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.
FliZ, a global regulatory protein under the control of the flagellar master regulator FlhDC, was shown to antagonize σ(S)-dependent gene expression in Escherichia coli. Thereby it plays a pivotal role in the decision between alternative life-styles, i.e. FlhDC-controlled flagellum-based motility or σ(S)-dependent curli fimbriae-mediated adhesion and biofilm formation. Here, we show that FliZ is an abundant DNA-binding protein that inhibits gene expression mediated by σ(S) by recognizing operator sequences that resemble the -10 region of σ(S)-dependent promoters. FliZ does so with a structural element that is similar to region 3.0 of σ(S). Within this element, R108 in FliZ corresponds to K173 in σ(S), which contacts a conserved cytosine at the -13 promoter position that is specific for σ(S)-dependent promoters. R108 as well as C(-13) are also crucial for DNA binding by FliZ. However, while a number of FliZ binding sites correspond to known σ(S)-dependent promoters, promoter activity is not a prerequisite for FliZ binding and repressor function. Thus, we demonstrate that FliZ also feedback-controls flagellar gene expression by binding to a site in the flhDC control region that shows similarity only to a -10 element of a σ(S)-dependent promoter, but does not function as a promoter.  相似文献   

3.
The effect of palmitoylation on the GTP-binding activity and conformation of Goalpha protein in hydrophobic and hydrophilic environments was studied. The binding assay was performed with an isotope labeled analog of GTP, GTP-gamma-35S, and its fluorescent analog, BODIPY FL-GTPgammaS was used to detect conformational change in the GTP-binding domain of Goalpha. Investigation of the GTP-gamma-35S binding activity of Goalpha shows that in a hydrophobic environment, mimicked by the presence of detergent, the apparent dissociation constant for palmitoylated Goalpha (K(D)=25.5x10(-9)+/-1.7x10(-9)M) increased threefold compared with that of non-palmitoylated Goalpha (K(D)=9.9x10(-9)+/-0.8x10(-9)M), while in an aqueous environment without detergent there is no significant difference between palmitoylated (K(D)=50.1 x 10(-9)+/-5.2x10(-9)M) and non-palmitoylated (K(D)=65.5x10(-9)+/-7.6x10(-9)M) Go(. This indicates that in a membrane environment palmitoylation may weaken the GTPgammaS binding ability of Go(. Fluorescent quenching studies using BODIPY FL-GTPgammaS as a probe showed that the conformation of the GTP-binding domain of Go( tends to become more compact after palmitoylation. These results imply that palmitoylation may regulate the GTP/GDP exchange of Goalpha by influencing the GTP-binding activity of Goalpha and facilitating the on-off switch function of the G protein in G protein-coupled signal transduction.  相似文献   

4.
Decameric vanadate (V(10)) inhibits the actin-stimulated myosin ATPase activity, noncompetitively with actin or with ATP upon interaction with a high-affinity binding site (K(i) = 0.27 +/- 0.05 microM) in myosin subfragment-1 (S1). The binding of V(10) to S1 can be monitored from titration with V(10) of the fluorescence of S1 labeled at Cys-707 and Cys-697 with N-iodo-acetyl-N'-(5-sulfo-1-naphthyl)ethylenediamine (IAEDANS) or 5-(iodoacetamido) fluorescein, which showed the presence of only one V(10) binding site per monomer with a dissociation constant of 0.16-0.7 microM, indicating that S1 labeling with these dyes produced only a small distortion of the V(10) binding site. The large quenching of AEDANS-labeled S1 fluorescence produced by V(10) indicated that the V(10) binding site is close to Cys-697 and 707. Fluorescence studies demonstrated the following: (i) the binding of V(10) to S1 is not competitive either with actin or with ADP.V(1) or ADP.AlF(4); (ii) the affinity of V(10) for the complex S1/ADP.V(1) and S1/ADP.AlF(4) is 2- and 3-fold lower than for S1; and (iii) it is competitive with the S1 "back door" ligand P(1)P(5)-diadenosine pentaphosphate. A local conformational change in S1 upon binding of V(10) is supported by (i) a decrease of the efficiency of fluorescence energy transfer between eosin-labeled F-actin and fluorescein-labeled S1, and (ii) slower reassociation between S1 and F-actin after ATP hydrolysis. The results are consistent with binding of V(10) to the Walker A motif of ABC ATPases, which in S1 corresponds to conserved regions of the P-loop which form part of the phosphate tube.  相似文献   

5.
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   

6.
The affinity of ortho-iodinated hexestrols for the estrogen binding protein from rat uterus, determined by competitive binding assay, decreases with progressive iodine substitution; 3-iodohexestrol (I-Hex) has a binding affinity 42% that of estradiol. Analysis of [3-H]-I-Hex binding in rat uterine cytosol by sucrose density gradient centrifugation shows both an estrogen-specific binding component (8 S) and a more abundant component (4 S) that is not estrogen specific. Scatchard analysis indicates that this latter binding is of high affinity (Kd equals to 3.7-8.3 times 10- minus-9 M) but is not uterine specific. Polyacrylamide gel electrophoresis shows that most of the [3-H]-I-Hex binding activity in serum and uterine cytosol is distinct from and anodic to the principal protein component (albumin), and that is comigrates with [14-C]thyroxine binding activity. In in vitro incubation of rat uteri, I-Hex can block the specific uptake of [3-H]estradiol into the nuclear fraction; it itself causes a translocation of estrogen-specific binding capacity (as measured by exchange) from cytoplasm to nuclei, and can induce the synthesis of an estrogen-specific uterine protein, all under conditions where it is not metabolically deiodinated to hexestrol. The uterotrophic activities of the iodohexestrols are in most cases comparable to that expected on the basis of their competitive binding affinities. However, selective, estrogen-specific uptake of [3-H]-I-Hex into rat uterus, either in vitro or in vivo, cannot be demonstrated.  相似文献   

7.
It is known that ternary complexes of myosin subfragment 1 (S1) with ADP and the Pi analogs beryllium fluoride (BeFx) and aluminum fluoride (AlF4-) are stable analogs of the myosin ATPase intermediates M* x ATP and M** x ADP x Pi, respectively. Using kinetic approaches, we compared the rate of formation of the complexes S1 x ADP x BeFx and S1 x ADP x AlF4- in the absence and in the presence of F-actin, as well as of the interaction of these complexes with F-actin. We show that in the absence of F-actin the formation of S1 x ADP x BeFx occurs much faster (3-4 min) than that of S1 x ADP x AlF4- (hours). The formation of these complexes in the presence of F-actin led to dissociation of S1 from F-actin, this process being monitored by a decrease in light scattering. The light scattering decrease of the acto-S1 complex occurred much faster after addition of BeFx (during 1 min) than after addition of AlF4- (more than 20 min). In both cases the light scattering of the acto-S1 complex decreased by 40-50%, but it remained much higher than that of F-actin measured in the absence of S1. The interaction of the S1 x ADP x BeFx and S1 x ADP x AlF4- complexes with F-actin was studied by the stopped-flow technique with high time resolution (no more than 0.6 sec after mixing of S1 with F-actin). We found that the binding of S1 x ADP x BeFx or S1 x ADP x AlF4- to F-actin is accompanied by a fast increase in light scattering, but it does not affect the fluorescence of a pyrene label specifically attached to F-actin. We conclude from these data that within this time range a "weak" binding of the S1 x ADP x BeFx and S1 x ADP x AlF4- complexes to F-actin occurs without the subsequent transition of the "weak" binding state to the "strong" binding state. Comparison of the light scattering kinetic curves shows that S1 x ADP x AlF4- binds to F-actin faster than S1 x ADP x BeFx does: the second-order rate constants for the "weak" binding to F-actin are (62.8 +/- 1.8) x 10(6) M-1 x sec-1 in the case of S1 x ADP x AlF4- and (22.6 +/- 0.4) x 10(6) M-1 x sec-1 in the case of S1 x ADP x BeFx. We conclude that the stable ternary complexes S1 x ADP x BeFx and S1 x ADP x AlF4- can be successfully used for kinetic studies of the "weak" binding of the myosin heads to F-actin.  相似文献   

8.
A monoclonal antibody detecting unusual Thy-1 determinants   总被引:3,自引:0,他引:3  
20-10-5S is a monoclonal antibody produced by the fusion of C3H anti-C3H.SW splenocytes with the SP2/0 cell line. The antibody appears to react with Thy-1 determinants by several criteria including cytotoxicity patterns, functional assays, genetic analyses, and competitive binding experiments. However, the antibody and the determinants it detects are unusual in that: 1) 20-10-5S is autoreactive; 2) the antibody shows allospecificity for Thy-1.2 vs Thy-1.1 antigens only on peripheral lymphocytes and not on thymocytes; and 3) the antibody reacts only with determinants on murine T cells and not with antigens on brain tissue or on rat thymocytes. It therefore seems that 20-10-5S reacts with murine T cell-specific Thy-1 determinants that are lost or modified during maturation of the cells on which they are expressed.  相似文献   

9.
10.
Gel filtration studies demonstrate that retinol receptors of chick retinal and pigment epithelial cytosols are (1) of very similar nature (2) of small molecular size (about 18000 daltons) and are different in character from serum proteins. Citral inhibits the binding of [3H]retinol to the retinal 2 S receptor. Retinol acetate competes with retinol for binding to 2 S receptor in both retina and pigment epithelium whereas retinol palmitate is an effective competitor only in the pigment epithelium. Dithiothreitol maximizes 2 S binding in retina and pigment epithelial cytosol; its absence does not lead to receptor aggregation however. A limited number of high affinity binding sites (2 S receptor) appear to be present in retina and pigment epithelium. A 5 S binding species is also present in pigment epithelium; it is similar in character to [3H]retinol binding in serum and may arise from serum contamination of the pigment epithelial preparation. Binding affinity in retina is high with possibly two classes of retinol binding sites present of KD about 1 - 10(-9) and 4 - 10(-8).  相似文献   

11.
A macromolecular component in the liver cytosol from laying hens as well as roosters, protein in nature and sedimenting at 4S, was shown to bind oestradiol. The dissociation constant (Kd) of the complex is approximately 5 X 10(-6)M. No binding component with a higher affinity for oestradiol was detectable in the cytosol. The binding is specific for the tissue and hormone, with the exception that progesterone also shows some affinity for this 4S component. The number of binding sites is about 330 pmol/mg cytosol protein. This number is not altered significantly after treatment of a rooster with oestrogen (24 h) or with cycloheximide (3 h). The cytoplasmic complex (oestradiol-4S-component) does not enhance the binding of oestradiol to the chromatin from rooster liver. The nuclear complex (oestradiol bound to the soluble nuclear receptor seems to be more effective in doing so.  相似文献   

12.
M A Peters  T A Walker  N R Pace 《Biochemistry》1982,21(10):2329-2335
Limited digestion of mouse 5.8S ribosomal RNA (rRNA) with RNase T2 generates 5'- and 3'-terminal "half-molecules". These fragments are capable of independently and specifically binding to 28S rRNA, so there exist at least two contacts in the 5.8S rRNA for the 28S rRNA. The dissociation constants for the 5.8S/28S, 5' 5.8S fragment/28S, and 3' 5.8S fragment/28S complexes are 9 x 10(-8) M, 6 x 10(-8) M, and 13 x 10(-8) M, respectively. Thus, each of the fragment binding sites contributes about equally to the overall binding energy of the 5.8S/28S rRNA complex, and the binding sites act independently, rather than cooperatively. The dissociation constants suggest that the 5.8S rRNA termini from short, irregular helices with 28S rRNA. Thermal denaturation data on complexes containing 28S rRNA and each of the half-molecules of 5.8S rRNA indicate that the 5'-terminal binding site(s) exist(s) in a single conformation while the 3'-terminal site exhibits two conformational alternatives. The functional significance of the different conformational states is presently indeterminate, but the possibility they may represent alternative forms of a conformational switch operative during ribosome function is discussed.  相似文献   

13.
A lambda-repressor mutant, S228N, which is defective in tetramer formation in the free state but retains full cooperativity, was studied in detail. Isolated single operator-bound S228N repressor shows association properties similar to those of the wild-type repressor. Fluorescence anisotropy studies with dansyl chloride-labeled repressor show a dimer-monomer dissociation constant of around 10(-5) M. The structure of the mutant repressor was studied by circular dichroism, acrylamide quenching and sulfhydryl reactivity at protein concentrations of < or =10(-6) M, where it is predominantly monomeric. The results suggest no significant perturbations in the structure of the S228N mutant repressor from that of the wild-type repressor. Urea denaturation studies also indicate no significant change in the stability of the repressor. The results were used to calculate energetics of loop formation in the cooperative binding process.  相似文献   

14.
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   

15.
Deuterium nuclear magnetic resonance is used to delineate the molecular dynamics of sugars bound to a lectin. 2H spin-spin relaxation times (from linewidth measurements) and reorientational correlation times are determined for N-acetylglucosamine specifically-labeled with 2H in the N-acetyl group and at carbon-3 of the pyranoside ring, in the presence and absence of wheat germ agglutinin. The correlation time for the 2H-label of N-acetylglucosamine-3-2H in the bound state is the same as that of the protein (3 X 10(-8)S), indicating that the six-membered ring has negligible motional freedom relative to the protein. The correlation time for the C2H3 group of N-acetyl-2H3-glucosamine (1.7 X 10(-9)S) shows that the N-acetyl side chain is also immobilized in the binding site, the only motion available being rotation of the C2H3 group about its threefold axis.  相似文献   

16.
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   

17.
Hillier W  Wydrzynski T 《Biochemistry》2000,39(15):4399-4405
The first determinations of substrate water binding to the O(2) evolving complex in photosystem II as a complete function of the S states have been made. H(2)(18)O was rapidly injected into spinach thylakoid samples preset in either the S(0), S(1), S(2), or S(3) states, and the rate of (18)O incorporation into the O(2) produced was determined by time-resolved mass spectrometry. For measurements at m/e = 34 (i.e., for the (16)O(18)O product), the rate of (18)O incorporation in all S states shows biphasic kinetics, reflecting the binding of the two substrate water molecules to the catalytic site. The slow phase kinetics yield rate constants at 10 degrees C of 8 +/- 2, 0.021 +/- 0.002, 2.2 +/- 0.3, and 1.9 +/- 0.2 s(-1) for the S(0), S(1), S(2), and S(3) states, respectively, while the fast phase kinetics yield a rate constant of 36.8 +/- 1.9 s(-1) for the S(3) state but remain unresolvable (>100 (s-1)) for the S(0), S(1), and S(2) states. Comparisons of the (18)O exchange rates reveal that the binding affinity for one of the substrate water molecules first increases during the S(0) to S(1) transition, then decreases during the S(1) to S(2) transition, but stays the same during the S(2) to S(3) transition, while the binding affinity for the second substrate water molecule undergoes at least a 5-fold increase on the S(2) to S(3) transition. These findings are discussed in terms of two independent Mn(III) substrate binding sites within the O(2) evolving complex which are separate from the component that accumulates the oxidizing equivalents. One of the Mn(III) sites may only first bind a substrate water molecule during the S(2) to S(3) transition.  相似文献   

18.
Elucidating the molecular mechanisms regulating multimodularity is a challenging task. Paenibacillus barcinonensis Xyn10C is a 120-kDa modular enzyme that presents the CBM22/GH10/CBM9 architecture found in a subset of large xylanases. We report here the three-dimensional structure of the Xyn10C N-terminal region, containing the xylan-binding CBM22-1–CBM22-2 tandem (Xyn10C-XBD), which represents the first solved crystal structure of two contiguous CBM22 modules. Xyn10C-XBD is folded into two separate CBM22 modules linked by a flexible segment that endows the tandem with extraordinary plasticity. Each isolated domain has been expressed and crystallized, and their binding abilities have been investigated. Both domains contain the R(W/Y)YYE motif required for xylan binding. However, crystallographic analysis of CBM22-2 complexes shows Trp-308 as an additional binding determinant. The long loop containing Trp-308 creates a platform that possibly contributes to the recognition of precise decorations at subsite S2. CBM22-2 may thus define a subset of xylan-binding CBM22 modules directed to particular regions of the polysaccharide. Affinity electrophoresis reveals that Xyn10C-XBD binds arabinoxylans more tightly, which is more apparent when CBM22-2 is tested against highly substituted xylan. The crystal structure of the catalytic domain, also reported, shows the capacity of the active site to accommodate xylan substitutions at almost all subsites. The structural differences found at both Xyn10C-XBD domains are consistent with the isothermal titration calorimetry experiments showing two sites with different affinities in the tandem. On the basis of the distinct characteristics of CBM22, a delivery strategy of Xyn10C mediated by Xyn10C-XBD is proposed.  相似文献   

19.
Binding experiments with the cytosolic androgen receptor from murine skeletal muscle yield with testosterone a biphasic saturation curve and a biphasic Scatchard plot. These binding characteristics result from the conversion of 8 S receptor (KD = 1,4 X 10(-10) M) into 4-5 S receptor (KD = 1,2 X 10(-9) M). This conversion is androgen dependent and is facilitated in vitro by either UV-irradiation or by methods known to activate steroid hormone receptor complexes to a nuclear binding form (e.g. high ionic strength or elevated temperature). The measured data show that both receptor forms are in a complex dissociation equilibrium. The reassociation of the 4-5 S receptor to form the 8 S complex is inhibited by RNase.  相似文献   

20.
Murine myeloma immunoglobulin (IgA, K) J539, which shows enhanced tryptophanyl fluorescence on ligand binding, and S10, which shows reverse-sign changes in tryptophanyl fluorescence on ligand binding (RLIF, see below), have been reduced, alkylated, and dissociated into their light (L) and heavy (H) chains. Two hybrid recombinants, H10L539 and H539L10, have been prepared and the 7S material has been isolated by chromatography. The binding behavior of these recombinants was studied with a number of ligands. Both recombinants showed activity with beta(1 leads to 6) linked galactose ligands comparable to the native immunoglobulins. The ligand-induced fluorescence changes of the recombinants paralleled those of the heavy chain donor. For the recombinant H10L539, two different galactose-ligands caused fluorescence changes in opposite directions. It was quantitatively shown that binding of these ligands, nevertheless, took place in the same combining region. The idiotype of each recombinant resembled that of the heavy chain donor.  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