the Rev responsive element (RRE) within the env gene of the HIV-1 RNA genome and is involved in transport of unspliced or partially spliced viral mRNA from the cell nucleus to the cytoplasm. Previous studies have shown that a short alpha-helical peptide derived from Rev (Rev 34-50), and a truncated form of the RRE sequence provide a useful in vitro system to study this interaction while still preserving the essential aspects of the native complex. We have selectively incorporated the fluorescent probe 2-aminopurine 2'-O-methylriboside (2-AP) into the RRE sequence in nonperturbing positions (A68 and U72) such that the binding of both Rev peptide and aminoglycoside ligands could be characterized directly by fluorescence methods. Rev peptide binding to the RRE-72AP variant resulted in a 2-fold fluorescence increase that provided a useful signal to monitor this binding interaction (K(D) = 20 +/- 7 nM). Using stopped-flow kinetic measurements, we have shown that specific Rev peptide binding occurs by a two-step process involving diffusion-controlled encounter, followed by isomerization of the RNA. Using the RRE-68AP and -72AP constructs, three classes of binding sites for the aminoglycoside neomycin were unambiguously detected. The first site is noninhibitory to Rev binding (K(D) = 0.24 +/- 0.040 microM), the second site inhibited Rev binding in a competitive fashion (K(D) = 1. 8 +/- 0.8 microM), and the third much weaker site (or sites) is attributed to nonspecific binding (K(D) >/= 40 microM). Complementary NMR measurements have shown that neomycin forms both a specific binary complex with RRE and a specific ternary complex with RRE and Rev. NMR data further suggest that neomycin occupies a similar high-affinity binding site in both the binary and ternary complexes, and that this site is located in the lower stem region of RRE.  相似文献   

5.
An Escherichia coli expression vector, containing the alkaline phosphatase promoter and the stII heat-stable enterotoxin signal sequence, along with the cDNA of the kringle 1 (K1) region of human plasminogen (HPg), has been employed to express into the periplasmic space amino acid residues 82-163 (E163----D) of HPg. This region of the molecule contains the entire K1 domain (residues C84-C162) of HPg, as well as two non-kringle amino-terminal amino acids (S82-E83) that are present in their normal locations in HPg and a carboxyl-terminal amino acid, D163, that results from mutation of the E163, normally present at this location in the HPg amino acid sequence. After purification of r-K1 by chromatographic techniques, we have investigated its omega-amino acid binding properties by titration calorimetry, intrinsic fluorescence, and differential scanning microcalorimetry (DSC). The antifibrinolytic agent, epsilon-aminocaproic acid (EACA), possesses a single binding site for r-K1. The thermodynamic properties of this interaction, studied by calorimetric titrations of the heats of binding with this ligand, reveal a Kd of 12 +/- 2 microM at 25 degrees C and pH 7.4, a corresponding delta G of -6.7 +/- 0.1 kcal/mol, a delta H of -3.6 +/- 0.1 kcal/mol, and a delta S of 10.5 +/- 0.8 eu. The intrinsic fluorescence of r-K1 decreases by approximately 44% when its binding site is saturated with EACA, and titrations of this perturbation with EACA lead to calculation of a Kd of approximately 13 microM, a value in good agreement with that obtained from titration calorimetric analysis. EACA represents the strongest binding ligand of a variety of simple aliphatic omega-amino acids examined. A cyclic analogue of EACA, trans-4-(aminomethyl)cyclohexanecarboxylic acid, interacts with r-K1 with an approximate 12-fold tighter Kd (1.0 +/- 0.2 microM). Investigations by DSC, at pH 7.4, demonstrate that a significant stabilization of the r-K1 structure occurs when EACA binds to this domain. The temperature of maximum heat capacity change (Tm) in the thermal denaturation of r-K1 increases from approximately 340.8 to 359.1 K as a consequence of EACA binding. These studies demonstrate that a fully functional EACA-binding kringle from HPg can be expressed and secreted in E. coli, purified by techniques that do not require refolding, and investigated as an independent structural unit.  相似文献   

6.
7.
Metal ions, such as Ca2+ and Mn2+, are necessary for the generation of cofactor activity following reconstitution of factor VIII from its isolated light chain (LC) and heavy chain (HC). Titration of EDTA-treated factor VIII with Mn2+ showed saturable binding with high affinity (K(d) = 5.7 +/- 2.1 microM) as detected using a factor Xa generation assay. No significant competition between Ca2+ and Mn2+ for factor VIII binding (K(i) = 4.6 mM) was observed as measured by equilibrium dialysis using 20 microM Ca2+ and 8 microM factor VIII in the presence of 0-1 mM Mn2+. The intersubunit affinity measured by fluorescence energy transfer of an acrylodan-labeled LC (fluorescence donor) and fluorescein-labeled HC (fluorescence acceptor) in the presence of 20 mM Mn2+ (K(d) = 53.0 +/- 17.1 nM) was not significantly different from the affinity value previously obtained in the absence of metal ion (K(d) = 53.8 +/- 14.2 nM). The sensitization of phosphorescence of Tb3+ bound to factor VIII subunits was utilized to detect Mn2+ binding to the subunits. Mn2+ inhibited the phosphorescence of Tb3+ bound to HC and LC, as well as the HC-derived A1 and A2 subunits with a relatively wide range of estimated inhibition constant values (K(i) values = 169-1147 microM), whereas Ca2+ showed no effect on Tb3+ phosphorescence. These results suggest that factor VIII cofactor activity can be generated by Mn2+ binding to site(s) on factor VIII that are different from the high-affinity Ca2+ binding site. However, like Ca2+, Mn2+ did not alter the affinity for HC and LC association. Thus, Mn2+appears to generate factor VIII cofactor activity by a similar mechanism as observed for Ca2+following its association at nonidentical sites on the protein.  相似文献   

8.
We recently identified residues 185-224 of the light chain of human high molecular weight kininogen (HMWK) as the binding site for plasma prekallikrein (Tait, J.F., and Fujikawa, K. (1986) J. Biol. Chem. 261, 15396-15401). In the present study, we have further defined the primary structure requirements for binding of HMWK to factor XI and prekallikrein. In a competitive fluorescence polarization binding assay, a 31-residue synthetic peptide (residues 194-224 of the HMWK light chain) bound to prekallikrein with a Kd of 20 +/- 6 nM, indistinguishable from the previously determined value of 18 +/- 5 nM for the light chain. We also prepared three shorter synthetic peptides corresponding to different portions of the 31-residue peptide (residues 205-224, 212-224, and 194-211), but these peptides bound to prekallikrein more than 100-fold more weakly. Factor XI also bound to the same region of the HMWK light chain, but at least 58 residues (185-242) were required for optimal binding (Kd = 69 +/- 4 nM for the light chain; Kd = 130 +/- 50 nM for residues 185-242). The four synthetic peptides inhibited kaolin-activated clotting of blood plasma with potencies paralleling their affinities for prekallikrein and factor XI. Peptide 194-224 can also be used for rapid affinity purification of prekallikrein and factor XI from plasma.  相似文献   

9.
Protein L is a bacterial surface protein with 4-5 immunoglobulin (Ig)-binding domains (B1-B5), each of which appears to have two binding sites for Ig, corresponding to the two edges of its beta-sheet. To verify these sites biochemically and to probe their relative contributions to the protein L-Ig kappa light chain (kappa) interaction, we compared the binding of PLW (the Y47W mutant of the B1 domain) to that of mutants designed to disrupt binding to sites 1 and 2, using gel filtration, BIAcore surface plasmon resonance, fluorescence titration, and solid-phase radioimmunoassays. Gel filtration experiments show that PLW binds kappa both in 1:1 complexes and multivalently, consistent with two binding sites. Covalent dimers of the A20C and V51C mutants of PLW were prepared to eliminate site 1 and site 2 binding, respectively; both the A20C and V51C dimers bind kappa in 1:1 complexes and multivalently, indicating that neither site 1 nor site 2 is solely responsible for kappa binding. The A20R mutant was designed computationally to eliminate site 1 binding while preserving site 2 binding; consistent with this design, the A20R mutant binds kappa in 1:1 complexes but not multivalently. To probe the contributions of amino acid side chains to binding, we prepared 75 point mutants spanning nearly every residue of PLW; BIAcore studies of these mutants revealed two binding-energy "hot spots" consistent with sites 1 and 2. These data indicate that PLW binds kappa at both sites with similar affinities (high nanomolar), with the strongest contributions to the binding energy from Tyr34 (site 2) and Tyr36 (site 1). Compared to other protein-protein complexes, the binding is insensitive to amino acid substitutions at these sites, consistent with the large number of main chain interactions relative to side chain interactions. The strong binding of protein L to Ig kappa light chains of various species may result from the ambidextrous binding of the B1-B5 domains and the unimportance of specific side chain interactions.  相似文献   

10.
In order to get insight into the origin of apparent negative cooperativity observed for F(1)-ATPase, we compared ATPase activity and ATPMg binding of mutant subcomplexes of thermophilic F(1)-ATPase, alpha((W463F)3)beta((Y341W)3)gamma and alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma. For alpha((W463F)3)beta((Y341W)3)gamma, apparent K(m)'s of ATPase kinetics (4.0 and 233 microM) did not agree with apparent K(m)'s deduced from fluorescence quenching of the introduced tryptophan residue (on the order of nM, 0.016 and 13 microM). On the other hand, in case of alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma, which lacks noncatalytic nucleotide binding sites, the apparent K(m) of ATPase activity (10 microM) roughly agreed with the highest K(m) of fluorescence measurements (27 microM). The results indicate that in case of alpha((W463F)3)beta((Y341W)3)gamma, the activating effect of ATP binding to noncatalytic sites dominates overall ATPase kinetics and the highest apparent K(m) of ATPase activity does not represent the ATP binding to a catalytic site. In case of alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma, the K(m) of ATPase activity reflects the ATP binding to a catalytic site due to the lack of noncatalytic sites. The Eadie-Hofstee plot of ATPase reaction by alpha((K175A/T176A/W463F)3)beta((Y341W)3)gamma was rather linear compared with that of alpha((W463F)3)beta((Y341W)3)gamma, if not perfectly straight, indicating that the apparent negative cooperativity observed for wild-type F(1)-ATPase is due to the ATP binding to catalytic sites and noncatalytic sites. Thus, the frequently observed K(m)'s of 100-300 microM and 1-30 microM range for wild-type F(1)-ATPase correspond to ATP binding to a noncatalytic site and catalytic site, respectively.  相似文献   

11.
The kinetics and energetics of the binding between barley alpha-amylase/subtilisin inhibitor (BASI) or BASI mutants and barley alpha-amylase 2 (AMY2) were determined using surface plasmon resonance and isothermal titration calorimetry (ITC). Binding kinetics were in accordance with a 1:1 binding model. At pH 5.5, [Ca(2+)] = 5 mM, and 25 degrees C, the k(on) and k(off) values were 8.3 x 10(+4) M(-1) s(-1) and 26.0 x 10(-4) s(-1), respectively, corresponding to a K(D) of 31 nM. K(D) was dependent on pH, and while k(off) decreased 16-fold upon increasing pH from 5.5 to 8.0, k(on) was barely affected. The crystal structure of AMY2-BASI shows a fully hydrated Ca(2+) at the protein interface, and at pH 6.5 increase of [Ca(2+)] in the 2 microM to 5 mM range raised the affinity 30-fold mainly due to reduced k(off). The K(D) was weakly temperature-dependent in the interval from 5 to 35 degrees C as k(on) and k(off) were only increasing 4- and 12-fold, respectively. A small salt dependence of k(on) and k(off) suggested a minor role for global electrostatic forces in the binding and dissociation steps. Substitution of a positively charged side chain in the mutant K140L within the AMY2 inhibitory site of BASI accordingly did not change k(on), whereas k(off) increased 13-fold. ITC showed that the formation of the AMY2-BASI complex is characterized by a large exothermic heat (Delta H = -69 +/- 7 kJ mol(-1)), a K(D) of 25 nM (27 degrees C, pH 5.5), and an unfavorable change in entropy (-T Delta S = 26 +/- 7 kJ mol(-1)). Calculations based on the thermodynamic data indicated minimal structural changes during complex formation.  相似文献   

12.
Fluorescence polarization spectroscopy and isothermal titration calorimetry were used to study the influence of osmolytes on the association of the anti-hen egg lysozyme (HEL) monoclonal antibody HyHEL-5 with bobwhite quail lysozyme (BWQL). BWQL is an avian species variant with an Arg-->Lys mutation in the HyHEL-5 epitope, as well as three other mutations outside the HyHEL-5 structural epitope. This mutation decreases the equilibrium association constant of HyHEL-5 for BWQL by over 1000-fold as compared to HEL. The three-dimensional structure of this complex has been obtained recently. Fluorescein-labeled BWQL, obtained by labeling at pH 7.5 and purified by hydrophobic interaction chromatograpy, bound HyHEL-5 with an equilibrium association constant close to that determined for unlabeled BWQL by isothermal titration calorimetry. Fluorescence titration, stopped-flow kinetics, and isothermal titration calorimetry experiments using various concentrations of the osmolytes glycerol, ethylene glycol, and betaine to perturb binding gave a lower limit of the uptake of approximately 6-12 water molecules upon formation of the HyHEL-5/BWQL complex.  相似文献   

13.
14.
Isothermal titration calorimetry has been used to determine thermodynamic parameters of substrate binding to the oxygenase domain of neuronal nitric oxide synthase (nNOS(oxy)) in the presence of the cofactor tetrahydrobiopterin. The intermediate N(omega)-hydroxy-L-arginine (NHA) has a larger affinity than L-Arginine (L-Arg) for nNOS(oxy), with K(d)=0.4+/-0.1 microM and 1.7+/-0.3 microM at 25 degrees C, respectively. nNOS(oxy) binds NHA and L-Arg with DeltaH -4.1+/-0.2 and -1.0+/-0.1 kcal/mol and DeltaS=15 and 23 cal/Kmol respectively. NHA binding is more exothermic probably due to formation of an extra hydrogen bond in the active site compared to L-Arg. The changes in heat capacity (DeltaC(p)) are relatively small for binding of both NHA and L-Arg (-53+/-18 and -95+/-23 cal/L mol, respectively), which indicates that hydrophobic interactions contribute little to binding.  相似文献   

15.
Lin SH  Lee JC 《Biochemistry》2002,41(39):11857-11867
The binding of adenosine 3',5'-cyclic monophosphate (cAMP) and its nonfunctional analogue, guanosine 3',5'-cyclic monophosphate (cGMP), to the adenosine 3',5'-cyclic monophosphate receptor protein (CRP) from Escherichia coli was investigated by means of fluorescence and isothermal titration calorimetry (ITC) at pH 7.8 and 25 degrees C. A biphasic fluorescence titration curve was observed, confirming the previous observation reported by this laboratory (Heyduk and Lee (1989) Biochemistry 28, 6914-6924). However, the triphasic titration curve obtained from the ITC study suggests that the cAMP binding to CRP is more complicated than the previous conclusion that CRP binds sequentially two molecules of cAMP with negative cooperativity. The binding data can best be represented by a model for two identical interactive high-affinity sites and one low-affinity binding site. Unlike cAMP, the binding of cGMP to CRP exhibits no cooperativity between the high-affinity sites. The effects of mutations on the bindings of cAMP and cGMP to CRP were also investigated. The eight CRP mutants studied were K52N, D53H, S62F, T127L, G141Q, L148R, H159L, and K52N/H159L. These sites are located neither in the ligand binding site nor at the subunit interface. The binding was monitored by fluorescence. Although these mutations are at a variety of locations in CRP, the basic mechanism of CRP binding to cyclic nucleotides has not been affected. Two cyclic nucleotide molecules bind to the high-affinity sites of CRP. The cooperativity of cAMP binding is affected by mutation. It ranges from negative to positive cooperativity. The binding of cGMP shows none. With the exception of the T127L mutant, the free energy change for DNA-CRP complex formation increases linearly with increasing free energy change associated with the cooperativity of cAMP binding. This linear relationship implies that the protein molecule modulates the signal in the binding of cAMP, even though the mutation is not directly involved in cAMP or DNA binding. In addition, the significant differences in the amplitude of fluorescent signal indicate that the mutations also affect the surface characteristics of CRP. These results imply that these mutations are not perturbing specific pathways of signal transmission. Instead, these results are more consistent with the concept that CRP exists as an ensemble of native states, the distribution of which can be altered by these mutations.  相似文献   

16.
The bacterial cocaine esterase, cocE, hydrolyzes cocaine faster than any other reported cocaine esterase. Hydrolysis of the cocaine benzoyl ester follows Michaelis-Menten kinetics with k(cat) = 7.8 s(-1) and K(M) = 640 nM. A similar rate is observed for hydrolysis of cocaethylene, a more potent cocaine metabolite that has been observed in patients who concurrently abuse cocaine and alcohol. The high catalytic proficiency, lack of observable product inhibition, and ability to hydrolyze both cocaine and cocaethylene make cocE an attractive candidate for rapid cocaine detoxification in an emergency setting. Recently, we determined the crystal structure of this enzyme, and showed that it is a serine carboxylesterase, with a catalytic triad formed by S117, H287, and D259 within a hydrophobic active site, and an oxyanion hole formed by the backbone amide of Y118 and the Y44 hydroxyl. The only enzyme previously known to use a Tyr side chain to form the oxyanion hole is prolyl oligopeptidase, but the Y44F mutation of cocE has a more deleterious effect on the specificity rate constant (k(cat)/K(M)) than the analogous Y473F mutation of prolyl oligopeptidase. Kinetic studies on a series of cocE mutants both validate the proposed mechanism, and reveal the relative contributions of active site residues toward substrate recognition and catalysis. Inspired by the anionic binding pocket of the cocaine binding antibody GNC92H2, we found that a Q55E mutation within the active site of cocE results in a modest (2-fold) improvement in K(M), but a 14-fold loss of k(cat). The pH rate profile of cocE was fit to the ionization of two groups (pK(a1) = 7.7; pK(a2) = 10.4) that likely represent titration of H287 and Y44, respectively. We also describe the crystal structures of both S117A and Y44F mutants of cocE. Finally, urea denaturation studies of cocE by fluorescence and circular dichroism show two unfolding transitions (0.5-0.6 M and 3.2-3.7 M urea), with the first transition likely representing pertubation of the active site.  相似文献   

17.
In previous studies we showed that the plasma protein histidine-rich glycoprotein (HRG) binds strongly to pooled human IgG. In the present work myeloma proteins consisting of different human IgG subclasses were examined for their ability to interact with human HRG. Using an IAsys optical biosensor we found initially that IgG subclasses differ substantially in their affinity of interaction with HRG. However, the most striking finding was the observation that the kinetics of the HRG interaction was dramatically affected by whether the IgG subclasses contained the kappa or lambda light (L)-chains. Thus, the on-rate for the binding of HRG to the kappa L-chain containing IgG1 and IgG2 (IgG1kappa and IgG2kappa) was approximately 4- and approximately 10-fold faster than that for the binding of HRG to lambda L-chain containing IgG1 and IgG2 (IgG1lambda and IgG2lambda), respectively, with the dissociation constants (K(d)) in the range 3-5 nM and 112-189 nM for the kappa and lambda isoforms, respectively. In contrast, the on-rate for the binding of HRG to IgG3kappa and IgG4kappa was found to be 9- and 20-fold slower than that for the binding of HRG to IgG3lambda and IgG4lambda, respectively, with the K(d) in the range 147-268 nM and 96-109 nM for the kappa and lambda isoforms, respectively. The binding of HRG to immunoglobulins containing the kappa L-chain (particularly IgG1kappa) was generally potentiated in the presence of a physiological concentration (20 microM) of Zn(2+) (K(d) decreased to 0.60 +/- 0.01 for IgG1kappa), but Zn(2+) had no effect or slightly inhibited the binding of HRG to immobilized IgG subclasses possessing the lambda L-chain. Interestingly, HRG also bound differentially to Bence Jones (BJ) proteins containing kappa and lambda L-chains, with HRG having a 14-fold lower K(d) for BJkappa than for BJlambda when 20 microM Zn(2+) was present. HRG also bound to IgM (IgMkappa), but the affinity of this interaction (K(d) approximately 1.99 +/- 0.05 microM) was markedly lower than the interaction with IgG, and the affinity was actually decreased 4-fold in the presence of Zn(2+). The results demonstrate that both the heavy (H)- and L-chain type have a profound effect on the binding of HRG to different IgG subclasses and provide the first evidence of a functional difference between the kappa and lambda L-chains of immunoglobulins.  相似文献   

18.
Lys(63)-linked polyubiquitin (poly-Ub) chains appear to play a nondegradative signaling and/or recruitment role in a variety of key eukaryotic cellular processes, including NF-kappaB signal transduction and DNA repair. A protein heterodimer composed of a catalytically active ubiquitin-conjugating enzyme (Ubc13) and its homologue (Mms2 or Uev1a) forms a catalytic scaffold upon which a noncovalently associated acceptor Ub and thiolester-linked donor Ub are oriented such that Lys(63)-linked poly-Ub chain synthesis is facilitated. In this study, we have used (1)H-(15)N nuclear magnetic resonance spectroscopy, in combination with isothermal titration calorimetry, to determine the thermodynamics and kinetics of the interactions between various components of the Lys(63)-linked poly-Ub conjugation machinery. Mms2 and Uev1a interact in vitro with acceptor Ub to form 1/1 complexes with macroscopic dissociation constants of 98 +/- 15 and 213 +/- 14 microM, respectively, and appear to bind Ub in a similar fashion. Interestingly, the Mms2.Ubc13 heterodimer associates with acceptor Ub in a 1/1 complex and binds with a dissociation constant of 28 +/- 6 microM, significantly stronger than the binding of Mms2 alone. Furthermore, a dissociation constant of 49 +/- 7 nM was determined for the interaction between Mms2 and Ubc13 using isothermal titration calorimetry. In connection with previous structural studies for this system, the thermodynamics and kinetics of acceptor Ub binding to the Mms2.Ubc13 heterodimer described in detail in this study will allow for a more thorough rationalization of the mechanism of formation of Lys(63)-linked poly-Ub chains.  相似文献   

19.
A fluorescent derivative of paclitaxel, 3'-N-m-aminobenzamido-3'-N-debenzamidopaclitaxel (N-AB-PT), has been prepared in order to probe paclitaxel-microtubule interactions. Fluorescence spectroscopy was used to quantitatively assess the association of N-AB-PT with microtubules. N-AB-PT was found equipotent with paclitaxel in promoting microtubule polymerization. Paclitaxel and N-AB-PT underwent rapid exchange with each other on microtubules assembled from GTP-, GDP-, and GMPCPP-tubulin. The equilibrium binding parameters for N-AB-PT to microtubules assembled from GTP-tubulin were derived through fluorescence titration. N-AB-PT bound to two types of sites on microtubules (K(d1) = 61 +/- 7.0 nM and K(d2) = 3.3 +/- 0.54 microM). The stoichiometry of each site was less than one ligand per tubulin dimer in the microtubule (n(1) = 0.81 +/- 0.03 and n(2) = 0.44 +/- 0.02). The binding experiments were repeated after exchanging the GTP for GDP or for GMPCPP. It was found that N-AB-PT bound to a single site on microtubules assembled from GDP-tubulin with a dissociation constant of 2.5 +/- 0.29 microM, and that N-AB-PT bound to a single site on microtubules assembled from GMPCPP-tubulin with a dissociation constant of 15 +/- 4.0 nM. It therefore appears that microtubules contain two types of binding sites for paclitaxel and that the binding site affinity for paclitaxel depends on the nucleotide content of tubulin. It has been established that paclitaxel binding does not inhibit GTP hydrolysis and microtubules assembled from GTP-tubulin in the presence of paclitaxel contain almost exclusively GDP at the E-site. We propose that although all the subunits of the microtubule at steady state are the same "GDP-tubulin-paclitaxel", they are formed through two paths: paclitaxel binding to a tubulin subunit before its E-site GTP hydrolysis is of high affinity, and paclitaxel binding to a tubulin subunit containing hydrolyzed GDP at its E-site is of low affinity.  相似文献   

20.
The thermodynamics of coenzyme binding to human cytochrome P450 reductase (CPR) and its isolated FAD-binding domain have been studied by isothermal titration calorimetry. Binding of 2',5'-ADP, NADP(+), and H(4)NADP, an isosteric NADPH analogue, is described in terms of the dissociation binding constant (K(d)), the enthalpy (DeltaH(B)) and entropy (TDeltaS(B)) of binding, and the heat capacity change (DeltaC(p)). This systematic approach allowed the effect of coenzyme redox state on binding to CPR to be determined. The recognition and stability of the coenzyme-CPR complex are largely determined by interaction with the adenosine moiety (K(d2)(')(,5)(')(-ADP) = 76 nM), regardless of the redox state of the nicotinamide moiety. Similar heat capacity change (DeltaC(p)) values for 2',5'-ADP (-210 cal mol(-)(1) K(-)(1)), NADP(+) (-230 cal mol(-)(1) K(-)(1)), and H(4)NADP (-220 cal mol(-)(1) K(-)(1)) indicate no significant contribution from the nicotinamide moiety to the binding interaction surface. The coenzyme binding stoichiometry to CPR is 1:1. This result validates a recently proposed one-site kinetic model [Daff, S. (2004) Biochemistry 43, 3929-3932] as opposed to a two-site model previously suggested by us [Gutierrez, A., Lian, L.-Y., Wolf, C. R., Scrutton, N. S., and Roberts, C. G. K. (2001) Biochemistry 40, 1964-1975]. Calorimetric studies in which binding of 2',5'-ADP to CPR (TDeltaS(B) = -13400 +/- 200 cal mol(-)(1), 35 degrees C) was compared with binding of the same ligand to the isolated FAD-binding domain (TDeltaS(B) = -11200 +/- 300 cal mol(-)(1), 35 degrees C) indicate that the number of accessible conformational substates of the protein increases upon 2',5'-ADP binding in the presence of the FMN-binding domain. This pattern was consistently observed along the temperature range that was studied (5-35 degrees C). This contribution of coenzyme binding energy to domain dynamics in CPR agrees with conclusions from previous temperature-jump studies [Gutierrez, A., Paine, M., Wolf, C. R., Scrutton, N. S., and Roberts, G. C. K. (2002) Biochemistry 41, 4626-4637]. A combination of calorimetry and stopped-flow spectrophotometry kinetics experiments showed that this linkage between coenzyme binding energetics and diffusional domain motion impinges directly on the molecular recognition of cytochrome c by CPR. Single-turnover reduction of cytochrome c by CPR (k(max) = 15 s(-)(1), K(d) = 37 microM) is critically coupled to coenzyme binding through ligand-induced motions that enable the FMN-binding domain to overcome a kinetically unproductive conformation. This is remarkable since the FMN-binding domain is not directly involved in coenzyme binding, the NADP(H) binding site being fully contained in the FAD-binding domain. Sequential rapid mixing measurements indicate that harnessing of coenzyme binding energy to the formation of a kinetically productive CPR-cytochrome c complex is a highly synchronized event. The inferred half-time for the decay of this productive conformation (tau(50)) is 330 +/- 70 ms only. Previously proposed structural and kinetic models are discussed in light of these findings.  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