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.
The solution NMR structure is reported for Ca(2+)-loaded S100B bound to a 12-residue peptide, TRTK-12, from the actin capping protein CapZ (alpha1 or alpha2 subunit, residues 265-276: TRTKIDWNKILS). This peptide was discovered by Dimlich and co-workers by screening a bacteriophage random peptide display library, and it matches exactly the consensus S100B binding sequence ((K/R)(L/I)XWXXIL). As with other S100B target proteins, a calcium-dependent conformational change in S100B is required for TRTK-12 binding. The TRTK-12 peptide is an amphipathic helix (residues W7 to S12) in the S100B-TRTK complex, and helix 4 of S100B is extended by three or four residues upon peptide binding. However, helical TRTK-12 in the S100B-peptide complex is uniquely oriented when compared to the three-dimensional structures of other S100-peptide complexes. The three-dimensional structure of the S100B-TRTK peptide complex illustrates that residues in the S100B binding consensus sequence (K4, I5, W7, I10, L11) are all involved in the S100B-peptide interface, which can explain its orientation in the S100B binding pocket and its relatively high binding affinity. A comparison of the S100B-TRTK peptide structure to the structures of apo- and Ca(2+)-bound S100B illustrates that the binding site of TRTK-12 is buried in apo-S100B, but is exposed in Ca(2+)-bound S100B as necessary to bind the TRTK-12 peptide.  相似文献   

2.
Trapp O  Schurig V 《Chirality》2002,14(6):465-470
The axially chiral allenes dimethyl-1,3-allenedicarboxylate 1 and diethyl-1,3-allenedicarboxylate 2 show characteristic plateau formation during enantioselective GC separation on the chiral stationary liquid phase Chirasil-beta-Dex. The elution profiles, obtained from temperature-dependent dynamic GC (DGC) experiments (1: 100-140 degrees C; 2: 110-150 degrees C) were evaluated with the recently derived approximation function (AF) k1(approx) = f(t(R)(A),t(R)(B),w(h)(A),h(plateau), N) to yield the enantiomerization rate constant directly k(1). These values were compared with those obtained by computer-aided simulation with ChromWin. The Eyring activation parameters of the experimental interconversion profiles were determined to be: DeltaG(#)(298.15 K) = 103.6 +/- 0.9 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.4 kJ mol(-1), DeltaS(#) = -198 +/- 7 J K(1) mol(-1) for dimethyl-1,3-allenedicarboxylate 1, and DeltaG(#)(298.15 K) = 103.5 +/- 1.1 kJ mol(-1), DeltaH(#) = 44.7 +/- 0.5 kJ mol(-1), DeltaS(#) = -197 +/- 9 J K(-1) mol(-1) for diethyl-1,3-allenedicarboxylate 2. The approximation function (AF) presented here allows the fast determination of rate constants k(1) and activation barriers of enantiomerization DeltaG(#) from chromatographic parameters without extensive computer simulation.  相似文献   

3.
We present herein the partitioning characteristics of anti-Salmonella and anti-Escherichia coli O157 immunomagnetic beads (IMB) with respect to the nonspecific adsorption of several nontarget food-borne organisms with and without an assortment of well-known blocking agents, such as casein, which have been shown to be useful in other immunochemical applications. We found several common food-borne organisms that strongly interacted with both types of IMB, especially with anti-Salmonella form (av DeltaG0=-20 +/- 4 kJ mol(-1)) even in the presence of casein [1% (w/v): DeltaG0=-18 +/- 3 kJ mol(-1); DeltaDeltaG0 approximately -2 kJ mol(-1)]. However, when one of the most problematic organisms (a native K12-like E. coli isolate; DeltaG0=-19 +/- 2 kJ mol(-1)) was tested for nonspecific binding in the presence of iota-carrageenan (0.03-0.05%), there was an average decline of ca. 90% in the equilibrium capture efficiency xi (DeltaG0=-11 +/- 4 kJ mol(-1); DeltaDeltaG0 approximately -8 kJ mol(-1)). Other anionic polysaccharides (0.1% kappa-carrageenan and polygalacturonic acid) had no significant effect (av DeltaG0=-19 +/- 1 kJ mol(-1); DeltaDeltaG0 approximately 0 kJ mol(-1)). Varying iota-carrageenan from 0% to 0.02% resulted in xi significantly diminishing from 0.69 (e.g., 69% of the cells captured; DeltaG0=-19 +/- 3 kJ mol(-1)) to 0.05 (DeltaG0=-11 +/- 2 kJ mol(-1); DeltaDeltaG0 approximately -9 kJ mol(-1)) at about 0.03% iota-carrageenan where xi leveled off. An optimum blocking ability was achieved with 0.04% iota-carrageenan suspended in 100 mM phosphate buffer. We also demonstrated that the utilization of iota-carrageenan as a blocking agent causes no great loss in the IMBs capture efficiency with respect to the capture of its target organisms, various salmonellae.  相似文献   

4.
Manganese(II) activation of the methionyl aminopeptidases from Escherichia coli (EcMetAP-I) and the hyperthermophilic archaeon Pyrococcus furiosus (PfMetAP-II) was investigated. Maximum catalytic activity for both enzymes was obtained with 1 equiv of Mn(II), and the dissociation constants (K(d)) for the first metal binding site were found to be 6 +/- 0.5 and 1 +/- 0.5 microM for EcMetAP-I and PfMetAP-II, respectively. These K(d) values were verified by isothermal titration calorimetry (ITC) and found to be 3.0 +/- 0.2 and 1.4 +/- 0.2 microM for EcMetAP-I and PfMetAP-II, respectively. The hydrolysis of MGMM was measured in triplicate between 25 and 85 degrees C at eight substrate concentrations ranging from 2 to 20 mM for PfMetAP-II. Both specific activity and K(m) values increased with increasing temperature. An Arrhenius plot was constructed from the kcat values and was found to be linear over the temperature range 25-85 degrees C. The activation energy for the Mn(II)-loaded PfMetAP-II hydrolysis of MGMM was found to be 25.7 kJ/mol while the remaining thermodynamic parameters calculated at 25 degrees C are DeltaG+ = 50.1 kJ/mol, DeltaH+ = 23.2 kJ/mol, and DeltaS++ = -90.2 J x mol(-1) x K(-1).  相似文献   

5.
S100A1 is a typical representative of a group of EF-hand calcium-binding proteins known as the S100 family. The protein is composed of two alpha subunits, each containing two calcium-binding loops (N and C). At physiological pH (7.2) and NaCl concentration (100 mm), we determined the microscopic binding constants of calcium to S100A1 by analysing the Ca(2+)-titration curves of Trp90 fluorescence for both the native protein and its Glu32 --> Gln mutant with an inactive N-loop. Using a chelator method, we also determined the calcium-binding constant for the S100A1 Glu73 --> Gln mutant with an inactive C-loop. The protein binds four calcium ions in a noncooperative way with binding constants of K(1) =4 +/- 2 x 10(3) m(-1) (C-loops) and K(2) approximately 10(2) m(-1) (N-loops). Only when both loops are saturated with calcium does the protein change its global conformation, exposing to the solvent hydrophobic patches, which can be detected by 2-p-toluidinylnaphthalene-6-sulfonic acid - a fluorescent probe of protein-surface hydrophobicity. S-Glutathionylation of the single cysteine residue (85) of the alpha subunits leads to a 10-fold increase in the affinity of the protein C-loops for calcium and an enormous - four orders of magnitude - increase in the calcium-binding constants of its N-loops, owing to a cooperativity effect corresponding to DeltaDeltaG = -6 +/- 1 kcal.mol(-1). A similar effect is observed upon formation of the mixed disulfide with cysteine and 2-mercaptoethanol. The glutathionylated protein binds TRTK-12 peptide in a calcium-dependent manner. S100A1 protein can act, therefore, as a linker between the calcium and redox signalling pathways.  相似文献   

6.
The Alzheimer-linked neural protein S100B is a signaling molecule shown to control the assembly of intermediate filament proteins in a calcium-sensitive manner. Upon binding calcium, a conformational change occurs in S100B exposing a hydrophobic surface for target protein interactions. The synthetic peptide TRTK-12 (TRTKIDWNKILS), derived from random bacteriophage library screening, bears sequence similarity to several intermediate filament proteins and has the highest calcium-dependent affinity of any target molecule for S100B to date (K(d) <1 microm). In this work, the three-dimensional structure of the Ca(2+)-S100B-TRTK-12 complex has been determined by NMR spectroscopy. The structure reveals an extended, contiguous hydrophobic surface is formed on Ca(2+)-S100B for target interaction. The TRTK-12 peptide adopts a coiled structure that fits into a portion of this surface, anchored at Trp(7), and interacts with multiple hydrophobic contacts in helices III and IV of Ca(2+)-S100B. This interaction is strikingly different from the alpha-helical structures found for other S100 target peptides. By using the TRTK-12 interaction as a guide, in combination with other available S100 target structures, a recognition site on helix I is identified that may act in concert with the TRTK-12-binding site from helices III and IV. This would provide a larger, more complex site to interact with full-length target proteins and would account for the promiscuity observed for S100B target protein interactions.  相似文献   

7.
In solution, S100B protein is a noncovalent homodimer composed of two subunits associated in an antiparallel manner. Upon calcium binding, the conformation of S100B changes dramatically, leading to the exposure of hydrophobic residues at the surface of S100B. The residues in the C-terminal domain of S100B encompassing Phe(87) and Phe(88) have been implicated in interaction with target proteins. In this study, we used two-hybrid technology to identify specific S100B target proteins. Using S100B as bait, we identify S100A6 and S100A11 as specific targets for S100B. S100A1, the closest homologue of S100B, is capable of interaction with S100B but does not interact with S100A6 or S100A11. S100B, S100A6, and S100A11 isoforms are co-regulated and co-localized in astrocytoma U373 cells. Furthermore, co-immunoprecipitation experiments demonstrated that Ca(2+)/Zn(2+) stabilizes S100B-S100A6 and S100B-S100A11 heterocomplexes. Deletion of the C-terminal domain or mutation of Phe(87) and Phe(88) residues has no effect on S100B homodimerization and heterodimerization with S100A1 but drastically decreases interaction between S100B and S100A6 or S100A11. Our data suggest that the interaction between S100B and S100A6 or S100A11 should not be viewed as a typical S100 heterodimerization but rather as a model of interaction between S100B and target proteins.  相似文献   

8.
The kinetics of charge recombination between the primary photoxidized donor (P(+)) and the secondary reduced quinone acceptor (Q(B)(-)) have been studied in reaction centers (RCs) from the purple photosynthetic bacterium Rhodobacter sphaeroides incorporated into lecithin vesicles containing large ubiquinone pools over the temperature range 275 K = (50 +/- 15) nm). Following these premises, we describe the kinetics of P(+)Q(B)(-) recombination with a truncated cumulant expansion and relate it to P(Q) and to the free energy changes for Q(A)(-)Q(B) --> Q(A)Q(B)(-) electron transfer (DeltaG(AB)(o)) and for quinone binding (DeltaG(bind)(o)) at Q(B). The model accounts well for the temperature and quinone dependence of the charge recombination kinetics, yielding DeltaG(AB)(o) = -7.67 +/- 0.05 kJ mol(-1) and DeltaG(bind)(o) = -14.6 +/- 0.6 kJ mol(-1) at 298 K.  相似文献   

9.
S100B(betabeta) is a dimeric Ca2+-binding protein that is known to inhibit the protein kinase C (PKC)-dependent phosphorylation of several proteins. To further characterize this inhibition, we synthesized peptides based on the PKC phosphorylation domains of p53 (residues 367-388), neuromodulin (residues 37-53), and the regulatory domain of PKC (residues 19-31), and tested them as substrates for PKC. All three peptides were shown to be good substrates for the catalytic domain of PKC. As for full-length p53 (Baudier J, Delphin C, Grunwald D, Khochbin S, Lawrence JJ. 1992. Proc Natl Acad Sci USA 89:11627-11631), S100B(betabeta) binds the p53 peptide and inhibits its PKC-dependent phosphorylation (IC50 = 10 +/- 7 microM) in a Ca2+-dependent manner. Similarly, phosphorylation of the neuromodulin peptide and the PKC regulatory domain peptide were inhibited by S100B(betabeta) in the presence of Ca2+ (IC50 = 17 +/- 5 microM; IC50 = 1 +/- 0.5 microM, respectively). At a minimum, the C-terminal EF-hand Ca2+-binding domain (residues 61-72) of each S100beta subunit must be saturated to inhibit phosphorylation of the p53 peptide as determined by comparing the Ca2+ dependence of inhibition ([Ca]IC50 = 29.3 +/- 17.6 microM) to the dissociation of Ca2+ from the C-terminal EF-hand Ca2+-binding domain of S100B(betabeta).  相似文献   

10.
1. Stellation of astrocytes in culture involves a complex rearrangement of microfilaments, intermediate filaments, and microtubules, which reflects in part the plasticity of these cells observed during development or after injury.2. An astrocytic calcium-binding protein, S100B, has been implicated in the regulation of plasticity due to its ability to interact with cytoskeletal proteins.3. We used digitonin-permeabilized astrocytes to introduce TRTK-12, a peptide that binds to the C-terminal of S100B and blocks its interaction with cytoskeletal proteins.4. TRTK-12 was able to block cAMP-induced astrocyte stellation and this effect was dependent on the concentration of the peptide. These results support the idea that S100B has a modulatory role on astrocyte morphology.  相似文献   

11.
P Mucha  A Szyk  P Rekowski  P A Weiss  P F Agris 《Biochemistry》2001,40(47):14191-14199
The contributions of the natural modified nucleosides to RNA identity in protein/RNA interactions are not understood. We had demonstrated that 15 amino acid long peptides could be selected from a random phage display library using the criterion of binding to a modified, rather than unmodified, anticodon domain of yeast tRNA(Phe) (ASL(Phe)). Affinity and specificity of the selected peptides for the modified ASL(Phe) have been characterized by fluorescence spectroscopy of the peptides' tryptophans. One of the peptides selected, peptide t(F)2, exhibited the highest specificity and most significant affinity for ASL(Phe) modified with 2'-O-methylated cytidine-32 and guanosine-34 (Cm(32) and Gm(34)) and 5-methylated cytidine-40 (m(5)C(40)) (K(d) = 1.3 +/- 0.4 microM) and a doubly modified ASL(Phe)-Gm(34),m(5)C(40) and native yeast tRNA(Phe) (K(d) congruent with 2.3 and 3.8 microM, respectively) in comparison to that for the unmodified ASL(Phe) (K(d) = 70.1 +/- 12.3 microM). Affinity was reduced when a modification altered the ASL loop structure, and binding was negated by modifications that disfavored hairpin formation. Peptide t(F)2's higher affinity for the ASL(Phe)-Cm(32),Gm(34),m(5)C(40) hairpin and fluorescence resonance energy transfer from its tryptophan to the hypermodified wybutosine-37 in the native tRNA(Phe) placed the peptide across the anticodon loop and onto the 3'-side of the stem. Inhibition of purified yeast phenylalanyl-tRNA synthetase (FRS) catalyzed aminoacylation of cognate yeast tRNA(Phe) corroborated the peptide's binding to the anticodon domain. The phage-selected peptide t(F)2 has three of the four amino acids crucial to G(34) recognition by the beta-structure of the anticodon-binding domain of Thermus thermophilus FRS and exhibited circular dichroism spectral properties characteristic of beta-structure. Thus, modifications as simple as methylations contribute identity elements that a selected peptide specifically recognizes in binding synthetic and native tRNA and in inhibiting tRNA aminoacylation.  相似文献   

12.
Chatani E  Nonomura K  Hayashi R  Balny C  Lange R 《Biochemistry》2002,41(14):4567-4574
To clarify the structural role of Phe46 inside the hydrophobic core of bovine pancreatic ribonuclease A (RNase A), thermal and pressure unfolding of wild-type RNase A and three mutant forms (F46V, F46E, and F46K) were analyzed by fourth-derivative UV absorbance spectroscopy. All the mutants, as well as the wild type, exhibited a two-state transition during both thermal and pressure unfolding, and both T(m) and P(m) decreased markedly when Phe46 was replaced with valine, glutamic acid, or lysine. The strongest effect was on the F46K mutant and the weakest on F46V. Both unfolding processes produced identical blue shifts in the fourth-derivative spectra, indicating that the tyrosine residues are similarly exposed in the temperature- and pressure-induced unfolded states. A comparison of Gibbs free energies determined from the pressure and temperature unfoldings, however, gave DeltaG(p)/DeltaG(t) ratios (r) of 1.7 for the wild type and 0.92 +/- 0.03 for the mutants. Furthermore, the DeltaV value for each mutant was larger than that for the wild type. CD spectra and activity measurements showed no obvious major structural differences in the folded state, indicating that the structures of the Phe46 mutants and wild type differ in the unfolded state. We propose a model in which Phe46 stabilizes the hydrophobic core at the boundary between two structural domains. Mutation of Phe46 decreases protein stability by weakening the unfolding cooperativity between these domains. This essential function of Phe46 in RNase A stability indicates that it belongs to a chain-folding initiation site.  相似文献   

13.
The spike H protein of bacteriophage phiX174 was prepared as a hexa histidine-tagged fusion (HisH). On enzyme-linked plate assaying, HisH was found to bind specifically to the lipopolysaccharides (LPSs) of phiX174-sensitive strains, Escherichia coli C and Salmonella typhimurium Ra chemotype, having the complete oligosaccharide sequence of the R-core on the LPSs. In sharp contrast, HisH bound weakly to the LPSs of phiX174-insensitive strains, i.e. E. coli F583 (Rd(2)) lacking some terminal saccharides and E. coli O111: B4 (smooth strain) having additional O-repeats on the R-core. The fluorescence spectra of HisH changed dose-dependently in the case of the LPS of E. coli C, the intensity increasing and the emission peak shifting to the shorter wavelength side, which was attributable to the hydrophobic interaction of HisH with the LPS. The binding equilibrium was analyzed by fluorometric titration to determine the dissociation constant K(d), 7.02 +/- 0.37 microM, and the Gibbs free energy change DeltaG(0), -29.1 kJ mol(-1) (at 22 degrees C, pH 7.4). Based on the temperature dependence of (K)d in a van't Hoff plot, the standard enthalpy change DeltaH(0) and the entropy change DeltaS(0) were calculated to be +23.7 kJ mol(-1) and 179 J mol(-1) K(-1) at 22 degrees C, respectively, and this binding was thereby concluded to be an entropy-driven reaction.  相似文献   

14.
The effects of electrostatic substitutions on the spectroscopic, ligand binding, and redox properties of the heme in leghemoglobin have been examined by replacement of the proximal leucine 88 residue with an aspartic acid residue (Leu88Asp). Electronic and resonance Raman spectra of the ferric derivative of Leu88Asp indicate a mixture of 6-coordinate, high-spin and 6-coordinate, low-spin hemes, analogous to that observed in the recombinant wild-type protein (rLb). At alkaline pH, formation of hydroxide-bound heme is indicated for Leu88Asp; the pK(a) for this transition (8.7 +/- 0.2, micro = 0.10 M, 25.0 degrees C) is 0.4 pH units higher than for rLb. Equilibrium dissociation constants (sodium phosphate, pH 7.0, micro = 0.10 M, 25.0 +/- 0.1 degrees C) for binding of anionic ligands (N(-)(3), nicotinate) to Leu88Asp are higher (K(d,nicotinate) = 6.8 +/- 0.2 microM; K(d,azide) = 33 +/- 0.6 microM) than the corresponding values for rLb (K(d,nicotinate) = 1.4 +/- 0.3 microM (pH 5.5, micro = 0.10 M, 25.0 +/- 0.1 degrees C); K(d,azide) = 4.8 +/- 0.2 microM). Resonance Raman spectra (sodium phosphate, pH 7.0, micro = 0.10 M) for the ferrous derivatives of Leu88Asp and rLb exhibit a strong nu(Fe-His) stretching frequency at 223 cm(-1) in both cases, indicating that the hydrogen bonding structure on the proximal side is not substantially altered in the variant. The reduction potential of Leu88Asp is -14 +/- 2 mV vs standard hydrogen electrode (SHE) (25.0 degrees C, micro = 0.10 M, pH 7.0), a decrease of 35 mV over the corresponding value for the wild-type protein under the same conditions (21 +/- 3 mV vs SHE). An assessment of these data in terms of electrostatic and hydrogen bonding considerations is presented.  相似文献   

15.
It is generally accepted that P-glycoprotein binds its substrates in the lipid phase of the membrane. Quantification and characterization of the lipid-transporter binding step are, however, still a matter of debate. We therefore selected 15 structurally diverse drugs and measured the binding constants from water to the activating (inhibitory) binding region of P-glycoprotein, K(tw(1)) (K(tw(2))), as well as the lipid-water partition coefficients, K(lw). The former were obtained by measuring the concentrations of half-maximum activation (inhibition), K(1) (K(2)), in living NIH-MDR-G185 mouse embryo fibroblasts using a Cytosensor microphysiometer, and the latter were derived from surface activity measurements. This allowed determination of the membrane concentration of drugs at half-maximum P-glycoprotein activation (C(b(1)) = (0.02 to 67) mmol/L lipid), which is much higher than the corresponding aqueous concentration (K(1) = (0.02 to 376) microM). Moreover we determined the free energy of drug binding from water to the activating binding region of the transporter (DeltaG degrees (tw(1)) = (-30 to -54) kJ/mol), the free energy of drug partitioning into the lipid membrane (DeltaG degrees (lw) = (-23 to -34) kJ/mol), and, as the difference of the two, the free energy of drug binding from the lipid membrane to the activating binding region of the transporter (DeltaG degrees (tl(1)) = (-7 to -27) kJ/mol). For the compounds tested DeltaG degrees (tl(1)) was less negative than DeltaG degrees (lw) but varied more strongly. The free energies of substrate binding to the transporter within the lipid phase, DeltaG degrees (tl(1)), are consistent with a modular binding concept, where the energetically most efficient binding module comprises two hydrogen bond acceptor groups.  相似文献   

16.
The thermodynamics of the conversion of aqueous L-aspartic acid to fumaric acid and ammonia have been investigated using both heat conduction microcalorimetry and high-pressure liquid chromatography. The reaction was carried out in aqueous phosphate buffer over the pH range 7.25-7.43, the temperature range 13-43 degrees C, and at ionic strengths varying from 0.066 to 0.366 mol kg(-1). The following values have been found for the conversion of aqueous L-aspartateH- to fumarate2- and NH4+ at 25 degrees C and at zero ionic strength: K = (1.48 +/- 0.10) x 10(-3), DeltaG degrees = 16.15 +/- 0.16 kJ mol(-1), DeltaH degrees = 24.5 +/- 1.0 kJ mol(-1), and DeltaC(p) degrees = -147 +/- 100 J mol(-1) K(-1). Calculations have also been performed which give values of the apparent equilibrium constant for the conversion of L-aspartic acid to fumaric acid and ammonia as a function of temperature, pH and ionic strength.  相似文献   

17.
The thermodynamics of the conversion of aqueous D-psicose to D-allose has been investigated using high-pressure liquid chromatography. The reaction was carried out in phosphate buffer at pH 7.4 over the temperature range 317.25-349.25 K. The following results are obtained for the conversion process at 298.15 K: DeltaG degrees = - 1.41 +/- 0.09 kJ mol(-1), DeltaH degrees = 7.42 +/- 1.7 kJ mol(-1), and DeltaC(p) degrees = 67 +/- 50 J mol(-1) K(-1). An approximate equilibrium constant of 0.30 is obtained at 333.15 K for the conversion of aqueous D-psicose to D-altrose. Available thermodynamic data for isomerization reactions involving aldohexoses and aldopentoses are summarized.  相似文献   

18.
The agonist, [3H](-)[S]-1-(2-amino-2-carboxyethyl)-5-fluoro-pyrimidine-2,4-dione ([3H](S)F-Willardiine) binding to functional alpha-amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA) receptors of resealed plasma membrane vesicles and nerve endings freshly isolated from the rat cerebral cortex displayed two binding sites (K(D1)=33+/-7 nM, B(MAX1)=1.6+/-0.3 pmol/mg protein, K(D2)=720+/-250 nM and B(MAX2)=7.8+/-4.0 pmol/mg protein). The drug which impairs AMPA receptor desensitisation, 6-chloro-3,4-dihydro-3-(2-norbornene-5-yl)-2H-1,2,4-benzothiadiazine-7-sulphonamide-1,1-dioxide (cyclothiazide, CTZ) fully displaced the [3H](S)F-Willardiine binding at a concentration of 500 microM. In the presence of 100 microM CTZ (K(I(CTZ))=60+/-6 microM), both the antagonist [3H]-1,2,3,4-tetrahydro-6-nitro-2,3-dioxo-benzo(F)quinoxaline-7-sulfonamide ([3H]NBQX: K(D)=24+/-4 nM, B(MAX)=12.0+/-0.1 pmol/mg protein) and the high-affinity agonist binding showed similar affinity reduction ([3H](S)F-Willardiine: K(D)=140+/-19 nM, B(MAX)=2.9+/-0.5 pmol/mg protein; [3H]NBQX: K(D)=111+/-34 nM, B(MAX)=12+/-3 pmol/mg protein). To disclose structural correlates underlying genuine allosteric binding interactions, molecular mechanics calculations of CTZ-induced structural changes were performed with the use of PDB data on extracellular GluR2 binding domain dimeric crystals available by now. Hydrogen-bonding and root mean square (rms) values of amino acid residues recognising receptor agonists showed minor alterations in the agonist binding sites itself. Moreover, CTZ binding did not affect dimeric subunit structures significantly. These findings indicated that the structural changes featuring the non-desensitised state could possibly occur to a further site of the extracellular GluR2 binding domain. The increase of agonist efficacy on allosteric CTZ binding may be interpreted in terms of a mechanism involving AMPA receptor desensitisation sequential to activation.  相似文献   

19.
ATP-sensitive potassium (K(ATP)) channels are composed of an ATP-binding cassette (ABC) protein (SUR1, SUR2A or SUR2B) and an inwardly rectifying K(+) channel (Kir6.1 or Kir6.2). Like other ABC proteins, the nucleotide binding domains (NBDs) of SUR contain a highly conserved "signature sequence" (the linker, LSGGQ) whose function is unclear. Mutation of the conserved serine to arginine in the linker of NBD1 (S1R) or NBD2 (S2R) did not alter the ability of ATP or ADP (100 microM) to displace 8-azido-[(32)P]ATP binding to SUR1, or abolish ATP hydrolysis at NBD2. We co-expressed Kir6.2 with wild-type or mutant SUR in Xenopus oocytes and recorded the resulting currents in inside-out macropatches. The S1R mutation in SUR1, SUR2A or SUR2B reduced K(ATP) current activation by 100 microM MgADP, whereas the S2R mutation in SUR1 or SUR2B (but not SUR2A) abolished MgADP activation completely. The linker mutations also reduced (S1R) or abolished (S2R) MgATP-dependent activation of Kir6.2-R50G co-expressed with SUR1 or SUR2B. These results suggest that the linker serines are not required for nucleotide binding but may be involved in transducing nucleotide binding into channel activation.  相似文献   

20.
We used an H2-purging culture vessel to replace an H2-consuming syntrophic partner, allowing the growth of pure cultures of Syntrophothermus lipocalidus on butyrate and Aminobacterium colombiense on alanine. By decoupling the syntrophic association, it was possible to manipulate and monitor the single organism's growth environment and determine the change in Gibbs free energy yield (DeltaG) in response to changes in the concentrations of reactants and products, the purging rate, and the temperature. In each of these situations, H2 production changed such that DeltaG remained nearly constant for each organism (-11.1 +/- 1.4 kJ mol butyrate(-1) for S. lipocalidus and -58.2 +/- 1.0 kJ mol alanine(-1) for A. colombiense). The cellular maintenance energy, determined from the DeltaG value and the hydrogen production rate at the point where the cell number was constant, was 4.6 x 10(-13) kJ cell(-1) day(-1) for S. lipocalidus at 55 degrees C and 6.2 x 10(-13) kJ cell(-1) day(-1) for A. colombiense at 37 degrees C. S. lipocalidus, in particular, seems adapted to thrive under conditions of low energy availability.  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