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.
Thermodynamic studies have been made on the effect of Cr+3 on the conformation and structure of bovine β lactoglobulin-A, (Blg-A) in 50 mM sodium chloride solution at 27°C using isothermal titration calorimetry (ITC), circular dichroism (CD) and fluorescence spectroscopy. There is a set of six identical and independent binding sites for Cr+3 by a dissociation binding constant of 124 μM and the molar enthalpy of binding −17.8 kJ/mol. Circular dichroism studies do not show any significant change in the secondary structure of the protein after the binding of chromium ion on the Blg-A. Fluorescence spectroscopy studies do not show any considerable change in the tertiary structure of Blg-A due to the increasing of Cr+3 in low concentration. However, occupation of fourth and fifth binding sites for chromium ions induce partially unfolding in the tertiary structure of the protein resulted from solvent – exposed hydrophobic patches on BLG-A.  相似文献   

2.
3′-azido-3′-deoxythymidine (AZT) is the first effective drug used clinically for the treatment of human immunodeficiency virus (HIV) infection. The drug interactions with DNA and protein are associated with its mechanism of action in vivo. This study was designed to examine the interaction of AZT with the Na,K-dependent adenosine triphosphatase (Na,K-ATPase) in H2O and D2O solutions at physiological pH using drug concentration of 0.1 μM to 1 mM and final protein concentration of 0.5 to 1 mg/mL. Ultraviolet absorption and Fourier transform infrared difference spectroscopy with its self-deconvolution second-derivative resolution enhancement, and curve-fitting procedures were used to characterize the drug-binding mode, the drug-binding constant, and the effects of drug interaction on the protein secondary structure Spectroscopic evidence showed that at low drug concentration (0.1 μM), AZT binds (H-bonding) mainly to the polypeptide C=O and C−N groups with two binding constants of K1=5.3×105 M −1 and K2=9.8×103 M −1. As drug content increased, AZT-lipid complex prevailed. At a high drug concentration (1 mM), drug binding resulted in minor protein secondary structural changes from that of the α-helix 19.8%; β-pleated 25.6%; turn 9.1%; β-antiparallel 7.5% and random 38%, in the free Na,K-ATPase to that of the α-helix 19%; β-pleated 21.1%; turn 10.1%; β-antiparallel 8.8% and random 41%, in the AZT-ATPase complexes.  相似文献   

3.
Calorimetric and Binding Dissections of HSA Upon Interaction with Bilirubin   总被引:1,自引:0,他引:1  
The interactions between bilirubin and human serum albumin (HSA) were studied by isothermal titration calorimetry (ITC) and UV–vis spectrophotometry at 27°C in 100 mM phosphate buffer pH 7.4 containing 1 mM EDTA. The biphasic shape of the HSA–bilirubin binding curve depicted the existence of two bilirubin binding sets on the HSA structure which had distinct binding interactions. Each binding set contained one or more bilirubin binding site. The first binding set at subdomain IIA included one binding site and had a more hydrophobic microenvironment than the other two binding sites in the second bilirubin binding set (subdomain IIIA). With our method of analysis, the calculated dissociation constant of the first binding site is 1.28×10−6 M and 4.80×10−4 M for the second and third binding sites. Here, the typical Boltzmann’s equation was used with a new approach to calculate the dissociation constants as well as the standard free energy changes for the HSA–bilirubin interactions. Interestingly, our calculations obtained using the Wyman binding potential theory confirmed that our analysis method had been correct (especially for the second binding phase). The molar extinction coefficient determined for the first bound bilirubin molecule depicted that the bilirubin molecules (in low concentrations) should interact with the nonpolar microenvironment of the first high affinity binding site. Binding of the bilirubin molecules to the first binding site was endothermic (ΔHo>0) and occurred through the large increase in the binding entropy established when the hydrophobic bilirubin molecules escaped from their surrounding polar water molecules and into the hydrophobic medium of the first binding site. On the other hand, the calculated molar extinction coefficient illustrated that the microenvironment of the second binding set (especially for the third binding site) was less hydrophobic than the first one but still more hydrophobic than the buffer medium. The binding of the third bilirubin molecule to the HSA molecule was established more through exothermic (electrostatic) interactions.  相似文献   

4.
Nowadays, understanding of interface between protein and drugs has become an active research area of interest. These types of interactions provide structural guidelines in drug design with greater clinical efficacy. Thus, structural changes in catalase induced by clofazimine were monitored by various biophysical techniques including UV‐visible spectrometer, fluorescence spectroscopy, circular dichroism, and dynamic light scattering techniques. Increase in absorption spectra (UV‐visible spectrum) confers the complex formation between drug and protein. Fluorescence quenching with a binding constants of 2.47 × 104 M−1 revealed that clofazimine binds with protein. Using fluorescence resonance energy transfer, the distance (r ) between the protein (donor) and drug (acceptor) was found to be 2.89 nm. Negative Gibbs free energy change (ΔG °) revealed that binding process is spontaneous. In addition, an increase in α‐helicity was observed by far‐UV circular dichroism spectra by adding clofazimine to protein. Dynamic light scattering results indicate that topology of bovine liver catalase was slightly altered in the presence of clofazimine. Hydrophobic interactions are the main forces between clofazimine and catalase interaction as depicted by molecular docking studies. Apart from hydrophobic interactions, some hydrogen bonding was also observed during docking method. The results obtained from the present study may establish abundant in optimizing the properties of ligand‐protein mixtures relevant for numerous formulations.  相似文献   

5.
Ribonuclease A (RNase A) with several high affinity binding sites is a possible target for many organic and inorganic molecules. 3'-Azido-3'-deoxythymidine (AZT) is the first clinically effective drug for the treatment of human immunodeficiency virus (HIV) infection. The drug interactions with protein and nucleic acids are associated with its mechanism of action in vivo. This study was designed to examine the interaction of AZT with RNase A under physiological conditions. Reaction mixtures of constant protein concentration (2%) and different drug contents (0.0001-0.1 mM) are studied by UV-visible, FTIR, and circular dichroism spectroscopic methods in order to determine the drug binding mode, the drug binding constant, and the effects of drug complexation on the protein and AZT conformations in aqueous solution. The spectroscopic results showed one major binding for the AZT-RNase complexes with an overall binding constant of 5.29 x 10(5) M(-1). An increase in the protein alpha helicity was observed upon AZT interaction, whereas drug sugar pucker remained in the C2'-endo/anti conformation in the AZT-RNase complexes.  相似文献   

6.
We determined the binding sites of curcumin (cur), resveratrol (res), and genistein (gen) with milk β-lactoglobulin (β-LG) at physiological conditions. Fourier transform infrared spectroscopy, circular dichroism, and fluorescence spectroscopic methods as well as molecular modeling were used to determine the binding of polyphenol–protein complexes. Structural analysis showed that polyphenols bind β-LG via both hydrophilic and hydrophobic contacts with overall binding constants of Kcurcumin–β-LG?=?4.4 (±?.4)?×?104 M?1, Kresveratrol–β-LG?=?4.2 (±?.2)?×?104 M?1, and Kgenistein–β-LG?=?1.2 (±?.2)?×?104?M?1. The number of polyphenol molecules bound per protein (n) was 1 (cur), 1.1 (res), and 1 (gen). Molecular modeling showed the participation of several amino acid residues in polyphenol–protein complexation with the free binding energy of ?12.67 (curcumin–β-LG), ?12.60 (resveratrol–β-LG), and ?10.68?kcal/mol (genistein–β-LG). The order of binding was cur?>?res?>?gen. Alteration of the protein conformation was observed in the presence of polyphenol with a major reduction of β-sheet and an increase in turn structure, causing a partial protein structural destabilization. β-LG might act as a carrier to transport polyphenol in vitro.  相似文献   

7.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

8.
The main aim of the present study was to evaluate potential of ternary complexation (comprising of drug, cyclodextrin and polymer) as an approach for taste masking. For this purpose famotidine with property of bitter taste was selected as a model drug. Improvement in taste masking capability of cyclodextrin towards famotidine was evaluated by formulating a ternary complex including hydrophilic polymer hydroxyl propyl methyl cellulose (HPMC 5 cps) as the third component. Phase solubility analysis at 25 °C was carried out for both the binary systems (viz. drug–cyclodextrin and drug–polymer) and the ternary system (drug–cyclodextrin–polymer). Ternary complex was prepared using solution method and was further characterized using XRD, DSC, FT-IR and microscopic studies. In vitro dissolution study was carried out to see the effect of ternary complexation on drug release. Taste perception study was carried out on human volunteers to evaluate the taste masking ability of ternary complexation. Results obtained from phase solubility analysis showed that the combined use of polymer and cyclodextrin effectively increased the stability constant of the complex [from 538 M−1 for binary system to 15,096 M−1 for ternary system]. Ternary system showed effective taste masking as compared to binary complex and at the same time showed no limiting effect on the drug release (D.E15min = 90%). The effective taste masking was attributed to the enhanced complexation of famotidine in ternary system compared to binary system and the same was confirmed from the characterization studies. In conclusion, the study confirmed that ternary complexation can be utilized as an alternative approach for effective taste masking.  相似文献   

9.
Breviscapine, a cerebrovascular drugs extracted from the Chinese herb Erigeron breviscapinus, has been frequently used to clinically treat cerebrovascular diseases such as cerebral thrombosis, cerebral infarction, and cerebral circulation insufficiency. In order to understand its pharmacology or toxicity, the binding mechanism of breviscapine to a model protein, human serum albumin (HSA), was probed by fluorescence, circular dichroism, Fourier transform infrared spectroscopy (FTIR), and electrochemical impedance spectroscopy approaches. The binding affinities and number of the drug with HSA were about 1.73 × 104 M?1 and 0.99 at 293 K, respectively. The conformation of the protein was slightly altered after interacting with breviscapine. The drug–protein complex was mainly stabilized by electrostatic forces.  相似文献   

10.
Chlorogenic acid, 3’-O-caffeoyl D-quinic acid, is an inherent ligand present inHelianthus annuus L. The effect of pH on chlorogenic acid binding to helianthinin suggests that maximum binding occurs at pH 6.0. The protein-polyphenol complex precipitates as a function of time. The association constant of the binding of chlorogenic acid to helianthinin, determined by equilibrium dialysis, at 31°C has a value of 3.5 ± 0.1 × 104M−-1 resulting in a ΔG value of − 6.32 ± 0.12 kcal /mol. The association constantK ais 1.0 ± 0.1 × 104M−1 as determined by ultraviolet difference spectral titration at 25°C with ΔG° of -5.46 ± 0.06 kcal/mol. From fluorescence spectral titration at 28°C, theK avalue is 1.38 ± 0.1 × 1 0 4M−1 resulting in a ΔG of − 5.70 ± 0.05 kcal/mol. The total number of binding sites on the protein are 420 ± 50 as calculated from equilibrium dialysis. Microcalorimetric data of the ligand-protein interaction at 23°C suggests mainly two classes of binding. The thermal denaturation temperature,T mof the protein decreases from 76°C to 72°C at 1 × 10−3M chlorogenic acid concentration upon complexation. This suggests that the complexation destabilizes the protein. The effect of temperature onK aof chlorogenic acid shows a nonlinear increase from 10.2°C to 45°C. Chemical modification of both lysyl and tryptophanyl residues of the protein decreases the strength of binding of chlorogenic acid. Lysine, tryptophan and tyrosine of protein are shown to be present at the binding site. Based on the above data, it is suggested that charge-transfer complexation and entropically driven hydrophobic interaction are the predominant forces that are responsible for binding of chlorogenic acid to the multisubunit protein, helianthinin. Publication No. 324.  相似文献   

11.
Multiple spectroscopic techniques, such as fluorescence, absorption, and circular dichroism along with in silico studies were used to characterize the binding of a potent inhibitor molecule, CCG1423 to the major transport protein, human serum albumin (HSA). Fluorescence and absorption spectroscopic results confirmed CCG1423–HSA complex formation. A strong binding affinity stabilized the CCG1423–HSA complex, as evident from the values of the binding constant (Ka = 1.35 × 106–5.43 × 105 M?1). The KSV values for CCG1423–HSA system were inversely correlated with temperature, suggesting the involvement of static quenching mechanism. Thermodynamic data anticipated that CCG1423–HSA complexation was mainly driven by hydrophobic and van der Waals forces as well as hydrogen bonds. In silico analysis also supported these results. Three-dimensional fluorescence and circular dichroism spectral analysis suggested microenvironmental perturbations around protein fluorophores and structural (secondary and tertiary) changes in the protein upon CCG1423 binding. CCG1423 binding to HSA also showed some protection against thermal denaturation. Site-specific marker-induced displacement results revealed CCG1423 binding to Sudlow’s site I of HSA, which was also confirmed by the computational results. A few common ions were also found to interfere with the CCG1423–HSA interaction.  相似文献   

12.
The interaction of deferiprone (1,2-dimethyl-3-hydroxy-pyrid-4-one) L1, the first clinically available oral iron chelator, with the tetrameric allosteric protein hemoglobin from human red blood cells has been investigated spectrofluorometrically and by circular dichroism spectroscopy. The interaction is hydrogenbond like electrostatic in nature, the binding constant being 4.54 × 103 M-1 in 0.15 M NaCl. Circular dichroism studies indicate a conformational change of hemoglobin in presence of deferiprone, helicity of hemoglobin being reduced in presence of increasing concentration of the drug L1.  相似文献   

13.
The interactions of three platinum(II)-based anticancer complexes [(5,6-dimethyl-1,10-phenanthroline)(1S,2S-diaminocyclohexane)platinum(II)]2+, [(5,6-dimethyl-1,10-phenanthroline)(1R,2R-diaminocyclohexane)platinum(II)]2+, and [(5,6-dimethyl-1,10-phenanthroline)(1,2-diaminoethane)platinum(II)]2+ (56MEEN) with BSA have been examined by circular dichroism (CD), fluorescence and 1H pulsed gradient spin–echo (PGSE) diffusion NMR spectroscopy. The number of association constants and sites differed depending upon the spectroscopic method. This may be because each technique monitors different types of interaction/s and/or as a consequence of the different concentration ranges required for each technique. The titration of BSA with the achiral 56MEEN as monitored by CD indicates a reduction in the α-helical nature of the albumin, with the association constant calculated to be ~5 × 106 M−1 for one site. Due to the chiral nature of the other two complexes, their association with albumin was not monitored using CD but was examined using fluorescence and PGSE diffusion NMR. Titration of BSA with any of the three metal complexes resulted in quenching of fluorescence, with the number of association sites calculated to be ~1.1, with an association constant of ~2 × 105 M−1. PGSE diffusion NMR provided insights into interactions occurring with the BSA in its entirety, rather than with individual regions. Metal complex binding sites were estimated (~10 equivalent) from the diffusion data, with the average association constant for all sites ~102–103M−1. These experiments highlight the information that can be elucidated from complementary spectroscopic techniques and demonstrate the usefulness of PGSE diffusion NMR in monitoring multiple weak binding sites, which is of great importance in studying drug-biomolecule interactions.  相似文献   

14.
Summary Callus of the mangrove plant, Sonneratia alba J. Smith, established from pistils of flower buds were cultured on solid Murashige and Skoog medium supplemented with 0 to 500 mM NaCl. Maximum growth was observed with 50 mM NaCl, and net growth of callus occurred for concentrations up to 200 mM NaCl. At 500 mM NaCl, growth of callus was completely inhibited, although a part of the tissue was still alive after 30 d. Cellular levels of Na+ and Cl were greatly increased by the treatment with NaCl. Uptake of K+ was also enhanced and was accompanied by increasing levels of Na+ and Cl so that the Na+/K+ ratio was almost constant (4.1–4.2) in callus grown with 50–200 mM NaCl. Levels of Mg2+ and Ca2+ were not changed significantly with 50–200 mM NaCl, whereas levels of free NH 4 + , NO 3 and SO 4 2− ions, which are convertible to organic compounds, were lowest in callus grown with 50 mM NaCl. The rate of conversion of 15NH 4 + into macromolecules during 30 d culture with 0–100 mM NaCl did not vary greatly, but 200 mM NaCl reduced the biosynthesis of macromolecules from this ion. The highest rate of conversion of 15NO 3 into macromolecules was observed at 50 mM NaCl. Identification of compatible solutes with NMR-spectroscopy indicated that mannitol is the compatible solute for intact plants of Sonneratia alba, but no accumulation of mannitol was found in calluses, not even in those grown at high concentrations of NaCl.  相似文献   

15.
Bovine pancreatic ribonuclease A (RNase A) catalyzes the cleavage of P-O5' bonds in RNA on the 3' side of pyrimidine to form cyclic 2',5'-phosphates. Even though extensive structural information is available on RNase A complexes with mononucleotides and oligonucleotides, the interaction of RNase A with tRNA has not been fully investigated. We report the complexation of tRNA with RNase A in aqueous solution under physiological conditions, using a constant RNA concentration and various amounts of RNase A. Fourier transform infrared, UV-visible, and circular dichroism spectroscopic methods were used to determine the RNase binding mode, binding constant, sequence preference, and biopolymer secondary structural changes in the RNase-tRNA complexes. Spectroscopic results showed 2 major binding sites for RNase A on tRNA, with an overall binding constant of K = 4.0 x 105 (mol/L)-1. The 2 binding sites were located at the G-C base pairs and the backbone PO2 group. Protein-RNA interaction alters RNase secondary structure, with a major reduction in alpha helix and beta sheets and an increase in the turn and random coil structures, while tRNA remains in the A conformation upon protein interaction. No tRNA digestion was observed upon RNase A complexation.  相似文献   

16.
The double-helical complex formed from m6m9A and poly U has been characterized by circular dichroism and u.v. spectrophotometry. The circular dichroism of the complex is similar to that of the double-helical poly A poly U complex both in shape and in magnitude and thus indicates a quite similar structure. The double Helix–coil transition has been studied at various nucleotide concentrations and at three different ionic strengths. As expected for the binding of a base to a polymer, the Helix–coil transition is shifted to higher temperatures by increasing nucleotide concentrations, but is not affected by changes of the ionic strength. The melting curves are analyzed according to a linear Ising model taking the stacking of the monomers into account. At 0°C the equilibrium constant for nucleation is found to be 2–5 M?1 and that for chain growth is 500 M?1. The enthalpy change associated with chain growth is ?11.2 ± 1 kcal M?1.  相似文献   

17.
Copper (II) complex of formulation [Cu–Phen–Tyr](H2O)](ClO4) (Phen = 1,10-phenanthroline, l-Tyr = l-tyrosine), has been prepared, and their induced DNA oxidative cleavage activity studied. The complex binds to DNA by intercalation, as deduced from the absorption and fluorescence spectral data. Scatchard plots constructed from the absorption titration data gave binding constant 2.44 × 104 M−1 of base pairs. Extensive hypochromism, broadening, and red shifts in the absorption spectra were observed. Upon binding to DNA, the fluorescence from the DNA–ethidium bromide system was efficiently quenched by the copper (II) complex. Stern–Volmer quenching constant 0.61 × 103 M−1 obtained from the linear quenching plots. [Cu–Phen–Tyr] complex efficiently cleave the supercoiled DNA to its nicked circular form with gallic acid as biological reductant at appropriate complex concentration. The gallic acid as reductant could observably improve copper (II) complex to DNA damage. The pseudo-Michaelis–Menten kinetic parameters (k cat, K M) were calculated to be 1.32 h−1 and 5.46 × 10−5 M for [Cu–Phen–Tyr] complex. Mechanistic studies reveal the involvement of superoxide anions and hydroxyl radical (HO·) as the reactive species under an aerobic medium.  相似文献   

18.
The drug–protein interaction has been the subject of increasing interest over the decades. In the present communication, the interaction of liver cystatin with anti-cancer (adriamycin) and anti-hepatitis (adevofir dipivoxil) drugs was studied by thiol-protease inhibitory assay, UV absorption, fluorescence spectroscopy and circular dichroism (CD). A static type of quenching was observed between the protein and the drug molecules. Binding constant (Ka) of adriamycin to liver cystatin (LC) was found to be 1.08 × 106 M−1. Moreover, binding site number was found to be 2. Importantly, cystatin loses its activity in the presence of adriamycin. However, intrinsic fluorescence studies in the presence of adevofir dipivoxil showed enhancement in the fluorescence intensity suggesting that binding of adevofir to LC caused unfolding of the protein. The unfolding of the test protein was also accompanied by significant loss of inhibitory activity. CD spectroscopy result showed, both adriamycin and adevofir dipivoxil caused perturbation in the secondary structure of liver cystatin. The possible implications of these results will help in combating drug induced off target effects.Abbreviations: LC, liver cystatin; ADR, adriamycin; CD, circular dichroism; Ka, binding constant; HBV, human hepatitis B virus  相似文献   

19.
The optical activity induced on binding the drug netrospin (NET) in the minor groove of DNA is studied in five oligonucleotides (OGNs) as a function of (1) the size of the binding site in (5′-(GC)2AATT(GC)2-3′)2 (OGN 1a) versus (5′-(GC)2AAATTT(GC)2-3′)2 (OGN 1b) and (2) the distance between two AATT binding sites in (5′-(GC)2AATT(GC) x AATT(GC)2-3′)2, with x = 1, 2, or 3 (OGNs 2a, b, c, respectively). NET binding is monitored via the induced circular dichroism (CD) at ~315 nm, where the nucleic acids are optically inactive. The CD titrations, fit to a tight binding model, yield lower limits for the binding constant, Ka, ≥8 × 107 M−1 for OGN 1a and ≥2 × 108 M−1 for OGNs 2a, b, c in 1 mM buffer. In 100 mM buffer, tight binding occurs in all five OGNs with Ka ≥ 8 × 107 M−1 for OGN 1a and ≥1 × 108 M−1 for OGNs 1b and 2a, b, c. In contrast, the elongated AAATTT binding site of OGN 1b results in weak binding of NET in 1 mM buffer, where competing electrostatic interactions with the solvent environment are lower. In the constructs with two binding sites, the increase in flexibility introduced by intervening GC base pairs does not induce co-operative binding, although differences in the number of binding sites, n (2.05–2.65), indicate that there may be differences in the way NET is bound in OGNs 2a, b, c. In addition, the large shifts in the absorption spectra induced in bound versus free NET, and effects on the CD spectral bands at higher energy, are discussed in terms of electrostatic and excitonic interactions.  相似文献   

20.
Rate and equilibrium measurements of ryanodine binding to terminal cysternae fractions of heavy sarcoplasmic reticulum vesicles demonstrate that its activation by high concentrations of monovalent salts is based on neither elevated osmolarity nor ionic strength. The effect of the ions specifically depends on their chemical nature following the Hofmeister ion series for cations (Li+ < NH+ 4 < K∼ Cs+≤ Na+) and anions (gluconate < Cl < NO3 ∼ ClO4 ∼ SCN) respectively, indicating that both are involved in the formation of the salt-protein complex that can react with ryanodine. Activation by rising salt concentrations exhibits saturation kinetics with different dissociation constants (25–11 m) and different degrees of cooperativity (n= 1.5–4.0) for the respective salts. Maximal second order binding rates between 40,000 and 80,000 (m −1· sec−1) were obtained for chlorides and nitrates of 1a group alkali ions with the exception of lithium supporting only rates of maximally 10,000 (M−1· sec−1). The nitrogen bases, NH+ 4 and Tris+, in combination with chloride or nitrate, behave divergently. High maximal binding rates were achieved only with NH4NO3. The dissociation constants for the ryanodine–protein complexes obtained by measurements at equilibrium proved to depend differently on salt concentration, yet, converging to 1–3 nm for the applied salts at saturating concentrations. The salts do not affect dissociation of the ryanodine protein complex proving that the effect of salts on the protein's affinity for ryanodine is determined by their effect on the on-rate of ryanodine binding. ATP and its analogues modify salt action resulting in elevated maximal binding rates and reduction or abolition of binding cooperativity. Linear relations have been obtained by comparing the rates of ryanodine binding at different salt concentrations with the rates or the initial amplitudes (15 sec) of salt induced calcium release from actively loaded heavy vesicles indicating that the various salts promote specifically and concentration dependently channel opening and its reaction with ryanodine. Received: 9 February 1998/Revised: 24 April 1998  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