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.
Plasma protein binding of oxybutynin (OXY) was investigated quantitatively and enantioselectively using high-performance frontal analysis (HPFA). An on-line HPLC system which consists of HPFA column, extraction column and analytical column was developed to determine the unbound concentrations of OXY enantiomers in human plasma, in human serum albumin (HSA) solutions, and in human alpha1-acid glycoprotein (AGP) solutions. OXY is bound in human plasma strongly and enantioselectively. The bound drug fraction in human plasma containing 2-10 microM (R)- or (S)-OXY was higher than 99%, and the unbound fraction of (R)-OXY was 1.56 times higher than that of (S)-isomer. AGP plays the dominant role in this strong and enantioselective plasma protein binding. The total binding affinities (nK) of (R)- and (S)-OXY to AGP were 6.86 x 10(6) and 1.53 x 10(7) M(-1), respectively, while the nK values of (R)- and (S)-OXY to HSA were 2.64 x 10(4) and 2.19 x 10(-4) M(-1), respectively. The binding affinity of OXY to AGP is much higher than that to HSA, and shows high enantioselectivity (SIR ratio of nK values is 2.2). It was found that both enantiomers are bound competitively at the same binding site on an AGP molecule. The binding property between OXY and low density lipoprotein (LDL) was investigated by using the frontal analysis method incorporated in high-performance capillary electrophoresis (HPCE/FA). It was found the binding is non-saturable and non-enantioselective.  相似文献   

2.
Plasma protein binding of N-desethyloxybytynin (DEOXY), a major active metabolite of oxybutynin (OXY), was investigated quantitatively and enantioselectively using high-performance frontal analysis (HPFA). An on-line HPLC system which consists of HPFA column, extraction column and analytical column was developed to determine the unbound concentrations of DEOXY enantiomers in human plasma, in human serum albumin (HSA) solutions, and in human alpha1-acid glycoprotein (AGP) solutions. DEOXY is bound in human plasma strongly and enantioselectively. The unbound drug fraction in human plasma samples containing 5 microM (R)- or (S)-DEOXY was 1.19 +/- 0.001 and 2.33 +/- 0.044%, respectively. AGP plays the dominant role in this strong and enantioselective plasma protein binding of DEOXY. The total binding affinity (nK) of (R)-DEOXY and (S)-DEOXY to AGP was 2.97 x 10(7) and 1.31 x 10(7) M(-1), respectively, while the nK values of (R)-DEOXY and (S)-DEOXY to HSA were 7.77 x 10(3) and 8.44 x 10(3) M(-1), respectively. While the nK value of (S)-DEOXY is weaker than that of (S)-OXY (1.53 x 10(7) M(-1)), the nK value of (R)-DEOXY is 4.33 times stronger than that of (R)-OXY (6.86 x I0(6) M(-1)). This suggests that the elimination of an ethyl group weakens the binding affinity of the (S)-isomer because of the decrease in hydrophobicity, while the binding affinity of the (R)-isomer is enhanced by the decrease in steric hindrance. The total binding affinity of DEOXY to HSA is much lower than that of DEOXY-AGP binding as well as OXY-HSA binding (2.64 x 10(4) and 2.19 x 10(4) M(-1) for (R)-OXY and (S)-OXY, respectively). The study on competitive binding between OXY and DEOXY indicated that DEOXY enantiomers and OXY enantiomers are all bound competitively at the same binding site of AGP molecule.  相似文献   

3.
Hong Y  Tang Y  Zeng S 《Chirality》2009,21(7):692-698
The interaction of propafenone (PPF) enantiomers with human plasma, human serum albumin (HSA), alpha(1)-acid glycoprotein (AGP), as well as with plasma from rat, rabbit, and cow was investigated using indirect chiral high performance liquid chromatography (HPLC) and ultrafiltration techniques. The stronger binding of the S-PPF found in human plasma was due to AGP. Two classes of binding sites in AGP were identified: one with high-affinity and small binding capacity (K(1(S)) = 7.65 x 10(6) M(-1), n(1(S)) = 0.50; K(1(R)) = 2.81 x 10(6) M(-1), n(1(R)) = 0.46), which revealed stereoselectivity; the other with low-affinity and high-binding capacity (n(2(S)) K(2(S)) = 9.95 x 10(3) M(-1); n(2(R)) K(2(R)) = 9.74 x 10(3) M(-1)). The binding to HSA was found to be weak and not enantioselective (nK(S) = 2.08 x 10(3) M(-1), nK(R) = 2.05 x 10(3) M(-1)). The interaction between enantiomers observed in human plasma was confirmed as a competitive type interacting at the high-affinity site in AGP. The binding mode of both enantiomers with AGP was mainly hydrophobic bond. PPF enantiomers had higher-binding affinity for the F-S variant of human AGP. Drug-drug binding interaction studies showed that verapamil, diazepam, nifedipine, furosemide, nitrendipine, and nimodipine did not affect the binding of PPF enantiomers except quinidine and aprindine at the therapeutic concentration. Comparative studies indicated considerable species-dependent binding stereoselectivity between plasma of the four species investigated.  相似文献   

4.
The interaction of propafenone enantiomers with human alpha 1-acid glycoprotein was studied using high-performance liquid chromatography. Each of the two optical antipodes interacted with one class of high-affinity binding sites characterized by Ka(R) = (6.18 +/- 0.93) x 10(5) M-1, n(R) = 1.34 +/- 0.09 for the (R)-isomer and Ka(S) = (8.93 +/- 1.82) x 10(5) M-1, n(S) = 0.99 +/- 0.08 for the (S)-isomer. Nonspecific binding to secondary low-affinity high-capacity binding site(s) was only slightly greater in the case of the (S)-enantiomer (n'k'(S) = (1.06 +/- 0.09) x 10(4) M-1) compared to the (R)-enantiomer (n'k'(R) = (6.87 +/- 0.72) x 10(3) M-1). It was concluded that both enantiomers interact with common single class of high-affinity binding sites on AAG (along with nonspecific binding) exhibiting only slight stereoselectivity for propafenone.  相似文献   

5.
Visy J  Fitos I  Mády G  Urge L  Krajcsi P  Simonyi M 《Chirality》2002,14(8):638-642
The binding of bimoclomol enantiomers to human plasma, its components, as well as to plasma from monkey, dog, rat, and mouse was investigated by ultrafiltration and equilibrium dialysis. The considerably stronger binding of the (-)-(S)-enantiomer found in human plasma is due to the alpha(1)-acid glycoprotein (AAG) component. The binding parameters for AAG (n(R)K(R) = 1.3 x 10(4) M(-1) and n(S)K(S) = 1.0 x 10(5) M(-1)) revealed high enantioselectivity, while the binding to human serum albumin was found to be weak (nK = 5 x 10(3) M(-1)) and not stereoselective. (-)-(S)-Bimoclomol was extensively displaced in the presence of specific marker ligands for the "FIS" subfraction of human AAG. Comparative binding studies indicated considerable differences between plasma of the five species investigated.  相似文献   

6.
The interaction of pirprofen enantiomers with human serum albumin (HSA) was investigated by means of high-performance liquid chromatography (HPLC), circular dichroism (CD), and 1H NMR spectroscopy. HPLC experiments indicated that both pirprofen enantiomers were bound to one class of high-affinity binding sites (n(+) = 1.91 +/- 0.13, K(+) = (4.09 +/- 0.64) x 10(5) M-1, n(-) = 2.07 +/- 0.13, K(-) = (6.56 +/- 1.35) x 10(5) M-1) together with nonspecific binding (n'K'(+) = (1.51 +/- 0.21) x 10(4) M-1, n'K'(-) = (0.88 +/- 0.13) x 10(-4) M-1). Slight stereoselectivity in specific binding was demonstrated by the difference in product n(+)K(+) = (0.77 +/- 0.08) x 10(6) M-1 vs. n(-)K(-) = (1.30 +/- 0.21) x 10(6) M-1, i.e., the ratio n(-)K(-)/n(+)K(+) = 1.7. CD measurements showed changes in the binding sites located on the aromatic amino acid side chains (a small positive band at 315 nm and a pronounced negative extrinsic Cotton effect in the region 250-280 nm). The protein remains, however, in its predominantly alpha-helical conformation. The 1H NMR difference spectra confirmed that both pirprofen enantiomers interacted with HSA specifically, most probably with site II on the albumin molecule.  相似文献   

7.
The binding of the (R)‐ and (S)‐enantiomers of amlodipine to bovine serum albumin (BSA), human serum albumin (HSA), α1‐acid glycoprotein (AGP), and human plasma (HP) was studied by equilibrium dialysis over the concentration range of 75–200 μM at a protein concentration of 150 μM. Unbound drug concentrations were determined by enantioselective capillary electrophoresis using 50 mM phosphate buffer, pH 2.5, containing 18 mM α‐cyclodextrin as background electrolyte. Saturation of the protein binding sites was not observed over the concentration range tested. Upon application of racemic amlodipine besylate, (S)‐amlodipine was bound to a higher extend by HSA and HP compared with (R)‐amlodipine, whereas the opposite binding of the enantiomers was observed for BSA and AGP. Scatchard analysis was used to illustrate the different binding affinities of amlodipine besylate enantiomers to BSA, HSA and AGP. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
The intravenous (0.5 mg/kg) and oral (5 mg/kg) dose kinetics of verapamil were studied in 6 dogs during steady-state oral verapamil dosing (5 mg/kg every 8 h for 3 days). Racemic verapamil and norverapamil, a metabolite of verapamil, were quantitated in plasma by HPLC-fluorescence detection. The verapamil peaks eluting off the column were collected and rechromatographed on an Ultron-OVM column, which resolved the two verapamil enantiomers. After intravenous administration, the systemic clearance and apparent volume of distribution of (?)-(S)-verapamil were nearly twice that of the (+)-(R)-isomer. There was no difference in the elimination half-lives between the two isomers. After oral administration, the oral clearance of (?)-(S)-verapamil was 20 times that of the (+)-(R)-isomer. The apparent bioavailability of (+)-(R)-verapamil was over 14 times that of (?)-(S)-verapamil. The plasma protein binding of the (+)-(R)-isomer was slightly higher by 5% than (?)-(S)-verapamil; however, this effect was not enough to account for the difference between the apparent volume of distribution of the enantiomers, indicating that the tissue binding of (?)-(S)-verapamil was greater than that of the (+)-(R)-isomer. This data on the disposition of the enantiomers of verapamil in the dog is similar to that reported for man and demonstrates that the dog may be an appropriate animal model for man in future studies on the disposition of the enantiomers of verapamil. © 1993 Wiley-Liss, Inc.  相似文献   

9.
An assay comprising two simple, selective and isocratic HPLC methods with UV detection was developed and validated for measuring warfarin enantiomers and all five warfarin monohydroxylated metabolites in patient blood plasma. Following liquid/liquid extraction from 1 ml of blood plasma a baseline separation of analytes was achieved on chiral (alpha(1) acid glycoprotein - AGP) and achiral (C(18)) column. Both methods were consistent (R.S.D.<6.9% for warfarin enantiomers and<8.9% for monohydroxylated metabolites) and linear (r>0.998). The limits of detection were 25 ng/ml for warfarin enantiomers, 25 ng/ml for 4'-, 10-, 6- and 7-hydroxywarfarin, 35 ng/ml for 8-hydroxywarfarin and 50 ng/ml for racemic warfarin. In a clinical study in 204 patients, it was confirmed that the assay is appropriate for evaluation of influences of genetic polymorphisms, demographic factors and concomitant drug treatment on warfarin metabolism.  相似文献   

10.
A sensitive and reproducible HPLC method utilizing a commercially available chiral α1-acid glycoprotein (AGP) phase has been developed to separate and quantify the enantiomers of nicotine. The method is suitable for routine use as indicated by column life. The quantification of (R/S:0.05/99.95)-nicotine or (R/S:99/1)-nicotine was possible. In addition, the separation or at least partial separation of the enantiomers of nornicotine and nornicotine-derived compounds was achieved. © 1993 Wiley-Liss, Inc.  相似文献   

11.
Główka FK  Caldwell J 《Chirality》2002,14(9):736-741
The binding of the enantiomers of indobufen (INDB) to human serum proteins was investigated using the racemic mixture or the pure (+)-S-enantiomer in a concentration range of 2.5-100.0 mg/L. In addition, the pharmacokinetics of free (unbound) and total INDB enantiomers were studied 1) following administration of a single 200 mg rac-INDB tablet to healthy volunteers, and 2) in obliterative atherosclerosis patients at steady state. The free fraction of INDB was obtained by ultrafiltration. Using the racemic mixture, the binding parameters of the two enantiomers were different, showing enantioselectivity in protein binding. The (-)-R-enantiomer was bound more strongly to human serum albumin, with association constant K = 11.95 +/- 0.98 x 10(5) M(-1) and n = 0.72 +/- 0.02 binding sites. The comparable data for the (+)-S-enantiomer were K = 4.65 +/- 0.02 x 10(5) M(-1), n = 0.92 +/- 0.01. When the binding of (+)-S-enantiomer was studied alone, the association constant K (2.10 +/- 0.18 x 10(5) M(-1)) was lower and the number of binding sites was increased, to n = 1.87 +/- 0.17. Competition occurred between the enantiomers, with the (-)-R-enantiomer displacing its antipode. The fraction of both enantiomers bound to serum proteins was 99.0%, which increased with decreasing initial concentration of the enantiomers. In healthy volunteers the (+)-S-enantiomer was eliminated faster than its (-)-R antipode, resulting in a lower AUC for the (+)-S-enantiomer. Significant differences were observed in the total INDB enantiomer concentrations. The mean unbound fraction of (-)-R- and (+)-S-INDB was 0.45% and 0.43%, respectively. Levels of the free (+)-S-enantiomer were higher than its (-)-R-antipode at steady state in patients with obliterative atherosclerosis who also took other drugs. The free enantiomer fraction increased to around 1% upon repeated administration. We conclude that the more rapid elimination of the (+)-S enantiomer is associated with its weaker binding to serum proteins.  相似文献   

12.
Coumarin-type anticoagulants, warfarin, phenprocoumon and acenocoumarol, were tested for their stereoselective binding to the human orosomucoid (ORM; AGP) genetic variants ORM 1 and ORM 2. Direct binding studies with racemic ligands were carried out by the ultrafiltration method; the concentrations of free enantiomers were determined by capillary electrophoresis. The binding of pure enantiomers was investigated with quinaldine red fluorescence displacement measurements. Our results demonstrated that all investigated compounds bind stronger to ORM 1 variant than to ORM 2. ORM 1 and human native AGP preferred the binding of (S)-enantiomers of warfarin and acenocoumarol, while no enantioselectivity was observed in phenprocoumon binding. Acenocoumarol possessed the highest enantioselectivity in AGP binding due to the weak binding of its (R)-enantiomer. Furthermore, a new homology model of AGP was built and the models of ORM 1 and ORM 2 suggested that difference in binding to AGP genetic variants is caused by steric factors.  相似文献   

13.
Imatinib is a selective tyrosine kinase inhibitor, successfully used for the treatment of chronic myelogenous leukaemia. Its strong plasma protein binding referred to alpha1-acid glycoprotein (AGP) component was found to inhibit the pharmacological activity. AGP shows genetic polymorphism and the two main genetic variants have different drug binding properties. The binding characteristics of imatinib to AGP genetic variants and the possibility of its binding interactions were investigated by various methods. The results proved that binding of imatinib to the two main genetic variants is very different, the high affinity binding belongs dominantly to the F1-S variant. This interaction is accompanied with specific spectral changes (induced circular dichroism, UV change, intrinsic fluorescence quenching), suggesting that the bound ligand has chiral conformation that would largely overlap with other ligands inside the protein cavity. Binding parameters of Ka=1.7(+/-0.2)x10(6)M(-1) and n=0.94 could be determined for the binding on the F1-S variant at 37 degrees . Imatinib binding on the A variant is weaker and less specific. The binding affinity of imatinib to human serum albumin (nKa approximately 3 x 10(4)M(-1)) is low. Pharmacologically relevant binding interactions with other drugs can be expected on the F1-S variant of AGP.  相似文献   

14.
A new method for preparing immobilized alpha1-acid glycoprotein (AGP) for use in drug-protein binding studies was developed and optimized. In this approach, periodate was used under mild conditions to oxidize the carbohydrate chains in AGP for attachment to a hydrazide-activated support. The final conditions chosen for this oxidation involved the reaction of 5.0 mg/mL AGP at 4 degrees C and pH 7.0 with 5-20 mM periodic acid for 10 min. These conditions helped maximize the immobilization of AGP without significantly affecting its activity. This method was evaluated by using it to attach AGP to silica for use in high-performance affinity chromatography and self-competition zonal elution studies. In work with R- and S-propranolol, only one type of binding site was observed for both enantiomers on the immobilized AGP, in agreement with previous studies using soluble AGP. The association equilibrium constants measured for the immobilized AGP with R- and S-propranolol at pH 7.4 and 37 degrees C were 2.7 x 10(6) and 4.2 x 10(6) M(-1), respectively, with linear van't Hoff plots being obtained between 5 and 37 degrees C. Work performed with other drugs also gave good agreement between the behavior seen for immobilized AGP and that for soluble AGP. The same immobilization method described in this work could be used to attach AGP to other materials, such as those used for surface plasmon resonance or alternative biosensors.  相似文献   

15.
Boulton DW  Devane CL 《Chirality》2000,12(9):681-687
Methadone enantiomers and EDDP, the main metabolite of methadone, were separated (R(s) = 2.0 for methadone enantiomers) following liquid-liquid extraction from human serum and urine followed by reverse-phase high-performance liquid chromatography on a derivatized beta-cyclodextrin column and quantified at therapeutic concentrations with ultraviolet detection. Detector response was linear (r(2) > 0.98) to 1,000 and 2,500 ng x mL(-1) for methadone enantiomers and EDDP, respectively. The limit of quantification from a 1-mL biological sample was 2.5 and 5 ng x mL(-1) for methadone enantiomers and EDDP, respectively. Interday variation was <13% and intraday variation was <8% for the analytes of interest. The assay was applied to plasma protein and erythrocyte binding studies and a 96-h pharmacokinetic study in two healthy female volunteers following oral dosing with rac-methadone. The binding of methadone to plasma proteins was enantioselective with the active (-)-(R) enantiomer having the highest free fraction (mean +/- SD: 21.2+/-7.6% vs. 13.3+/-6.2% for (+)-(S)-methadone, n = 8). Binding of methadone to erythrocytes was not apparently enantioselective (38.6+/-1.3% and 38.1+/-1.4% bound for (-)-(R)- and (+)-(S)-methadone, respectively). The pharmacokinetic study revealed enantioselective disposition of methadone in one volunteer but not in the other. EDDP was observed in urine but was only in small or undetectable concentrations in serum. The method is applicable to in vitro and pharmacokinetic studies of rac-methadone disposition in humans.  相似文献   

16.
Non-steroidal anti-inflammatory drugs (NSAIDs) are strongly bound to human serum albumin (HSA), mainly to sites I and II. The aim of this study was to characterize the binding site(s) of etodolac enantiomers under physiological conditions (580 μM HSA) using equilibrium dialysis. The protein binding of etodolac enantiomers, alone or in various ratios, was studied in order to evaluate the potential competition between them. Our results showed that (S)-etodolac was more strongly bound to HSA than (R)-etodolac. The displacement of one enantiomer by its antipode was observed only at high concentrations of the competitor, and was more pronounced for the (S)-form. Displacement studies of the enantiomers by specific probes of sites I and II of albumin, dansylamide, and dansylsarcosine, respectively, showed that (R)-etodolac was slightly displaced by both these probes whereas the free concentration of (S)-etodolac increased markedly in the presence of dansylsarcosine. Moreover, the binding of ligands to sites I and II is usually affected by alkaline pH, by chloride ions, and by fatty acids. For etodolac, the presence of 0.1 and 1 M chloride ions and increasing pH (5.5-9) decreased the binding of both enantiomers. The same result was obtained with addition of octanoic acid. Conversely, the addition of oleic, palmitic, or stearic acid to the protein solution increased the binding of (R)-etodolac, but decreased that of its antipode. All these findings suggest that (R)- and (S)-etodolac interact mainly with site II of HSA, and that the (R)-isomer is also bound to site I under physiological conditions. © 1996 Wiley-Liss, Inc.  相似文献   

17.
The role of the branching glycan structure of human alpha1-acid glycoprotein (AGP) in the interaction with basic drugs was investigated in terms of enantioselectivity in binding ability. AGP was separated by concanavalin A lectin affinity chromatography into two subfractions, the unretained AGP (UR-AGP) which has no biantennary glycan chain and the retained AGP (R-AGP) which possesses biantennary oligosaccharide chain(s). The unbound concentrations of propranolol (PRO) enantiomers and verapamil (VER) enantiomers in UR-AGP solution and R-AGP solution were determined by high-performance frontal analysis combined with capillary electrophoresis. It was found that (S)-PRO is bound to UR-AGP and R-AGP more strongly than (R)-PRO, whereas the reverse applies to VER enantiomers, and that such enantioselectivity is common to these proteins. This suggests that the branching type of glycan chains of AGP does not play significant role in the chiral recognition in binding these basic drugs.  相似文献   

18.
Stereoselective degradation of tebuconazole in rat liver microsomes   总被引:1,自引:0,他引:1  
Shen Z  Zhu W  Liu D  Xu X  Zhang P  Zhou Z 《Chirality》2012,24(1):67-71
The aim of this study was to assess the stereoselectivity of two tebuconazole [(RS)-1-p-chlorophenyl-4,4-dimethyl-3-(1H-1,2,4-triazol-1-ylmethyl)pentan-3-ol] enantiomers in in vitro system (rat liver microsomes). The analytes were extracted with acetic ether and concentrations were determined by high performance liquid chromatography (HPLC) with a cellulose tris(3,5-dimethylphenylcarbamate)-based chiral stationary phase. The degradation of rac-tebuconazole (15 μM) followed first-order kinetics, and the degradation of the S-tebuconazole (t(1/2) = 22.31 min) was faster than that of the R-tebuconazole (t(1/2) = 48.76 min), but no significant difference between the enantiomers was found in the respective incubation (7.5 μM for each). Kinetic assays showed that the K(m) was different between the two enantiomers (K(mR) = 14.83 ± 2.19, K(mS) = 12.23 ± 2.72). The interaction results revealed that there was competitive inhibition between S- and R-form, and there was a significant difference between the IC(50) of R- to S-tebuconazole and S- to R-tebuconazole (IC(50R/S)/IC(50S/R) = 4.98).  相似文献   

19.
o-Succinylbenzoate synthase (OSBS) from Amycolatopsis, a member of the enolase superfamily, catalyzes the Mn2+-dependent exergonic dehydration of 2-succinyl-6R-hydroxy-2,4-cyclohexadiene-1R-carboxylate (SHCHC) to 4-(2'-carboxylphenyl)-4-oxobutyrate (o-succinylbenzoate or OSB) in the menaquinone biosynthetic pathway. This enzyme first was identified as an N-acylamino acid racemase (NAAAR), with the optimal substrates being the enantiomers of N-acetyl methionine. This laboratory subsequently discovered that this protein is a much better catalyst of the OSBS reaction, with the value of k(cat)/K(M), for dehydration, 2.5 x 10(5) M(-1) s(-1), greatly exceeding that for 1,1-proton transfer using the enantiomers of N-acetylmethionine as substrate, 3.1 x 10(2) M(-1) s(-1) [Palmer, D. R., Garrett, J. B., Sharma, V., Meganathan, R., Babbitt, P. C., and Gerlt, J. A. (1999) Biochemistry 38, 4252-8]. The efficiency of the promiscuous NAAAR reaction is enhanced with alternate substrates whose structures mimic that of the SHCHC substrate for the OSBS reaction, for example, the value of k(cat)/K(M) for the enantiomers of N-succinyl phenylglycine, 2.0 x 10(5) M(-1) s(-1), is comparable to that for the OSBS reaction. The mechanisms of the NAAAR and OSBS reactions have been explored using mutants of Lys 163 and Lys 263 (K163A/R/S and K263A/R/S), the putative acid/base catalysts identified by sequence alignments with other OSBSs, including the structurally characterized OSBS from Escherichia coli. Although none of the mutants display detectable OSBS or NAAAR activities, K163R and K163S catalyze stereospecific exchange of the alpha-hydrogen of N-succinyl-(S)-phenylglycine with solvent hydrogen, and K263R and K263 catalyze the stereospecific exchange the alpha-hydrogen of N-succinyl-(R)-phenylglycine, consistent with formation of a Mn2+-stabilized enolate anion intermediate. The rates of the exchange reactions catalyzed by the wild-type enzyme exceed those for racemization. That this enzyme can catalyze two different reactions, each involving a stabilized enediolate anion intermediate, supports the hypothesis that evolution of function in the enolase superfamily proceeds by pathways involving functional promiscuity.  相似文献   

20.
The binding of Ru(phen)(2)dppz(2+) (dppz=dipyrido[3,2-a:2',3'-c]phenazine) to DNA was investigated at pH 7.0 and 25 degrees C using stopped-flow and spectrophotometric methods. Equilibrium measurements show that two modes of binding, whose characteristics depend on the polymer to dye ratio (C(P)/C(D)), are operative. The binding mode occurring for values of C(P)/C(D) higher than 3 exhibits positive cooperativity, which is confirmed by kinetic experiments. The reaction parameters are K=2 x 10(3)M(-1), omega=550, n=1, k(r)=(1.9+/-0.5) x 10(7)M(-1)s(-1) and k(d)=(9.5+/-2.5)x10(3)s(-1) at I=0.012 M. The results are discussed in terms of prevailing surface interaction with DNA grooves accompanied by partial intercalation of the dppz residue. The other binding mode becomes operative for C(P)/C(D)<3 and the equilibria analysis shows this is an ordinary intercalation mode (K=1.3 x 10(6) M(-1), n=1.5 at I=0.012 M and K=2 x 10(5) M(-1), n=1.2 at I=0.21 M). Similar behaviour is displayed by double-stranded poly(A).  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