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.
The thermal depolymerization procedure of Stephens (1970. J. Mol. Biol. 47:353) has been employed for solubilization of Strongylocentrotus purpuratus sperm tail outer doublet microtubules with the use of a buffer during solubilization which is of optimal pH and ionic strength for the preservation of colchicine binding activity of chick embryo brain tubulin. Colchicine binding values were corrected for first-order decay during heat solubilization at 50°C (t½ = 5.4 min) and incubation with colchicine at 37°C in the presence of vinblastine sulfate (t½ = 485 min). The colchicine binding properties of heat-solubilized outer doublet tubulin were qualitatively identical with those of other soluble forms of tubulin. The solubilized tubulin (mol wt, 115,000) bound 0.9 ± 0.2 mol of colchicine per mol of tubulin, with a binding constant of 6.3 x 105 liters/mol at 37°C. The colchicine binding reaction was both time and temperature dependent, and the binding of colchicine was prevented in a competitive manner by podophyllotoxin (Ki = 1.3 x 10-6 M). The first-order decay of colchicine binding activity was substantially decreased by the addition of the vinca alkaloids, vinblastine sulfate or vincristine sulfate, thus demonstrating the presence of a vinca alkaloid binding site(s) on the outer doublet tubulin. Tubulin contained within the assembled microtubules did not decay. Intact outer doublet microtubules bound less than 0.001 mol of colchicine per mol of tubulin contained in the microtubules, under conditions where soluble tubulin would have bound 1 mol of colchicine per mol of tubulin (saturating concentration of colchicine, no decay of colchicine binding activity). The presence of colchicine had no effect on the rate of solubilization of outer doublet microtubules during incubation at 37°C. Therefore, the colchicine binding site on tubulin is blocked (not available to bind colchicine) when the tubulin is in the assembled outer doublet microtubules.  相似文献   

2.
In kinesin X-ray crystal structures, the N-terminal region of the α-1 helix is adjacent to the adenine ring of the bound nucleotide, while the C-terminal region of the helix is near the neck-linker (NL). Here, we monitor the displacement of the α-1 helix within a kinesin monomer bound to microtubules (MTs) in the presence or absence of nucleotides using site-directed spin labeling EPR. Kinesin was doubly spin-labeled at the α-1 and α-2 helices, and the resulting EPR spectrum showed dipolar broadening. The inter-helix distance distribution showed that 20% of the spins have a peak characteristic of 1.4–1.7 nm separation, which is similar to what is predicted from the X-ray crystal structure, albeit 80% were beyond the sensitivity limit (>2.5 nm) of the method. Upon MT binding, the fraction of kinesin exhibiting an inter-helix distance of 1.4–1.7 nm in the presence of AMPPNP (a non-hydrolysable ATP analog) and ADP was 20% and 25%, respectively. In the absence of nucleotide, this fraction increased to 40–50%. These nucleotide-induced changes in the fraction of kinesin undergoing displacement of the α-1 helix were found to be related to the fraction in which the NL undocked from the motor core. It is therefore suggested that a shift in the α-1 helix conformational equilibrium occurs upon nucleotide binding and release, and this shift controls NL docking onto the motor core.  相似文献   

3.
Pyridostigmine bromide (PB), an inhibitor of acetylcholinesterase, has been used as a prophylactic for nerve gas poisoning. N,N-diethyl-m-toluamide (DEET) is the active ingredient in most insect repellents and is thought to interact synergistically with PB. Since PB can inhibit the binding of organophosphates to tubulin and since organophosphates inhibit microtubule assembly, we decided to examine the effects of PB and DEET on microtubule assembly as well as their interactions with tubulin, the subunit protein of microtubules. We found that PB binds to tubulin with an apparent K d of about 60 M. PB also inhibits microtubule assembly in vitro, although at higher concentrations PB induces formation of tubulin aggregates of high absorbance. Like PB, DEET is a weak inhibitor of microtubule assembly and also induces formation of tubulin aggregates. Many tubulin ligands stabilize the conformation of tubulin as measured by exposure of sulfhydryl groups and hydrophobic areas and stabilization of colchicine binding. PB appears to have very little effect on tubulin conformation, and DEET appears to have no effect. Neither compound interferes with colchicine binding to tubulin. Our results raise the possibility that PB and DEET may exert some of their effects in vivo by interfering with microtubule assembly or function, although high intracellular levels of these compounds would be required.  相似文献   

4.
5.
The study aimed to define the key factors involved in the modulation of actinomycete mannanases. We focused on the roles of carbohydrate-binding modules (CBMs) and bivalent ions. To investigate the effects of these factors, two actinomycete mannanase genes were cloned from Streptomyces thermoluteus (StManII) and Streptomyces lividans (SlMan). CBMs fused to mannanase catalytic domains do not affect the thermal stability of the proteins. CBM2 of StManII increased the catalytic efficiency toward soluble-mannan and insoluble-mannan by 25%–36%, and CBM10 of SlMan increased the catalytic efficiency toward soluble-mannan by 40%–50%. Thermal stability of wild-type and mutant enzymes was enhanced by calcium and manganese. Thermal stability of SlMandC was also slightly enhanced by magnesium. These results indicated that bivalent ion-binding site responsible for thermal stability was in the catalytic domains. Thermal stability of mannanase differed in the kinds of bivalent ions. Isothermal titration calorimetry revealed that the catalytic domain of StManII bound bivalent ions with a Ka of 5.39 ± 0.45 × 103–7.56 ± 1.47 × 103 M− 1, and the catalytic domain of SlMan bound bivalent ions with a Ka of 1.06 ± 0.34 × 103–3.86 ± 0.94 × 103 M− 1. The stoichiometry of these bindings was consistent with one bivalent ion-binding site per molecule of enzyme. Circular dichroism spectrum revealed that the presence of bivalent ions induced changes in the secondary structures of the enzymes. The binding of certain bivalent ion responsible for thermal stability was accompanied by a different conformational change by each bivalent ion. Actinomycete mannanases belong to GHF5 which contained various hemicellulases; therefore, the information obtained from mannanases applies to the other enzymes.  相似文献   

6.
A new photoaffinity analogue of colchicine, (2-nitro-4-azidophenyl)deacetylcolchicine (NAPDAC), bound to two classes of sites on bovine renal tubulin and photolabeled both the alpha- and beta-subunits. The apparent Ki for the photoaffinity analogue was 1.40 +/- 0.17 microM (mean +/- SD, n = 3) as measured by competition with [3H] colchicine. Values of the apparent KdS for the two sites, as measured by the direct binding of the [3H]NAPDAC to tubulin, were 0.48 +/- 0.11 microM and 11.6 +/- 3.5 microM (mean +/- SD, n = 6), and the corresponding stoichiometries of binding of the two sites were 0.25 +/- 0.06 and 1.3 +/- 0.4 mol/mol of tubulin (mean +/- SD, n = 6). NAPDAC was a potent inhibitor of microtubule formation as detected by electron microscopy. When tubulin was photolabeled with NAPDAC at 25 degrees C, 15 +/- 3 mol % (mean +/- SD, n = 6) of the [3H]NAPDAC was covalently bound to the alpha-subunit, and 67 +/- 9 mol % (mean +/- SD, n = 6) was covalently bound to the beta-subunit. Since NAPDAC is a mixture of two interconvertible diastereomers, the photoincorporation of each was also examined. One diastereomer photolabeled both alpha- and beta-tubulin; however, the other did not significantly photolabel either subunit. Tubulin photolabeled with NAPDAC (1:1 mole ratio) exhibited a 23% decrease in colchicine binding. Preblocking and prephotolysis experiments with colchicine, NAPDAC, or ANPAH-CLC [Williams et al. (1985) J. Biol. Chem. 260, 13794-13802] provided evidence for conformational changes in tubulin upon colchicine binding. Peptide maps of [3H]NAPDAC-labeled alpha- and beta-tubulin, using Staphylococcus aureus V8 protease, demonstrated the presence of NAPDAC in one peptide of the alpha-subunit and in five peptides of the beta-subunit as detected by autoradiography. NAPDAC provides the first direct evidence for two colchicine binding sites on tubulin.  相似文献   

7.
Reduction of Complex I (NADH:ubiquinone oxidoreductase I) from Escherichia coli by NADH was investigated optically by means of an ultrafast stopped-flow approach. A locally designed microfluidic stopped-flow apparatus with a low volume (0.2 μl) but a long optical path (10 mm) cuvette allowed measurements in the time range from 270 μs to seconds. The data acquisition system collected spectra in the visible range every 50 μs. Analysis of the obtained time-resolved spectral changes upon the reaction of Complex I with NADH revealed three kinetic components with characteristic times of < 270 μs, 0.45–0.9 ms and 3–6 ms, reflecting reduction of different FeS clusters and FMN. The rate of the major (τ = 0.45–0.9 ms) component was slower than predicted by electron transfer theory for the reduction of all FeS clusters in the intraprotein redox chain. This delay of the reaction was explained by retention of NAD+ in the catalytic site. The fast optical changes in the time range of 0.27–1.5 ms were not altered significantly in the presence of 10-fold excess of NAD+ over NADH. The data obtained on the NuoF E95Q variant of Complex I shows that the single amino acid replacement in the catalytic site caused a strong decrease of NADH binding and/or the hydride transfer from bound NADH to FMN.  相似文献   

8.
The inhibition of the polymerization of tubulin from cultured cells of rose (Rosa. sp. cv. Paul's scarlet) by colchicine and the binding of colchicine to tubulin were examined in vitro and compared with data obtained in parallel experiments with bovine brain tubulin. Turbidimetric measurements of taxol-induced polymerization of rose microtubules were found to be sensitive and semiquantitative at low tubulin concentrations, and to conform to some of the characteristics of a nucleation and condensation-polymerization mechanism for assembly of filamentous helical polymers. Colchicine inhibited the rapid phase of polymerization at 24°C with an apparent inhibition constant (K i) of 1.4·10-4 M for rose tubulin and an apparent K i=8.8·10-7 M for brain tubulin. The binding of [3H]colchicine to rose tubulin to form tubulin-colchicine complex was mildly temperature-dependent and slow, taking 2–3 h to reach equilibrium at 24°C, and was not affected by vinblastine sulfate. The binding of [3H]colchicine to rose tubulin was saturable and Scatchard analysis indicated a single class of low-affinity binding sites having an apparent affinity constant (K) of 9.7·102 M-1 and an estimated molar binding stoichiometry (r) of 0.47 at 24°C. The values for brain tubulin were K=2.46·106 M-1 and r=0.45 at 37°C. The binding of [3H]colchicine to rose tubulin was inhibited by excess unlabeled colchicine, but not by podophyllotoxin or tropolone. The data demonstrate divergence of the colchicine-binding sites on plant and animal tubulins and indicate that the relative resistance of plant microtubule polymerization to colchicine results from a low-affinity interaction of colchicine and tubulin.Abbreviations MT microtubule - TC tubulin-colchicine complex  相似文献   

9.
The tubulins of Antarctic fishes possess adaptations that favor microtubule formation at low body temperatures (Detrich et al.: Biochemistry 28:10085-10093, 1989). To determine whether some of these adaptations may be present in a domain of tubulin that participates directly or indirectly in lateral contact between microtubule protofilaments, we have examined the energetics of the binding of colchicine, a drug thought to bind to such a site, to pure brain tubulins from an Antarctic fish (Notothenia gibberifrons) and from a mammal (the cow, Bos taurus). At temperatures between 0 and 20 degrees C, the affinity constants for colchicine binding to the fish tubulin were slightly smaller (1.5-2.6-fold) than those for bovine tubulin. van't Hoff analysis showed that the standard enthalpy changes for colchicine binding to the two tubulins were comparable (delta H degrees = +10.6 and +7.4 kcal mol-1 for piscine and bovine tubulins, respectively), as were the standard entropy changes (delta S degrees = +61.3 eu for N. gibberifrons tubulin, +51.2 eu for bovine tubulin). At saturating concentrations of the ligand, the maximal binding stoichiometry for each tubulin was approximately 1 mol colchicine/mol tubulin dimer. The data indicate that the colchicine-binding sites of the two tubulins are similar, but probably not identical, in structure. The apparent absence of major structural modifications at the colchicine site suggests that this region of tubulin is not involved in functional adaptation for low-temperature polymerization. Rather, the colchicine site of tubulin may have been conserved evolutionarily to serve in vivo as a receptor for endogenous molecules (i.e., "colchicine-like" molecules or MAPs) that regulate microtubule assembly.  相似文献   

10.
11.
Thirty seven N-(5-methoxyphenyl)-4-methoxybenzenesulphonamide with methoxy or/and bromo substitutions (series 1-4) and with different substituents on the sulphonamide nitrogen have been synthesised. 21 showed sub-micromolar cytotoxicity against HeLa and HT-29 human tumour cell lines, and were particularly effective against MCF7. The most potent series has 2,5-dimethoxyanilines, especially the 4-brominated compounds 23–25. The active compounds inhibit microtubular protein polymerisation at micromolar concentrations, thus pointing at tubulin as the target. Co-treatment with the MDR inhibitor verapamil suggests that they are not MDR substrates. Compound 25 showed nanomolar antiproliferative potency. It severely disrupts the microtubule network in cells and arrests cells at the G2/M cell-cycle phase, thus confirming tubulin targeting. 25 triggered apoptotic cell death, and induced autophagy. Docking studies suggest binding in a distinct way to the colchicine site. These compounds are promising new antitumor agents acting on tubulin.  相似文献   

12.
Studies investigating the association between interleukin-13 (IL-13) single nucleotide polymorphism (SNP) rs20541 and allergic rhinitis (AR) risk have reported conflicting results. The aim of the present study was to conduct a meta-analysis assessing the possible association of IL-13 SNP rs20541 with AR risk. Eight studies were included in the present meta-analysis (2153 cases and 3931 controls). The combined results based on all studies showed that IL-13 SNP rs20541 was associated with increased AR risk (Gln versus Arg: odds ratio (OR) = 1.18, 95% confidence interval (CI) = 1.08–1.30; Gln/Gln versus Arg/Arg: OR = 1.52, 95% CI = 1.20–1.92; Arg/Gln + Gln/Gln versus Arg/Arg: OR = 1.19, 95% CI = 1.06–1.33; Gln/Gln versus Arg/Gln + Arg/Arg: OR = 1.42, 95% CI = 1.13–1.79). When stratifying for race, IL-13 SNP rs20541 exhibited increased AR risk in Asians (Gln versus Arg: OR = 1.20, 95% CI = 1.06–1.36; Gln/Gln versus Arg/Arg: OR = 1.57, 95% CI = 1.17–2.12; Arg/Gln + Gln/Gln versus Arg/Arg: OR = 1.22, 95% CI = 1.04–1.44; Gln/Gln versus Arg/Gln + Arg/Arg: OR = 1.45, 95% CI = 1.09–1.93), while no significant association was detected in Caucasians (Gln versus Arg: OR = 1.28, 95% CI = 0.93 ~ 1.78; Gln/Gln versus Arg/Arg: OR = 1.42, 95% CI = 0.96–2.11; Arg/Gln + Gln/Gln versus Arg/Arg: OR = 1.35, 95% CI = 0.89–2.05; Gln/Gln versus Arg/Gln + Arg/Arg: OR = 1.37, 95% CI = 0.93–2.02). This meta-analysis supported that IL-13 SNP rs20541 was associated with AR, particularly in Asians.  相似文献   

13.
An engineered monomeric chorismate mutase (mMjCM) has been found to combine high catalytic activity with the characteristics of a molten globule. To gain insight into the dramatic structural changes that accompany binding of a transition-state analog, we examined mMjCM by isothermal calorimetry and compared it with its dimeric parent protein, MjCM (CM from Methanococcus jannaschii), a thermostable and conventionally folded enzyme. As expected for a ligand-induced ordering process, there is a large entropic penalty for binding to the monomer relative to the dimer (− TΔΔS = 5.1 ± 0.5 kcal/mol, at 20 °C). However, this unfavorable entropy term is largely offset by enthalpic gains (ΔΔH = − 3.5 ± 0.4 kcal/mol), presumably arising from tightening of non-covalent interactions throughout the monomeric complex. Stopped-flow kinetic measurements further reveal that the catalytic molten globule binds and releases ligands significantly faster than its natural counterpart, demonstrating that partial structural disorder can speed up molecular recognition. These results illustrate how structural plasticity may strongly perturb the thermodynamics and kinetics of transition-state recognition while negligibly affecting catalytic efficiency.  相似文献   

14.
The interactions of tubulin with colchicine analogues in which the tropolone methyl ether ring had been transformed into a p-carbomethoxybenzene have been characterized. The analogues were allocolchicine (ALLO) and 2,3,4-trimethoxy-4'-carbomethoxy-1,1'-biphenyl (TCB), the first being transformed colchicine and the second transformed colchicine with ring B eliminated. The binding of both analogues has been shown to be specific for the colchicine binding site on tubulin by competition with colchicine and podophyllotoxin. Both analogues bind reversibly to tubulin with the generation of ligand fluorescence. The binding of ALLO is slow, the fluorescence reaching a steady state in the same time span as colchicine; that of TCB is rapid. The displacement of ALLO by podophyllotoxin proceeds with a half-life of ca. 40 min. Binding isotherms generated from gel filtration and fluorescence measurements have shown that both analogues bind to tubulin with a stoichiometry of 1 mol of analogue/mol of alpha-beta tubulin. The equilibrium binding constants at 25 degrees C have been found to be (9.2 +/- 2.5) x 10(5) M-1 for ALLO and (1.0 +/- 0.2) X 10(5) M-1 for TCB. Binding of both analogues was accompanied by quenching of protein fluorescence, perturbation of the far-ultraviolet circular dichroism of tubulin, and induction of the tubulin GTPase activity, similarly to colchicine binding. Both inhibited microtubule assembly in vitro, ALLO substoichiometrically, and both induced the abnormal cooperative polymerization of tubulin, which is characteristic of the tubulin-colchicine complex. Analysis in terms of the simple bifunctional ligand binding mechanism developed for colchicine [Andreu, J.M., & Timasheff, S.N. (1982) Biochemistry 21, 534-543] and comparison with the binding of the colchicine two-ring analogue, 2-methoxy-5-(2,3,4-trimethoxyphenyl)-2,4,6-cycloheptatrien-1-one [Andreu, J. M., Gorbunoff, M. J., Lee, J. C., & Timasheff, S. N. (1984) Biochemistry 23, 1742-1752], have shown that transformation of the tropolone methyl ether part of colchicine into p-carbomethoxybenzene weakens the standard free energy of binding to tubulin by 1.4 +/- 0.1 kcal/mol, while elimination of ring B weakens it by 1.0 +/- 0.1 kcal/mol. The roles of rings C and B of colchicine in the thermodynamic and kinetic mechanisms of binding to tubulin were analyzed in terms of these findings.  相似文献   

15.
[125I]IodoDPA-713 [125I]1, which targets the translocator protein (TSPO, 18 kDa), was synthesized in seven steps from methyl-4-methoxybenzoate as a tool for quantification of inflammation in preclinical models. Preliminary in vitro autoradiography and in vivo small animal imaging were performed using [125I]1 in a neurotoxicant-treated rat and in a murine model of lung inflammation, respectively. The radiochemical yield of [125I]1 was 44 ± 6% with a specific radioactivity of 51.8 GBq/μmol (1400 mCi/μmol) and >99% radiochemical purity. Preliminary studies showed that [125I]1 demonstrated increased specific binding to TSPO in a neurotoxicant-treated rat and increased radiopharmaceutical uptake in the lungs of an experimental inflammation model of lung inflammation. Compound [125I]1 is a new, convenient probe for preclinical studies of TSPO activity.  相似文献   

16.
The binding affinity of the two substrate–water molecules to the water-oxidizing Mn4CaO5 catalyst in photosystem II core complexes of the extremophilic red alga Cyanidioschyzon merolae was studied in the S2 and S3 states by the exchange of bound 16O-substrate against 18O-labeled water. The rate of this exchange was detected via the membrane-inlet mass spectrometric analysis of flash-induced oxygen evolution. For both redox states a fast and slow phase of water-exchange was resolved at the mixed labeled m/z 34 mass peak: kf = 52 ± 8 s− 1 and ks = 1.9 ± 0.3 s− 1 in the S2 state, and kf = 42 ± 2 s− 1 and kslow = 1.2 ± 0.3 s− 1 in S3, respectively. Overall these exchange rates are similar to those observed previously with preparations of other organisms. The most remarkable finding is a significantly slower exchange at the fast substrate–water site in the S2 state, which confirms beyond doubt that both substrate–water molecules are already bound in the S2 state. This leads to a very small change of the affinity for both the fast and the slowly exchanging substrates during the S2 → S3 transition. Implications for recent models for water-oxidation are briefly discussed.  相似文献   

17.
RIG-I (retinoic acid-inducible gene I) is an essential cytosolic pathogen recognition receptor that binds to a variety of viral RNA or DNA to induce type I interferons. In the present study, insert–deletion polymorphisms in promoter and introns of CiRIG-I (Ctenopharyngodon idella RIG-I) were explored, their associations with resistance/susceptibility to grass carp reovirus (GCRV) were analyzed. To this end, genomic sequence of CiRIG-I gene was obtained, and twenty pairs of primers were prepared for the detection of insert–deletion polymorphisms. Five insert–deletion mutations were found, a 2-bp mutation and an 8-bp mutation existed in the promoter and other three sizes in 74 bp, 146 bp and 53 bp were sited in the intron 8. After a challenge experiment, only the genotype and allele of − 740 insert–deletion mutation in the promoter and allele of 6804 insert–deletion mutation were significantly associated with resistance/susceptibility to GCRV among the five mutations (P < 0.05). To further identify this correlation, another independent challenge test was carried out. The result revealed that the cumulative mortality in ins/ins genotype individuals (43.75%) at − 740 insert–deletion mutation was significantly lower than that in ins/del (72.09%) and del/del (74.19%) genotypes (P < 0.05). Linkage disequilibrium and haplotype analysis showed 6610 insert–deletion mutation and 6804 insert–deletion mutation were linkage disequilibrium. The haplotype ins–ins (6610ins–6804ins) was significantly susceptible to GCRV, and ins–del (6610ins–6804del) was significantly resistant to GCRV (P < 0.05). Those could be potential gene markers for the future molecular selection of strains that are resistant to GCRV.  相似文献   

18.
Two modes of high-speed counter-current chromatography (HSCCC) were successfully applied to the separation of alkaloids from crude extract of Nelumbo nucifera leaves. The conventional HSCCC separations were performed with a two-phase solvent system composed of tetrachloromethane–CHCl3–methanol–0.1 M HCl at a volume ratio of 1:3:3:2 (v/v/v/v), and 120 mg crude extract could be successfully separated. pH-Zone-refining CCC was performed with a two-phase solvent system composed of petroleum ether (60–90 °C)–ethyl acetate–methanol–water (5:5:2:8, v/v/v/v) where triethylamine (10 mM) was added to the upper organic stationary phase as a retainer and hydrochloric acid (5 mM) to the aqueous mobile phase as an eluent. From 4.0 g of the crude extract, 120 mg N-nornuciferine, 1020 mg nuciferine and 96 mg roemerine were obtained in a single run each with a purity of over 98% as determined by HPLC. The structures of the isolated compounds were identified by ESI-MS, 1H NMR and 13C NMR.  相似文献   

19.
Gupta K  Panda D 《Biochemistry》2002,41(43):13029-13038
The dietary flavonoid quercetin has a broad range of biological activities, including potent antitumor activity against several types of tumors. Recently, it has been shown that quercetin inhibits cancer cells proliferation by depleting cellular microtubules and perturbing cellular microtubule functions. However, the direct interactions of quercetin with tubulin and microtubules have not been examined so far. Here, we found that quercetin inhibited polymerization of microtubules and depolymerized microtubules made from purified tubulin in vitro. The binding of quercetin with tubulin was studied using quercetin fluorescence and intrinsic tryptophan fluorescence of tubulin. Quercetin bound to tubulin at a single site with a dissociation constant of 5-7 microM, and it specifically inhibited colchicine binding to tubulin but did not bind at the vinblastine site. In addition, quercetin perturbed the secondary structure of tubulin, and the binding of quercetin stimulated the intrinsic GTPase activity of soluble tubulin. Further, quercetin stabilized tubulin against decay and protected two cysteine residues of tubulin toward chemical modification by 5,5'-dithiobis-2-nitrobenzoic acid. Our data demonstrated that the binding of quercetin to tubulin induces conformational changes in tubulin and a mechanism through which quercetin could perturb microtubule polymerization dynamics has been proposed. The data suggest that quercetin inhibits cancer cells proliferation at least in part by perturbing microtubule functions through tubulin binding.  相似文献   

20.
Understanding the structure, folding, and interaction of membrane proteins requires experimental tools to quantify the association of transmembrane (TM) helices. Here, we introduce isothermal titration calorimetry (ITC) to measure integrin αIIbβ3 TM complex affinity, to study the consequences of helix–helix preorientation in lipid bilayers, and to examine protein-induced lipid reorganization. Phospholipid bicelles served as membrane mimics. The association of αIIbβ3 proceeded with a free energy change of − 4.61 ± 0.04 kcal/mol at bicelle conditions where the sampling of random helix–helix orientations leads to complex formation. At bicelle conditions that approach a true bilayer structure in effect, an entropy saving of > 1 kcal/mol was obtained from helix–helix preorientation. The magnitudes of enthalpy and entropy changes increased distinctly with bicelle dimensions, indicating long-range changes in bicelle lipid properties upon αIIbβ3 TM association. NMR spectroscopy confirmed ITC affinity measurements and revealed αIIbβ3 association and dissociation rates of 4500 ± 100 s− 1 and 2.1 ± 0.1 s− 1, respectively. Thus, ITC is able to provide comprehensive insight into the interaction of membrane proteins.  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