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 156 毫秒
1.
The osmotic pressure equation for nonideal, associating systems of the type nA +mB ? AnBm, has been derived, by using the assumption yA nB m/yA nyB m = 1. This treatment can also be applied to related associations such as nA + mB ? AB + AB2 + A2B + …. From osmotic pressure experiments on the pure reactants it is possible to obtain the molecular weights (MA and MB) of the reactants and also the virial coefficients (BAA and BBB) of the reactants. The osmotic pressure of a nonreacting mixture of A and B can be calculated from these measurements. It can be used along with osmotic pressure measurements on equilibrium mixtures of A and B to obtain expressions containing the equilibrium constant (or constants) and the cross-virial coefficients (BAB and BBA). Several procedures are described for the evaluation of the equilibrium constant (or constants) and the BAB or BBA terms. It appears that this procedure is a general one which is applicable to associations of the type nA + mB ? AB + A2B + AB2 + …. By correcting for nonideal behavior, one should then be able to apply it to any method available for analyzing ideal associations of the types considered here. In addition it is possible, subject to certain restrictions, to analyze associations of the type 3A + B ? A2 + AB.  相似文献   

2.
A thermal perturbation curve (TPC) is defined to be the derivative of the fractional degree of saturation, f, with respect to temperature, considered as a function of the natural logarithm of free ligand concentration, y. The theoretical framework for the use of such curves in the thermodynamic analysis of ligand binding to macromolecules is presented. The thermal perturbation curve either provides or complements the information obtained from the derivative binding isotherm ?f/?y. For a single set of identical and independent sites the TPC is identical to the derivative binding isotherm. Analysis of such a curve directly yield ΔH0 and ΔG0 for the binding reaction. In actual experimental work, however, the TPC can only be approximated because of “self-buffering” effects relations between the parameter of the approximate curve and the thermodynamic quantities have been developed. This technique is applied to the proflavin-α-chymotryspin system to demonstrate its usefulness. The general features of thermal perturbation curves for cases of multiple sets of independent sites and cooperatively interacting sites have also been developed. The analysis of thermal perturbation curves in combination with other methods should provide a more powerful approach to the characterization of ligand-macromolecule interactions.  相似文献   

3.
The solution three-dimensional structure of the protonated [Leu7]-surfactin, an hepta-peptide extracted from Bacillus subtilis, has been determined from two-dimensional 1Hnmr performed in 2H6-dimethylsulfoxide and combined with molecular modeling. Experimental data included 9 coupling constants, 61 nuclear Overhauser effect derived distances, NH temperature coefficients, and 13C relaxation times. Two distance geometry (DISMAN) protocols converged toward models of the structure and the best of them were refined by restrained and unrestrained molecular dynamics (GROMOS). Two structures in accord with the set of experimental constraints are presented. Both are characterized by a “horse saddle” topology for ring atoms on which are attached the two polar Glu and Asp side chains showing an orientation clearly opposite to that of the C11–13 aliphatic chain. Amphipathic and surface properties of surfactin are certainly related to the existence of such minor polar and a major hydrophobic domains. The particular “claw” configuration of acidic residues observed in surfactin gives important clues for the understanding of its cation binding and transporting ability. © 1994 John Wiley & Sons, Inc.  相似文献   

4.
Experimental evidence for the appearance of synchronized bioelectric activity in neurons under applied extremely low frequency (ELF) magnetic fields is shown. We have studied the synchronizing process by recording the intracellular bioelectric activity from pairs of neurons randomly chosen from the brain ganglia of the snail Helix aspersa. The recordings were made in real time under exposure to sinusoidal low frequency (50 Hz) weak (B0=1–15 mT) magnetic fields. Synchronization was observed in 27% of the pairs tested. A linear dependence of the firing frequency f with the energy density of the applied magnetic field (i.e., fB02) was presented. The ability of low frequency sinusoidal weak magnetic fields to promote “magnetic synchronization” is exciting and opens new avenues for induced electromagnetic field bioeffects.  相似文献   

5.
The steroid hormone-receptor complex concentrations measured by “differential dissociation” techniques have to be corrected to obtain the true concentrations of receptor binding sites (Bs). For the calculation of Bs, the parameters kn (product of the equilibrium association constant and the concentration of binding sites of the “nonspecific” component) and f (fraction of the nonspecific binding measured in the experimental estimates of bound ligand by a given technique), previously proposed by Blondeau and Robel (J. P. Blondeau and P. Robel, 1975, Eur. J. Biochem.55, 375–384) are important. A new parameter of interest, ? [? = knf(kn + 1)], is discussed. The measurement of this parameter ? for three “differential dissociation” techniques allows the comparison of their efficiency and their reliability under various conditions for hormone receptor measurement in cytosol. Charcoal and hydroxylapatite methods are more efficient than the Sephadex G-25 filtration method. It is demonstrated that the “isotopic dilution” correction generally used for the estimation of the background of a given technique may be incorrect whatever the method of correction. A new method, the “double concentration measurement,” is developed. This method is simple, rapid, and precise. It requires two receptor binding measurements at two different saturating concentrations of ligand. This method allows the measurement of the estradiol receptor binding activity from calf uterine cytosol, with an error of less than 5% in samples containing the receptor either free or previously complexed with radioactive hormone, even in the presence of very high concentrations (≤0.5 μm) of radioactive steroid.  相似文献   

6.
Conformational preferences of modified nucleoside, N(4)-acetylcytidine, ac4C have been investigated using quantum chemical semi-empirical RM1 method. Automated geometry optimization using PM3 method along with ab initio methods HF SCF (6-31G**), and density functional theory (DFT; B3LYP/6-31G**) have also been made to compare the salient features. The most stable conformation of N(4)-acetyl group of ac4C prefers “proximal” orientation. This conformation is stabilized by intramolecular hydrogen bonding between O(7)···HC(5), O(2)···HC2′, and O4′···HC(6). The “proximal” conformation of N(4)-acetyl group has also been observed in another conformational study of anticodon loop of E. coli elongator tRNAMet. The solvent accessible surface area (SASA) calculations revealed the role of ac4C in anticodon loop. The explicit molecular dynamics simulation study also shows the “proximal” orientation of N(4)-acetyl group. The predicted “proximal” conformation would allow ac4C to interact with third base of codon AUG/AUA whereas the ‘distal’ orientation of N(4)-acetyl cytidine side-chain prevents such interactions. Single point energy calculation studies of various models of anticodon–codon bases revealed that the models ac4C(34)(Proximal):G3, and ac4C(34)(Proximal):A3 are energetically more stable as compared to models ac4C(34)(Distal):G3, and ac4C(34)(Distal):A3, respectively. MEPs calculations showed the unique potential tunnels between the hydrogen bond donor–acceptor atoms of ac4C(34)(Proximal):G3/A3 base pairs suggesting role of ac4C in recognition of third letter of codons AUG/AUA. The “distal” conformation of ac4C might prevent misreading of AUA codon. Hence, this study could be useful to understand the role of ac4C in the tertiary structure folding of tRNA as well as in the proper recognition of codons during protein biosynthesis process.  相似文献   

7.
Many herbicides of different chemical structure inhibit photosynthetic electron flow by interrupting the photosyn‐thetic electron flow by interrupting the photosynthetic electron transport chain between the primary acceptor (QA) and the secondary acceptor (QB) of photosystem 2 (PS2). Thermoluminescence (TL) originates from PS2, and the bands of the glow curve can be related to the charge recombination between positively charged donors and negatively charged acceptors. The glow curve of TL is strongly influenced by addition of PS2 herbicides. The herbicide treatment shifts the peak position and activation energy of the TL band related to QA, suggesting that herbicide binding affects the midpoint redox potential not only of Q B but also that of QA. On the basis of the band shift the herbicides of various chemical structures can be classified into different “thermodynamical” groups which relfect the differences in the binding properties of these herbicides. As a new approach TL seems to be a useful technique in studying the mechanism and site of action of herbicides that inhibit electron transport of PS2.  相似文献   

8.
Quasielastic light scattering methods were used to study calf thymus DNA in solutions of LiCl, NaCl, NH4Ac, and NH4Cl. Plots of the reciprocal relaxation time (1/τ) vs sin2(θ/2), where θ is the scattering angle, exhibit two linear regions, in accordance with theories for semiflexible polymers based on the t → 0 approximation. In these theories the slope of the linear region at low angles is associated with the translational diffusion coefficient (Dt), whereas the slope of the linear region at high angles is associated with the segmental diffusion coefficient (Ds = kT/?s). The midpoint of the “transition” between these two linear regions is associated with the mean displacement between segments (b). Data presented here indicate that the Rouse-Zimm parameters b and ?s are significantly different for DNA in 0.4M NH4Cl relative to the other salts at comparable ionic strengths. It is suggested that this difference reflects local solvent structure and that both b and Ds are sensitive to the local water structure.  相似文献   

9.
Y. Xu  J. Feng  Y. Hu 《Molecular simulation》2013,39(5):375-383
The microphase-separation behaviors of two types of star-diblock copolymers (A)4(B)4 and (AB)4 are investigated through the dissipative particle dynamics (DPD). The simulated phase diagrams show similar ordered mesostructures as those of corresponding linear and cyclic ones, such as lamellas (LAM), perforated lamellas (PL), hexagonal (HEX) cylinders and body-centered-cubic (BCC) spheres, besides, a series of gyroid (GR) morphologies predicted by theoretical research have been identified. In the regions between the totally ordered and disordered phases, we have found some melted morphologies that can be thought as locally ordered, such as micelles (M), liquid rods (LR) and random network (RN), which have not been identified in relevant theoretical predictions. The simulated threshold for a totally ordered mesostructure to appear is higher than theoretical predictions, which can be ascribed mainly to the increasing fluctuation with finite chain length, and the star architecture can facilitate microphase separation, which is in agreement with the theoretical predictions. In addition, it is easier for the (A) n (B) n copolymers than for corresponding (AB) n ones to trigger microphase separation under the same conditions. The relations between the root-mean-square radius of gyration (RMSGR) and the composition f A in the two types of star copolymers are almost contrary, which can be attributed to the differences in their structural characteristics.  相似文献   

10.
Total mercury concentrations (THg) in lake and stream sediments generally decrease with wet-area coverage (AW) per upslope basin area (AB). This was determined by delineating the wet-area component of 12,653 basins above as many sediment-sampling locations of the Geological Survey of Canada. These locations represent four climate regions (maritime, boreal, arctic, alpine) comprising six stream and six lake study areas. The dependence of sediment THg on AW/AB was examined by dividing the 0 < AW/AB < 1 range into 40 equal segments, and obtaining the mean sediment THg value for each segment. The results were evaluated by way of regression analysis using the following equation: mean sediment THg = a (1 ? AW/AB)b + c AW/AB, with a, b and c as area-specific coefficients. The “a” and “c” coefficients could – in part – be inferred from bedrock type, annual atmospheric Hg deposition, and mean monthly air temperatures, and mean annual precipitation. Both “a” and “c” increased with increasing atmospheric Hg deposition for lake sediments. For stream sediments, only “a” did so. The geogenic influence on the THg variations per study area was addressed through multiple regression analyses, using sediment concentrations of other heavy elements and organic matter as independent variables.  相似文献   

11.
In this paper, we use density functional theory to study the effect of the charge of solvophilic beads and concentration on the mesoscale structures of polyelectrolyte solution. The polyelectrolyte A6B12A6 was selected as the triblock polymer, and the solvophobic B blocks have no charges, while the solvophilic A blocks are charged. The simulation results showed: at higher concentration (above 50% systems), relatively small charges on the solvophilic block do not alter the bicontinuous phase inherent to uncharged solution, but at moderate concentrations (50% system), even though the charge per solvophilic bead is very small, the order lamellar structures become disturbed. Figure The density slice of A block in A6B12A6 solution at dimensionless time τ=5,000. 2D cut through the middle of the box shown in Fig. 2a for a (z orientation) and b (y orientation); in Fig. 2g for c (z orientation) and d (y orientation); and for 80% system with the charge of z A =0.5 for e (z orientation) and d (y orientation)  相似文献   

12.
A phenylthiophenyl-bearing Ru(II) complex of [Ru(bpy)2(Hbptip)](PF6)2 {bpy?=?2,2′-bipyridine, Hbptip?=?2-(4-phenylthiophen-2-yl)-1H-imidazo[4,5-f][1,10]phenanthroline} was synthesized and characterized by elemental analysis, 1H NMR spectroscopy, and electrospray ionization mass spectrometry. The ground- and excited-state acid–base properties of the complex were studied by UV–visible absorption and photoluminescence spectrophotometric pH titrations and the negative logarithm values of the ground-state acid ionization constants were derived to be pK a1?=?1.31?±?0.09 and pK a2?=?5.71?±?0.11 with the pK a2 associated deprotonation/protonation process occurring over 3 pK a units more acidic than thiophenyl-free parent complex of [Ru(bpy)2(Hpip)]2+ {Hpip?=?2-phenyl-1H-imidazo[4,5-f][1,10]phenanthroline}. The calf thymus DNA-binding properties of [Ru(bpy)2(Hbptip)]2+ in Tris–HCl buffer (pH 7.1 and 50?mM NaCl) were investigated by DNA viscosities and density functional theoretical calculations as well as UV–visible and emission spectroscopy techniques of UV–visible and luminescence titrations, steady-state emission quenching by [Fe(CN)6]4?, DNA competitive binding with ethidium bromide, DNA melting experiments, and reverse salt effects. The complex was evidenced to bind to the DNA intercalatively with binding affinity being greater than those for previously reported analogs of [Ru(bpy)2(Hip)]2+, [Ru(bpy)2(Htip)]2+, and [Ru(bpy)2(Haptip)]2+ {Hip?=?1H-imidazo[4,5-f][1,10]phenanthroline, Htip?=?2-thiophenimidazo[4,5-f][1,10]phenanthroline, Haptip?=?2-(5-phenylthiophen-2-yl)-1H-imidazo[4,5-f][1,10]phenanthroline}.  相似文献   

13.
Abstract

Systems containing a base or a base pair and 25 water molecules, as well as a helical stack and 30 water molecules per base pair, have been simulated. Changes in the base hydration shell structure, after the bases have been included into the pair and then into the base pair stack are discussed. Hydration shells of several configurations of the base pair stacks are discussed. Probabilities of formation of the hydrogen-bonded bridges of 1, 2 and 3 water molecules between hydrophilic centres have been estimated. The hydration shell structure was shown to depend on the nature of the base pair and on the stack configuration, while dependence of the global hydration shell characteristics on the stack configuration has been proved to be rather slight. The most typical structural elements of hydration shells, in the glycosidic (minor in B-like conformation) and non-glycosidic (major) grooves, for different configurations of AU and GC stacks, have been found and discussed. The number of hydrogen bonds between water molecules and bases per water molecule was shown to change upon transformation of the stack from A to B configuration. This result is discussed in connection with the reasons for B to A conformational transition and the concept of “water economy”. Hydration shell patterns of NH2-groups of AU and GC helical stacks differ significantly.  相似文献   

14.
Definitions are proposed for the independent and joint contributions that the chemical groups A and B make to the free energy of association of the ligand A?B with a receptor. The definitions are independent of the choice of the standard state and are consistent with the basic thermodynamic cycle relating the association of the ligands A?B, A?Y and X?B to the receptor Rappaport 1976. The basic idea is the use of the excess free energy of association of the ligand A?Y over the free energy of association of the reference ligand X?Y as the measure of the “independent” contribution of the group A to the binding. This definition allows the free energy of association of the ligand A?B to be written as the sum of the independent contributions of the groups A and B, their joint contribution, and an invariant free energy of association of the reference ligand with any receptor. With the appropriate definition of the receptor-reference ligand complex, water can be chosen as the reference ligand. Using ΔG(A?OH)?AG(HOH), ΔG(H?B?H)?ΔG(HOH) and ΔG(HO?C)?ΔG(HOH) as the definitions of the “independent” contributions of the chemical groups A, B and C to the binding of the ligand A?B?C, the joint contribution of the groups A and C to the binding is ΔG(A?B?C) ? ΔG(A?B?H) ? ΔG(H-B-C) + ΔG(H?B?H).  相似文献   

15.
Dairy propionibacteria are essential starters for Emmental cheese manufacture. The behavior of three commercial strains of Propionibacterium freudenreichii subsp. shermanii (P.f. 1, P.f. 2 and P.f 3) were studied in a liquid medium under air and N2 atmosphere using an on‐line pressure measurement technique. Growth kinetics and metabolite production were characterized under conditions usually reported as “optimal conditions” (pH 6.5, NaCl 0 %, temperature 30 °C) and also evaluated under “stressful conditions” (pH 5.2, NaCl 2 %, temperature 20 °C) simulating the cheese ripening conditions. In both cases, the effects of oxygen on growth were strain‐dependent. Under “stressful conditions”, two of the three strains were inhibited by oxygen under conditions of air atmosphere, while all three strains grew under conditions of N2 atmosphere. In the latter case, the duration of the lag phase and the maximum rate of pressure variation were significantly different, however, no significant differences were found between the strains with regard to the total fermentation time. Under “optimal conditions” metabolite production was strain‐dependent. In an air atmosphere, all strains produced more acetate and CO2 and less propionate than in a nitrogen atmosphere.  相似文献   

16.
Consider k independent exponential populations with location parameters μ1,…, μk and a common scale parameter or standard deviation θ. Let μ(k) be the largest of the μ's and define a population to be good if its location parameter exceeds μ(k) –Δ1. A selection procedure is proposed to select a subset of the k populations which includes the good populations with probability at least P*, a pre-assigned value. Simultaneous confidence intervals, that can be derived with the proposed selection procedure, are discussed. Moreover, if populations with locations below μ(k) –δ2, (δ2 > δ1) are “bad”, a selection procedure is proposed and a sample size is determined so that the probability of omitting a “good” population or selecting a “bad” population is at most 1 – P*.  相似文献   

17.
Estimation of coefficient of coancestry using molecular markers in maize   总被引:6,自引:0,他引:6  
Summary The coefficient of coancestry (fAB) between individuals A and B is the classical measure of genetic relationship. fAB is determined from pedigree records and is the probability that random alleles at the same locus in A and B are copies of the same ancestral allele or identical by descent (ibd). Recently, the proportion of molecular marker variants shared between A and B (SAB) has been used to measure genetic relationship. But SAB is an upwardly-biased estimator of fAB, especially between distantly-related lines. fAB, SAB, and adjusted (to remove bias) estimates of molecular marker similarity (f AB M ) were compared. RFLP banding patterns at 46 probe-restriction enzyme combinations were obtained for 23 maize inbred lines derived from the Iowa Stiff Stalk Synthetic (BSSS) maize (Zea mays L.) population, and for 4 non-BSSS lines. f AB M was estimated as , where A (or B) was the average proportion of RFLP variants shared between inbred A (or inbred B) and the non-BSSS lines. The average fAB among 253 pairwise combinations of BSSS lines was 0.212, whereas the average SAB was 0.397. The average f AB M was 0.162, indicating that the upward bias in SAB was effectively removed. SAB and fAB were significantly different ( = 0.05) in 76.3% of the comparisons, whereas 24.9% of the f AB M values differed significantly from fAB. The latter result suggests that selection and/or drift were present during inbred line development and that fAB may not be an accurate measure of the true proportion of ibd alleles between two lines. Cluster analyses based on S AB M and f AB M grouped lines according to pedigree, although several exceptions were noted. The presence of shared molecular marker variants between unrelated lines must be considered when setting SAB-based minimum distances for varietal protection. Under simplified conditions, more than 250 molecular marker loci are necessary to obtain sufficiently precise estimates of coefficient of coancestry using molecular markers.A contribution from Limagrain Genetics, a Group Limagrain company  相似文献   

18.
Service-theoretic concepts and methods, widely used in other fields (e.g., telecommunication and operations research), are useful also in a biochemical setting because the treatment of biocatalysts (enzymes, receptors) as servers and their ligands as customers, based on the established formal methods of service or queuing theory, may lead to insights and results unobtainable by conventional, mass-action-law-based theories. In this article, we apply the service-theoretic approach to receptor-agonist systems and show how by changing the stochastic time pattern of “operationally relevant” point events (e.g., instants of agonist arrival, instants of postclimax agonist departure) a great variety of dose-response curves may be generated, even in very simple reaction schemes, which, according to mass action kinetics, invariably lead to hyperbolic r(A) curves (r and A stand for response and agonist concentration, respectively). The molecular timing inherent to a hyperbolic response system is not optimal: for instance, at the agonist concentration A50, half of the agonist molecules are rejected (“lost”) because of unfortunate timing of the arrival events. The fraction of lost arrivers can be diminished considerably if the arrivals are better timed: “sub-Poisson” arrivals improve the timing and, thus, convert hyperbolic r(A) curves into “lifted” nonhyperbolic ones. Conversely, “super-Poisson” arrivals make the nonoptimal timing in hyperbolic response systems even worse and, thus, convert hyperbolic r(A) curves into “depressed” nonhyperbolic ones. Furthermore, under special timing conditions, nonhyperbolic r(A) curves can be generated, which are partly lifted, partly depressed relative to the reference hyperbola, and which resemble in shape well-known nonhyperbolic forms of enzyme and receptor kinetics (negatively cooperative, positively cooperative, and sigmoidal kinetics). In addition unusual (undulatory and sawtooth-like) r(A) curves can be generated solely by changing the temporal pattern of arrival and service completion instants. Virtually any shape of dose-response curves may be obtained by allowing for probability distributions whose characteristic shape varies with their mean; we call such distributions “variomorphic” and apply them to the arrival process of agonist molecules.  相似文献   

19.
Wayne L. Mattice 《Biopolymers》1985,24(12):2231-2242
The intramolecular formation of multiple clusters of interacting helices has been characterized in a homopolymer. The configuration partition function permits the formation of clusters in which the number of interacting helices may be as large as the greatest integer in n/2, where n denotes the number of amino acid residues in the chain. The theoretical formulation has its origin in a recent [Mattice, W. L. & Scheraga, H. A. (1984) Biopolymers 23 , 1701–1724], tractable matrix expression for the configuration partition function for intramolecular antiparallel β-sheet formation. Reassignment of the expression for one of the n(n+3)/2 elements in the sparse statistical weight matrix, along with a simple change in notation, converts that treatment into a matrix formulation of the configuration partition function for a chain containing multiple clusters of interacting antiparallel helices. The five statistical weights used are δ, fl, w, and the Zimm-Bragg σ and s. Each tight bend that connects two interacting helices contributes a factor of δ, fl is used in the weight for larger loops between interacting helices, and w arises from helix–helix interaction. The influence of the helix–helix interaction is well illustrated by two helix–coil transitions in a chain with n = 156 and σ = 0.001. In the absence of helix–helix interaction, the transition occurs by the nucleation and subsequent elongation of a small number of helices. When helix–helix interaction is attractive, the transition can occur by a different mechanism. Formation of a single pair of interacting helices is followed by addition of new helices to the initial cluster. In the latter process, individual helices experience relatively little growth after they are formed.  相似文献   

20.
RNA methylase genes are common antibiotic resistance determinants for multiple drugs of the macrolide, lincosamide, and streptogramin B (MLSB) families. We used molecular methods to investigate the diversity, distribution, and abundance of MLSB methylases in waste lagoons and groundwater wells at two swine farms with a history of tylosin (a macrolide antibiotic structurally related to erythromycin) and tetracycline usage. Phylogenetic analysis guided primer design for quantification of MLSB resistance genes found in tylosin-producing Streptomyces (tlr(B), tlr(D)) and commensal/pathogenic bacteria (erm(A), erm(B), erm(C), erm(F), erm(G), erm(Q)). The near absence of tlr genes at these sites suggested a lack of native antibiotic-producing organisms. The gene combination erm(ABCF) was found in all lagoon samples analyzed. These four genes were also detected with high frequency in wells previously found to be contaminated by lagoon leakage. A weak correlation was found between the distribution of erm genes and previously reported patterns of tetracycline resistance determinants, suggesting that dissemination of these genes into the environment is not necessarily linked. Considerations of gene origins in history (i.e., phylogeny) and gene distributions in the landscape provide a useful “molecular ecology” framework for studying environmental spread of antibiotic resistance.  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