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.
Decomposition of phenyl acridinium-9-carboxylate is monitored using electrogenerated chemiluminescence in a flow system. The formation of the pseudobase from the acridinium ester [AE] is described by rate = k1[AE] + k1[AE][OH?]0.5, where k1 = 0.020 ± 0.006 s?1 and k1 = 2.1 ± 0.8 (L/mol)?0.5 s?1. Irreversible decomposition of the pseudobase is described by rate = k2[AE][OH?], where k2 = 20.1 ± 3.8 (L/mol s). These kinetic equations, plus measurement of variation in emission intensity for constant acridinium ester concentration, are used to predict the resulting emission intensity v. pH behaviour given various contact times (in the 0.25 to 25 s range) for the acridinium ester to be in an alkaline solution prior to initiation of the chemiluminescence reaction.  相似文献   

2.
The antifungal activity of 441-acyl derivatives of 3-(3,5-dichlorophenyl)-2,4-imidazol- idinedione against Botrytis cinerea, and of 10 1-sulfonyl compounds against Aiternaria kikuchi- ana were assayed by the agar medium dilution method. The structure-activity relationships for the substituents of the acyl and sulfonyl moieties were analyzed with such physicochemical parameters as hydrophobic π, inductive electronic σ1, and steric Ecs and B1 values by multiple regression. The activity of the acyl derivatives against B. cinerea was related parabolically to the hydrophobicity of the substituents. The stronger the electron-donating power, the larger the overall steric bulkiness, and the smaller the minimum width in the direction perpendicular to the bond axis of the substituents, the greater was the activity. The activity of the sulfonyl derivatives against A. kikuciana was related only to the hydrophobicity of the substituents.  相似文献   

3.
Abstract

Substances eliciting, at very high concentrations, a lower maximal response of a particular biological system than a defined standard, are defined as partial agonists. The convention rests on the definition of a standard substance that achieves a ‘full’ maximal response; partial agonism being, therefore, relative. Various mechanisms lie behind this phenomenon: 1. Receptor-related mechanisms: the agonist-receptor complex exists in several conformational states from which only one, or only a few, activate the cell signaling pathway. This may occur when the receptor itself, or the agonist, exists in multiple states (e.g., in the form of enantiomers or stereoisomers), or when the agonist-receptor complex changes its conformation (receptor switch: two-state model of receptor activation). Furthermore, a steric hindrance by a ‘wrong-way binding’ of a part of the agonist's molecules may prevent the full ‘correct’ occupancy of receptors. 2. Mechanisms based on the efficacy of the stimulus-response coupling. The efficacy is then proportional to the sum of probabilities that receptors in individual states activate the cell-signaling pathway. Doses (concentrations) eliciting the half maximal response (EC50), or similar response sensitivity parameters, are not included in the definition of partial agonism. However, tight correlations exist between maximal response and EC50 in many, but not all, generic groups of agonistically acting substances. These relationships are frequently linear; intercepts and slopes of these ‘E, KE plots’ are characteristic for individual, putative mechanisms. Dose-response curves of partial agonists are akin to those obtained for a response to a full agonist after a stepwise partial inactivation of receptors by an irreversible inhibitor. Also, the E, KE plots obtained in these instances are similar to those of partial agonists. The receptor reserve, rather vaguely defined in early reports, is therefore closely linked to the phenomenon of partial agonism.  相似文献   

4.
The crystal structure of the model tripeptide Boc-Aib-Gly-Leu-OMe ( 1 ) reveals two independent molecules in the asymmetric unit that adopt “enantiomeric” type I and type I′ β-turn conformations with the Aib and Gly residues occupying the corner (i + 1 and i + 2) positions. 13C cross polarization and magic angle sample spinning spectra in the solid state also support the coexistence of two conformational species. 13C-nmr in CDCl3 establishes the presence of a single species or rapid exchange between conformations. 400 MHz 1H-nmr provides evidence for conformational exchange involving a major and minor species, with β-turn conformations supported by the low solvent exposure of Leu(3) NH and the observation of NiH ↔ Ni+1H nuclear Overhauser effects. CD bands in the region 190–230 nm are positive, supporting a major population of type I′ β-turns. The isomeric peptide, Boc-Gly-Leu-Aib-OMe ( 2 ), adopts an “open” type II′ β-turn conformation in crystals. Solid state and solution nmr support population of a single conformational species. Chiral perturbation introduced outside the folded region of peptides may provide a means of modulating screw sense in achiral sequences. © 1998 John Wiley & Sons, Inc. Biopoly 45: 191–202, 1998  相似文献   

5.
Rigid and flexible geometry calculations are described for N-acetylglycine-N′-methylamide, N-acetylalanine-N′-methylamide, and their retro-inverso analogs, bis(acetamido) methane, 1,1-bis(acetamido) ethane, N,N′-dimethylmalonamide, and N,N′-dimethyl-2-methyl-malonamide. The significance of relaxing all degrees of freedom, especially angular flexibility is demonstrated. The flexible geometry approach yields energy maps similar to those from rigid geometry, but the energy barriers between minima are substantially reduced, leading in general, to more probable transitions and a higher volume of accessible conformational space. Whereas the glycine and alanine derivatives exhibit their lowest energy minima in the C region, the gem-diaminoalkyl and malonyl residues show their lowest minima in the “α-helical” regions. With respect to the effect of side chains (H versus CH3), the greatest conformational influence appears with the gem-diaminoalkyl residues. These results indicate significantly different conformational behavior of retro peptides and the implications of these pairwise incorporations of retro-inverso residues in peptide chains, are discussed.  相似文献   

6.
Stomatal conductance and transpiration were measured concurrently in an irrigated Eucalyptus globulus Labill. plantation. Canopy stomatal conductance, canopy boundary layer conductance and the dimensionless decoupling coefficient (Ω) were calculated (a) summing the conductance of three canopy layers (gc) and (b) weighting the contribution of foliage according to the amount of radiation received (gc′). Canopy transpiration was then calculated from gc and gc′ for Ω = 1 (Eeq), Ω = 0 (Eimp) and by weighting Eeq and Eimp using Ω (EΩ). Eeq, Eimp and EΩ were compared to transpiration estimated from measurements of heat pulse velocity. The mean value of Ω was 0·63. Transpiration calculated using gc and assuming perfect coupling (12·5 ± 0·9 mmol m?2 s?1) significantly overestimated measured values (8·7 ± 0·8 mmol m?2 s?1). Good estimates of canopy transpiration were obtained either (a) calculating EΩ separately for the individual canopy layers or (b) treating the canopy as a single layer and using gc′ in a calculation of Eimp (Ω = 0). The latter approach only required measurement of stomatal conductance at a single canopy position but would be unsuitable for use in combined models of canopy transpiration and assimilation. It should however, be suitable for estimating transpiration in forests regardless of the degree of coupling.  相似文献   

7.
Abstract

1,3,5-Trimethyl-N4-hydroxycytosine, an analogue of the promutagenic N4-hydroxycytosine and 5-methyl-N4-hydroxycytosine nucleosides, crystallizes in the monoclinic space group P 21/n with cell dimensions at ?147°C: a = 7.1481(7), b = 9.2565(5), c = 13.3086(12) Å, β = 97.90(2)°, V = 872.24(13) Å3, ρc = 1.426 Mg m?3, Z = 4, F(000) = 401.39, μ = 0.91 mm?1, λ(Cu) = 1.54056 Å, 20(max) = 139.3°. The crystal structure has been solved by X-ray difraction and refined to R = 3.7 % for 1457 reflections. Notwithstandin the steric hindrance imposed by methyl groups at both N(3) and C(5), the exocyclic N4-OH group is located essentially in the plane of the ring, giving rise to an “overcrowded” molecule, like that of 1,5,N4,N4-tetramethylcytosine. The conformational parameters have also been compared with those of a number of related and previously reported N(1)-substituted cytosines. In the present compound the N4-OH rotamer is in the anti conformation relative to the ring N(3), hence similar to that of one of the rotamers in N(1)-substituted N4-hydroxycytosine, which permits normal Watson-Crick base pairing of the latter, relevant to the mechanism of hydroxylamine mutagenesis.  相似文献   

8.
Dehydroamino acids are non‐coded amino acids that offer unique conformational properties. Dehydrophenylalanine (ΔPhe) is most commonly used to modify bioactive peptides to constrain the topography of the phenyl ring in the side chain, which commonly serves as a pharmacophore. The Ramachandran maps (in the gas phase and in CHCl3 mimicking environments) of ΔPhe analogues with methyl groups at the β position of the side chain as well as at the C‐terminal amide were calculated using the B3LYP/6‐31 + G** method. Unexpectedly, β‐methylation alone results in an increase of conformational freedom of the affected ΔPhe residue. However, further modification by introducing an additional methyl group at C‐terminal methyl amide results in a steric crowding that fixes the torsion angle ψ of all conformers to the value 123°, regardless of the Z or E position of the phenyl ring. The number of conformers is reduced and the accessible conformational space of the residues is very limited. In particular, (Z)‐Δ(βMe)Phe with the tertiary C‐terminal amide can be classified as the amino acid derivative that has a single conformational state as it seems to adopt only the β conformation. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
The interaction of the uterine estrogen receptor (R) with three anti-estrogens, Dimethylstilbestrol1, Tamoxifen and Nafoxidine, have been studied either indirectly by competitive experiments or directly using radioactive compounds. The affinity of these anti-estrogens for R was found much higher when determined directly without estradiol (E2) than when evaluated by competitive experiments. In the presence of E2, but not in its absence, the inhibitory activity of the anti-estrogens decreased slowly with time. The present report strongly suggests that R is transformed by E2 into a form less sensitive to anti-estrogen. This “desensitization” of R to the estrogen antagonists is proposed as another in vitro test for the E2 induced “activation” of its receptor.  相似文献   

10.
We reported recently (Yoshikawa, K. and Terada, H. (1982) J. Am. Chem. Soc. 104, 7644–7646) that the potent uncoupler of oxidative phosphorylation SF-6847 ((3,5-di-tert-butyl-4-hydroxybenzylidene)malononitrile) shows unique intramolecular restricted rotation of the malononitrile moiety. In this study, values for the activation energy Ea of the restricted rotation of SF-6847 derivatives with the same alkyl chain R in both ortho positions of the phenolic hydroxyl group were determined from the temperature-dependent change in the1H-NMR signals of their aromatic protons. The Ea values of the neutral forms of these derivatives were found to be the same irrespective of R, but those of the anionic forms increased with increase in the alkyl chain length of R. It was found that the restricted rotation of the malononitrile moiety regulates its electron-withdrawing ability in such a way as to keep the acid dissociability of these derivatives similar, overcoming the effect of steric hindrance by R. The protonophoric activity of these derivatives, in a phospholipid bilayer membrane and their uncoupling activity in rat-liver mitochondria were both found to depend on Ea of their anionic forms. The stability of the uncoupler anions regulated by the restricted rotation of the malononitrile group in a nonpolar membrane environment was found to be important for exhibition of these activities. The hydrophobicity of the anionic forms of these derivatives was suggested also to be affected by the intramolecular rotation.  相似文献   

11.
12.
The effect of controlling the redox potential (Eh) on chalcopyrite bioleaching kinetics was studied as a new aspect of redox control during chalcopyrite bioleaching, and its mechanism was investigated by employing the “normalized” solution redox potential (Enormal) and the reaction kinetics model. Different Eh ranges were established by use of different acidophiles (Sulfobacillus acidophilus YTF1; Sulfobacillus sibiricus N1; Acidimicrobium ferrooxidans ICP; Acidiplasma sp. Fv-AP). Cu dissolution was very susceptible to real-time change in Eh during the reaction. It was found that efficiency of bioleaching of chalcopyrite can be effectively evaluated on the basis of Enormal, since it is normalized for real-time fluctuations of concentrations of major metal solutes during bioleaching. For steady Cu solubilization during bioleaching at a maximum rate, it was important to maintain a redox potential range of 0 ≤ Enormal ≤ 1 (?0.35 mV optimal) at the mineral surface by employing a “weak” ion-oxidizer. This led to a copper recovery of > 75%. At higher Enormal levels (Enormal > 1 by “strong” microbial Fe2+ oxidation), Cu solubilization was slowed by diffusion through the product film at the mineral surface (< 50% Cu recovery) caused by low reactivity of the chalcopyrite and by secondary passivation of the chalcopyrite surface, mainly by jarosite.  相似文献   

13.
 The interaction of the two chiral isomers of the new anticancer agent [Pt(ampyr)(cbdca)] (ampyr=aminomethylpyrrolidine, cbdca=cyclobutanedicarboxylate) with 5′-GMP and with short G-containing oligonucleotides has been studied using 1H and 31P NMR, UV-vis spectroscopy and molecular modelling. Each isomer loses the cbdca ligand upon binding to the DNA fragments. Two geometrical isomers of the DNA adducts are formed owing to the presence of the unsymmetric ampyr ligand. These isomers prove to be GG-N7,N7 chelates for d(GpG), d(pGpG) and d(CpGpG). A slight preference for the formation of one geometrical isomer is found in the case of DNA fragments having a phosphate moiety and/or a C base at the 5′-site of the GG sequence. H-bonding interactions from the NH2 moiety towards the 5′-phosphate group and/or the O atom of the C base clearly favour the formation of one geometrical isomer. The presence of these H-bonds, together with the bulky pyrrolidine ring, has resulted in the unique observation (by 1H NMR) of NH protons of coordinated amines that do not rapidly exchange in a 99.95% D2O solution. Temperature-dependence studies show an extremely slow stack ⇄ destack conformational change for the CGG adducts of the S isomer, which could be related to these stable H-bonds of the amine protons towards the oligonucleotide. For the R isomer this stack ⇄ destack conformational change is faster, probably owing to more steric hindrance of the pyrrolidine ring as deduced from the NOESY data, and as also suggested by molecular modelling. The observation of extremely slow rotation around the Pt-N7 bond for [Pt(R-ampyr)(GMP-N7)2] provides further evidence for increased steric hindrance of the R isomer compared to the S isomer. The rate of binding of the drug to G bases proved to be second order for both isomers; in fact the (toxic) S isomer is about two times more reactive than the (non-toxic) R isomer, as seen from k 2 values of 0.17±0.01 M–1 s–1 for [Pt(S-ampyr)(cbdca)] and 0.09±0.01 M–1 s–1 for [Pt(R-ampyr)(cbdca)]. No solvent-assisted pathway is involved in these reactions, since the complexes prove to be stable in solution for weeks and therefore only a direct attack of the G base on the Pt must be involved. Because hardly any intermediate species can be detected during the reaction, coordination of the second G base must occur much faster than the binding of the first G base. Since direct attack of the nucleobases takes place, steric interactions become extremely important and therefore are likely to determine the reactivity, activity and even the toxicity of such Pt complexes. Received: 12 January 1999 / Accepted: 17 June 1999  相似文献   

14.
Stem CO2 efflux (E s) has been estimated from a temperature-related equation, but sap flux often affects measurements of E s, which leads to misunderstanding real stem respiration. In order to observe the relationship between E s and stem temperature and to analyze the effect of sap velocity on E s, stem temperature, E s and sap flux were measured from a subtropical Schima superba plantation in South China on three trees for consecutive 3 days in July and October 2009. Stem temperature, E s and sap velocity were significantly higher in July than in October. Stem temperature could explain 17–41 and 54–75% variations of E s in July and October, respectively. A negative relationship between E s and stem temperature was found during 1800–2300 hours in July. The daytime E s was 9.2, 4.3 and 2.4% higher than the predicted for three trees in July, and this occurred only on Tree 1 in October. Sap velocity was positively correlated with E s for three trees in July, and the increase of E s with the increase of sap velocity was only observed on Tree 1 in October. These results demonstrated that the occurrence of sap flux could account for the increase of daytime E s, and the effect of sap velocity on E s varied with the seasons from the S. superba stem.  相似文献   

15.
Although chiral anionic compounds, notably a large number of organic acids, have been found to be readily separated into enantiomers on BSA-based columns, the structural requirements for an efficient enantiomer discrimination by the protein is still not very well known. Since it is often observed that very hydrophobic acids, like many of the antiinflammatory “profens,” can be resolved with large separation factors for the enantiomers, a systematic study of a series of racemic α-substituted alkanoic acids was made. The series of analytes was prepared from α-amino acids, RCH(NH2)CO2H (where R = C1-C6), by reaction with N-(chloroformyl)-carbazole. A rapid increase in the capacity ratios of both enantiomers was found with increasing length of R. The effect, however, was larger for the last eluted enantiomer, leading to a substantial increase in the separation factor; this being 7.3 for R = C6 in 20 mM phosphate buffer (pH 8.0) with 30% of acetonitrile. Further, the separation factor also increased with decreasing organic modifier content. Thus when the R = C6-analyte was run at a mobile phase concentration of 20% acetonitrile and a flow rate of 1.5 ml/min, the time difference between the two eluted enantiomers exceeded 20 hr. A reasonable interpretation of our results seems to be that enantioselectivity is promoted by increased hydrophobic interaction. Since the anionic charge of the analyte is also taking part in the retention mechanism, a tight binding of the analyte will result from simultaneous electrostatic and hydrophobic interaction. When the latter is increased, less conformational freedom will be left for the analyte and the steric configuration at the α-carbon atom will become more and more important. Steric hindrance by the α-substituent in the first eluted enantiomer will counteract the tight binding caused by the combined binding interactions and lead to a smaller increase in the capacity ratio.  相似文献   

16.
Abstrac A study is made of the nonlocal nature of the electron energy distribution function in the positive column of a glow discharge in a tube filled with pure oxygen. The distribution function and the axial (E z) and radial (E r) electric fields as functions of radius are measured using an array of mobile probes. The experimentally obtained spatial profiles of the distribution function are used to test the applicability of the two-term approximation to the distribution function of the electrons with a nonlocal energy spectrum. The distribution function in a specified electric field E=E z +E r (where E zE r) is calculated by solving the coordinate-dependent Boltzmann equation in the two-term approximation and by directly integrating the equations of electron motion using the Monte Carlo method. A comparison between the experimental data and the results of simulations carried out for a broad parameter range shows that, in the case of a highly nonlocal electron energy spectrum, the two-term approximation makes it possible to calculate the electron distribution function with a fairly good accuracy, in which case, however, in imposing the boundary conditions, the electron losses at the plasma surface should be treated in the kinetic approximation. It is shown that using the reflection coefficient of the plasma surface for electrons instead of the loss cone in space makes it possible to accurately calculate the electron energy distribution function over the entire parameter range under consideration, including the transient region in which the electron-energy relaxation length is comparable to the characteristic plasma dimension. __________ Translated from Fizika Plazmy, Vol. 26, No. 11, 2000, pp. 1038–1045. Original Russian Text Copyright ? 2000 by Ivanov, Klopovskii, Lopaev, Rakhimov, Rakhimova.  相似文献   

17.
Quasi-elastic light scattering studies on some polyelectrolyte systems exhibit a somewhat “bizarre” behavior in the profile of the apparent diffusion coefficient Dapp as a function of the salt concentration Cs. As Cs is decreased, Dapp first increases in accordance with polyelectrolyte theories, and then undergoes a precipitous drop in value by over an order of magnitude at a well-defined critical value Cs = C. This “transition” from Cs > C (ordinary) to Cs < C (extraordinary) is referred to as the “ordinary-extraordinary” (o-e) transition. Ghosh, Peitzsch, and Reed [(1992) Biopolymers, Vol. 32, pp. 1105–1122] proposed a “filterable aggregate” (FA) and “other particle” interpretation for the o-e transition and its reversibility in regard to ionic strength changes. The present communication examines in detail the FA model as applied to the o-e transition. It is shown that the FA model fails to account of the established characteristics of the o-e transition. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
Abstract: Extracts of the nematode Caenorhabditis elegans contain five molecular forms of acetylcholinesterase (AChE) activity that can be separated by a combination of selective solubilization, velocity sedimentation, and ion-exchange chromatography. These are called form IA (5.2s), form IB (4.9.s), form II (6.7s), form III (11.3s), and form IV (13.0s). All except form III are present in significant amounts in rapidly prepared extracts and are probably native; form III is probably derived autolytically from form IV. Most of forms IA and IB can be solubilized by repeated extractions without detergent, whereas forms II, III, and IV require detergent for effective solubilization and may therefore be membrane-bound. High salt concentrations are not required for, and do not aid in, the solubilization of these forms. For all forms, molecular weights and frictional ratios have been estimated by a combination of gel permeation chromatography and velocity sedimentations in both H2O and D2O. The molecular weight estimates range from 83,000 to 357,000 and only form II shows extensive asymmetry. The separated forms have been characterized with respect to substrate affinity, substrate specificity, inhibitor sensitivity, thermal inactivation, and detergent sensitivity. Judging by these properties, C. elegans is like other invertebrates in that none of its cholinesterase forms resembles either the “true” or the “pseudo” cholinesterase of vertebrates. However, internal comparison of the C. elegans forms clearly distinguishes forms IA, III, and IV as a group from forms IB and II; the former are therefore designated “class A” forms, the latter “class B” forms. Genetic evidence indicates that separate genes control class A and class B forms, and that these two classes overlap functionally. Several factors, including kinetic properties, molecular asymmetry, molecular size, and solubility, all suggest that a molecular model of the multiple cholinesterase forms observed in vertebrate electric organs probably does not apply in C. elegans. Potential functional roles and subunit structures of the multiple AChE forms within each C. elegans class are discussed.  相似文献   

19.
A. Vilenkin 《Biopolymers》1977,16(8):1657-1675
A new technique is presented for treating the ground state of an heteropolymer with a random sequence of components. An exact system of equations is found for determining the ground state energy E which is equal to the polymer free energy f in the lowest-order approximation in T/V (V/2 is the large “surface” energy arising at the boundaries between coiled and “helical” sections: V ? T, Uk; U1 and –U2 are the free energies of the components counted from the corresponding coiled state energies). These equations are essentially simplified at certain fixed values of the ratio U1/U2. For integer values of U2/U1 and U1/U2 a solution is obtained with an accuracy exp(–V/Uk). The ground-state energy as a function of U1 and U2 is shown to be highly irregular: its derivatives have jumps at an infinite number of points. These jumps provide a fine structure of the melting curves. A smoothed over the jumps function E′ is found by way of analytic continuation from the integer values of U1/U2 and U2/U1. The accuracy of the approximation fE is estimated and the correctional term of order T/V is determined.  相似文献   

20.
Flecainide, an antiarrythmic agent, and its analogs were resolved on a high performance liquid chromatographic chiral stationary phase (CSP) based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid with the use of a mobile phase consisting of methanol‐acetonitrile‐trifluoroacetic acid‐triethylamine (80/20/0.1/0.3, v/v/v/v). The chiral resolution was quite successful, the separation factors (α) and the resolutions (RS) for 20 analytes including flecainide being in the range of 1.19–1.82 and 1.73–6.80, respectively. The ortho‐substituent of the benzoyl group of analytes was found to cause decrease in the retention times of analytes probably because of the conformational deformation of analytes originated from the steric hindrance exerted by the ortho‐substituent. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  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