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.
周正虎  王传宽 《生态学报》2017,37(7):2428-2436
土地利用方式的变化导致土壤碳氮含量及其化学计量关系的变化,然而土壤微生物化学计量及其驱动的碳氮矿化过程如何响应这种变化仍不明确。以帽儿山地区天然落叶阔叶林、人工红松林、草地和农田4种不同土地利用类型为对象,测定其土壤有机碳(C_(soil))、全氮(N_(soil))、微生物生物量碳和氮(C_(mic)和N_(mic))、土壤碳和氮矿化速率(C_(min)和N_(min)),旨在比较不同土地利用方式对土壤、微生物碳氮化学计量特征及矿化速率的影响,探索土壤-微生物-矿化之间碳氮化学计量特征的相关性,揭示微生物对土壤碳氮化学计量变化的响应和调控机制。结果显示:C_(soil)、N_(soil)、C_(mic)、N_(mic)和C_(min)均呈现天然落叶阔叶林人工红松林草地农田,而天然落叶阔叶林和草地的N_(min)显著高于人工红松林和农田。土地利用方式显著影响土壤和微生物碳氮比(C∶N_(soil)和C∶N_(mic)),均呈现农田最高。不同土地利用方式的数据综合分析发现:碳氮矿化速率比与C∶N_(mic)呈负相关,而和微生物与土壤碳氮化学计量不平衡性(C∶N_(imb))显著正相关。单位微生物生物量的碳矿化速率(qCO_2)随着C∶N_(mic)的增加而降低,而单位微生物生物量的氮矿化速率(qAN)随着C∶N_(mic)的增加而增加。C∶N_(imb)与qCO_2正相关,与qAN负相关。以上结果表明,微生物会通过改变自身碳氮化学计量、调整碳氮之间相对矿化速率,以适应土地利用变化导致的土壤碳氮及其化学计量的变异性,以满足自身生长和代谢的碳氮需求平衡。  相似文献   

2.
The protein BBL undergoes structural transitions and acid denaturation between pH 1.2 and 8.0. Using NMR spectroscopy, we measured the pKa values of all the carboxylic residues in this pH range. We employed 13C direct-detection two-dimensional IPAP (in-phase antiphase) CACO NMR spectroscopy to monitor the ionization state of different carboxylic groups and demonstrated its advantages over other NMR techniques in measuring pKa values of carboxylic residues. The two residues Glu161 and Asp162 had significantly lowered pKa values, showing that these residues are involved in a network of stabilizing electrostatic interactions, as is His166. The other carboxylates had unperturbed values. The pH dependence of the free energy of denaturation was described quantitatively by the ionizations of those three residues of perturbed pKa, and, using thermodynamic cycles, we could calculate their pKas in the native and denatured states as well as the equilibrium constants for denaturation of the different protonation states. We also measured 13Cα chemical shifts of individual residues as a function of pH. These shifts sense structural transitions rather than ionizations, and they titrated with pH consistent with the change in equilibrium constant for denaturation. Kinetic measurements of the folding of BBL E161Q indicated that, at pH 7, the stabilizing interactions with Glu161 are formed mainly in the transition state. We also found that local interactions still exist in the acid-denatured state of BBL, which attenuate somewhat the flexibility of the acid-denatured state.  相似文献   

3.
We introduce a pair of compartment models for the honey bee nest-site selection process that lend themselves to analytic methods. The first model represents a swarm of bees deciding whether a site is viable, and the second characterizes its ability to select between two viable sites. We find that the one-site assessment process has two equilibrium states: a disinterested equilibrium (DE) in which the bees show no interest in the site and an interested equilibrium (IE) in which bees show interest. In analogy with epidemic models, we define basic and absolute recruitment numbers (R0 and B0) as measures of the swarm's sensitivity to dancing by a single bee. If R0 is less than one then the DE is locally stable, and if B0 is less than one then it is globally stable. If R0 is greater than one then the DE is unstable and the IE is stable under realistic conditions. In addition, there exists a critical site quality threshold Q* above which the site can attract some interest (at equilibrium) and below which it cannot. We also find the existence of a second critical site quality threshold Q** above which the site can attract a quorum (at equilibrium) and below which it cannot. The two-site discrimination process, in which we examine a swarm's ability to simultaneously consider two sites differing in both site quality and discovery time, has a stable DE if and only if both sites’ individual basic recruitment numbers are less than one. Numerical experiments are performed to study the influences of site quality on quorum time and the outcome of competition between a lower quality site discovered first and a higher quality site discovered second.  相似文献   

4.
Analysis based on the integration of differential inequalities is employed to derive upper and lower bounds on the total populationN(t) = ∫ R θ(x 1,x 2,t) dx 1 dx 2 of a biological species with an area-density distribution function θ=θ(x 1,x 2,t) (≥0) governed by a reaction-diffusion equation of the form ∂θ/∂t =D2θ +fθ −gθ n+1 whereD (>0),n (>0),f andg are constant parameters, θ=0 at all points on the boundary ∂R of an (arbitrary) two-dimensional regionR, and the initial distribution (θ(x 1,x 2, 0) is such thatN(0) is finite. Forg≥0 withR the entire two-dimensional Euclidean space, a lower bound onN(t) is obtained, showing in particular thatN(∞) is bounded below by a finite positive quantity forf≥0 andn>1. An upper bound onN(t) is obtained for arbitrary bounded or unbounded)R withn=1,f andg negative, and ∫ R θ(x 1,x 2, 0)2 dx 1 dx 2 sufficiently small in magnitude, implying that the population goes to extinction with increasing values of the time,N(∞)=0. Forg≥0 andR of finite area, the analysis yields upper bounds onN(t), predicting eventual extinction of the population if eitherf≤0 or if the area ofR is less than a certain grouping of the parameters in cases for whichf is positive. These results are directly applicable to biological species with distributions satisfying the Fisher equation in two spatial dimensions and to species governed by certain specialized population models.  相似文献   

5.
Methods are described for extracting the rate constants, governing the transitions between states, for a generalized, linear three discrete state model for the Na conductance,g Na. It is shown that the problem cannot be solved given only the time course ofg Na during some test potential step with fixed initial conditions. However, if the effects of any two durations of some conditioning potential step on the peakg Na during the test step are taken into account, and the physically required assumption that the steady state is an equilibrium state is made, then the rate constants for both conditioning and test potentials can be solved for simultaneously. These methods were applied to thev 5 model of Goldman (5) in which theg Na(t) was described in terms of a generalized linear second order variable. After extensive computations involving some 20,000 combinations of experimentally determined input values, it was not found possible to extract a set of rate constants that were all real and positive as is required. It was concluded that if the transitions between states are first order and there are no subunit interactions, then the Na gating unit displays more than three states.  相似文献   

6.
Pierre Bancal 《Annals of botany》2009,103(8):1315-1324

Background and Aims

Nitrogen (N) remobilization is the major source of N for grain filling in wheat, the other being N uptake after anthesis (Nup); however, variations in remobilization efficiency are not fully understood. It is hard to tell whether the source or the sink effects predominate, because N in the culm at anthesis (Nant) correlates strongly with both N remobilization (Nrem) and grain number (Gn), respectively the main source and the main sink.

Methods

A pot experiment was thus designed to assess the relative contributions of the source and sink to Nrem regulation. Using two cultivars of winter wheat (Triticum aestivum, ‘Apache’ and ‘Autan’), three pre-anthesis and two post-anthesis N fertilization levels were applied in order to vary the N sources, while ear trimming at anthesis reduced sink size.

Key Results

Unlike results observed at a scale of m2, the equation binding Nant to Nrem exhibited a negative intercept, challenging the concept of nitrogen remobilization efficiency. Before ear trimming, Gn fitted well to Nant, with a slope dependent on genotype. To obtain a sink variable that was less correlated with Nant, the difference δGn was calculated between actual grain number and that which could be predicted from culm N before trimming. A multiple regression then predicted Nrem (r2 = 0·95) from Nant, Nup and δGn, with fitting unbiased by fertilization treatment, trimming or genotype.

Conclusions

In untrimmed culms, δGn had a negligible effect, so that Nrem could be fitted to Nant and Nup only: grain N filling appeared to be determined by sources only (Nant and Nup), not by sink, and the reduction of Nrem by Nup was quantified. In these ‘normal’ cases, the regulation of Nrem should thus be located within the N sources themselves. In contrast, ear-trimming needs to be considered with caution as it introduced a sink limitation on Nrem; moreover one with an important genotype effect.Key words: Triticum aestivum, winter wheat, source/sink, grain filling, nitrogen uptake, grain number, nitrogen harvest index, nitrogen remobilization efficiency, genotype × environment  相似文献   

7.
Characterizing the denatured state ensemble is crucial to understanding protein stability and the mechanism of protein folding. The aim of this research was to see if fluorescence could be used to gain new information on the denatured state ensemble. Ribonuclease Sa (RNase Sa) contains no Trp residues. We made five variants of RNase Sa by adding Trp residues at locations where they are found in other members of the microbial ribonuclease family. To better understand the protein denatured state, we also studied the fluorescence properties of the following peptides: N-acetyl-Trp-amide (NATA), N-acetyl-Ala-Trp-Ala-amide (AWA), N-acetyl-Ala-Ala-Trp-Ala-Ala-amide (AAWAA), and the five pentapeptides with the same sequence as the Trp substitution sites in RNase Sa. The major conclusions are: 1), the wavelength of maximum fluorescence intensity, λmax, does not differ significantly for the peptides and the denatured proteins; 2), the fluorescence intensity at λmax, IF, differs significantly for the five Trp containing variants of RNase Sa; 3), the IF differences for the denatured proteins are mirrored in the peptides, showing that the short-range effects giving rise to the IF differences in the peptides are also present in the proteins; 4) the IF values for the denatured proteins are more than 30% greater than for the peptides, showing the presence of long-range effects in the proteins; 5), fluorescence quenching of Trp by acrylamide and iodide is more than 50% greater in the peptides than in the denatured proteins, showing that long-range effects limit the accessibility of the quenchers to the Trp side chains in the proteins; and 6), these results show that nonlocal effects in the denatured states of proteins influence Trp fluorescence and accessibility significantly.  相似文献   

8.
Various types of two-state models, classified by the type of direct receptorionophore coupling, were formulated based on the previously presented generalized two-state model of cooperativity (Kijima &; Kijima, 1978) and their dose-response relationships were examined. Hill coefficient at the mid-point of dose-response curve nHo the measure of the cooperativity of curves, is restricted for partial agonists in any two-state models because nHo is expressed by the product of two terms, one of which decreases when the other increases. In the independent gating unit model in which the channel opens only when the independent gating units are all in the activated state, the restriction of nHo is the most stringent: it never exceeds 2. In 2 ÷ 1·39 even for full agonist. It appears to be incompatible with most of the cooperative responses observed on chemically excitable membrane. In the basic model or one protomer-one channel model, nHo never exceeds 2·0 when 〈p, the maximum fraction of open-channel, is less than 23. In the cooperative gating unit model, nHo is the least restricted, which is less than 2·8 when 〈p ≤ 0·5, but if the number of gating units, N in a receptor is practically reasonably small (N ≤ 12), nHo ≤ 2·0 when 〈p ≤ 0·58. It is discussed whether or not several representative drug-receptive membranes can be accounted for by two-state models. Response of the insect sugar receptor is out of the above limitations of two-state models and can be accounted for by three-state model. The origin of cooperative interaction can be inferred by the shapes of dose-response curves. Cooperative dose-response curves of two dimensional lattices or oligomerc systems with large number of protomers weakly interacting by long range forces bend upward more markedly at lower region than the curves of strongly interacting oligomers, when curves with the same nHo are compared.  相似文献   

9.
Dynamic force spectroscopy has become indispensable for the exploration of the mechanical properties of proteins. In force-ramp experiments, performed by utilizing a time-dependent pulling force, the peak forces for unfolding transitions in a multimeric protein (D)N are used to map the free energy landscape for unfolding for a protein domain D. We show that theoretical modeling of unfolding transitions based on combining the observed first (f1), second (f2), …, Nth (fN) unfolding forces for a protein tandem of fixed length N, and pooling the force data for tandems of different length, n1 < n2 < … < N, leads to an inaccurate estimation of the distribution of unfolding forces for the protein D, ψD(f). This problem can be overcome by using Order statistics theory, which, in conjunction with analytically tractable models, can be used to resolve the molecular characteristics that determine the unfolding micromechanics. We present a simple method of estimation of the parent distribution, ψD(f), based on analyzing the force data for a tandem (D)n of arbitrary length n. Order statistics theory is exemplified through a detailed analysis and modeling of the unfolding forces obtained from pulling simulations of the monomer and oligomers of the all-β-sheet WW domain.  相似文献   

10.
The second harmonic generation (SHG) signal intensity sourced from skeletal muscle myosin II strongly depends on the polarization of the incident laser beam relative to the muscle fiber axis. This dependence is related to the second-order susceptibility χ(2), which can be described by a single component ratio γ under generally assumed symmetries. We precisely extracted γ from SHG polarization dependence curves with an extended focal field model. In murine myofibrillar preparations, we have found two distinct polarization dependencies: With the actomyosin system in the rigor state, γrig has a mean value of γrig = 0.52 (SD = 0.04, n = 55); in a relaxed state where myosin is not bound to actin, γrel has a mean value of γrel = 0.24 (SD = 0.07, n = 70). We observed a similar value in an activated state where the myosin power stroke was pharmacologically inhibited using N-benzyl-p-toluene sulfonamide. In summary, different actomyosin states can be visualized noninvasively with SHG microscopy. Specifically, SHG even allows us to distinguish different actin-bound states of myosin II using γ as a parameter.  相似文献   

11.
We consider simple lattice models for short peptide chains whose states can be exhaustively enumerated to find the lowest energy conformation. Using these exact results and numerical simulations, we compute the distributions for the mean time tN, required to find the global minimum energy state by simulated annealing (SA), as a function of N, the number of units in the chain. On the basis of scaling arguments, the time tN, to find the global minimum energy of longer chains, beyond the range covered by exhaustive enumeration, can be estimated. On the basis of the observed exponential increase in folding time of the standard SA algorithms, it is imperative that better algorithms be found for minimizing longer chains. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
The (β/α)8-barrel is one of the most common folds functioning as enzymes. The emergence of two (β/α)8-barrel enzymes involved in histidine biosynthesis, each of which has a twofold symmetric structure, has been proposed to be a consequence of tandem duplication and fusion of a (β/α)4-half-barrel. However, little evidence has been found for the existence of an ancestral half-barrel in the evolution of other (β/α)8-barrel proteins. In order to detect remnants of an ancestral half-barrel in the (β/α)8-barrel structure of Escherichia coli N-(5′-phosphoribosyl)anthranilate isomerase, we engineered three potential half-barrel units, (β/α)1-4, (β/α)3-6, and (β/α)5-8. Among these three arrangements, only (β/α)3-6 is stable; it exists in equilibrium between monomeric and dimeric forms. Thus, the central segment of N-(5′-phosphoribosyl)anthranilate isomerase from E. coli can serve as a half-barrel precursor. A tandem duplication of (β/α)3-6 yielded predominantly monomeric structures that were quite stable. This result exemplified that the structural characteristics of noncovalently assembled half-barrels could be improved by duplication and fusion. Moreover, our results may provide information regarding the local structural units that encompass interactions important for the early folding events of this ubiquitous protein conformation.  相似文献   

13.
There is much confusion and error in published treatments of data for multiple binding of ligands (e.g., substrates) by proteins (e.g., enzymes). There is a widespread impression that if the equilibrium binding, r, of ligand, A, by a protein with n sites can be fitted to an equation with two hyperbolic terms, i.e., r=nαkα(A)1+kα(A)+nβkβ(A)1+kβ(A) (nα+nβ=n) then kβ and kβ are always the intrinsic binding constants for two sets of sites. Such a conclusion is often incorrect. For example, in many cases, the protein is constituted of identical protomers with initially identical sites for binding ligands, and yet graphical representations of the binding data appear to behave as if the sites are partitioned between two classes. Although the use of a linear combination of hyperbolic terms to represent binding of ligands by macromolecules always yields empirical parameters kα, kβkλ, they cannot correspond to site binding constants when there are interactions between sites. In some circumstances their values may even be imaginary, complex numbers. On the other hand, stoichiometric binding constants can be assigned unambiguously without making any assumption regarding the nature of the interactions among binding sites. These principles are illustrated concretely by analyses of binding measurements for several different proteins containing two to six sites.  相似文献   

14.
Consider the family Gn of all n neuron networks whose dynamical behaviors are described by Caianiello's neuronic equations, and also its subfamily Rn of all reverberating networks each of whose neuronic equations have only periodic solution (states), i.e., without having any transient states. This paper is specifically concerned with characterizations of the subfamily Rn. First, we show that Rn is contained in a subfamily Sn of Gn consisting of all self-dual networks. We introduce Chow's matrix corresponding to each network of Sn, using Chow parameters and some algebraic operations {2α} applied to the coefficient matrix of the network in Sn, such as interchanges of coefficients between two neurons or changes of their signs. Then we give some necessary conditions on Chow's matrix under which any network in Sn belong to Rn, and a necessary and sufficient condition on the coefficient matrix. We also discuss relations between Chow's matrix and the maximum period of reverberations. In particular, it is shown that Chow's matrix of a network in Rn is symmetric if and only if the maximum period of reverberations is less than three. By virtue of these results, we propose two methods of construction of networks in Rn. The first method is an inductive construction. The second is based on the algebraic operations.  相似文献   

15.
The structural and functional organization of the spinach chloroplast photosystems (PS) I, IIα and IIβ was investigated. Sensitive absorbance difference spectrophotometry in the ultraviolet (?A320) and red (?A700) regions of the spectrum provided information on the relative concentration of PS II and PS I reaction centers. The kinetic analysis of PS II and PS I photochemistry under continuous weak excitation provided information on the number (N) of chlorophyll (Chl) molecules transferring excitation energy to PS IIα, PS IIβ and PS I. Spinach chloroplasts contained almost twice as many PS II reaction centers compared to PS I reaction centers. The number Nα of chlorophyll (Chl) molecules associated with PS IIα was 234, while Nβ = 100 and NPS I = 210. Thus, the functional photosynthetic unit size of PS II reaction centers was different from that of PS I reaction centers. The relative electron-transport capacity of PS II was significantly greater than that of PS I. Hence, under light-limiting green excitation when both Chl a and Chl b molecules are excited equally, the limiting factor in the overall electron-transfer reaction was the turnover of PS I. The Chl composition of PS I, PS IIα and PS IIβ was analyzed on the basis of a core Chl a reaction center complex component and a Chl ab-LHC component. There is a dissimilar Chl ab-LHC composition in the three photosystems with 77% of total Chl b associated with PS IIα only. The results indicate that PS IIα, located in the membrane of the grana partition region, is poised to receive excitation from a wider spectral window than PS IIβ and PS I.  相似文献   

16.
Several platinum(II) complexes of the general type [Pt(OND)X] have been prepared and characterized, the ligand (OND) representing the phenolate anion of the tridentate Schiff bases N-(2- diethylaminoethyl)-salicylaldimine (D = NEt2), N-(2- ethylaminoethyl)-salicylaldimine (D = NHEt) and N- (3-thia-n-pentyl)-salicylaldimine (D= SEt) and X= Cl, NO3. As shown by conductimetric studies the nitrato complexes [Pt(OND)NO3] dissociate completely in methanol according to:
Spectrophotometry (normal and stopped-flow) has been used to study the kinetics of solvent substitution according to
with a variety of neutral and anionic nucleophiles Y in methanol at 20 °C and constant ionic strength, I= 0.2 M (NaClO4). The substitution follows a one- term rate law, v = kobs[Pt(OND)(H2O)+] = kY[Y]- [Pt(OND)(H2O)+]. The kY data obtained for 13 (D = NEt2) and 7 (D = NHEt; SEt) different nucleophiles Y cannot be adequately correlated with their npt0 values according to the well-known relationship log kY = snpt0 + log ks. The deviations are strongest for large and bulky nucleophiles such as Y=Ph3P, Bu3P, Ph3As, I- and for D = NEt2, from which it is concluded that steric crowding hinders the formation of the 5-coordinate transition state. The rate reducing steric cis-effect observed is of the order kY(D = NEt2):kY(D = NHEt):kY(D = SEt) = 1:35:63 for small nucleophiles Y and as large as 1:192:2640 for Y = Ph3P. The introduction of substituents X in the salicylaldehyde ring in ortho (X3), meta (X4) and para position (X5) to the phenolic oxygen proves the existence of rather small electronic effects (X4, X5) and much stronger steric effects of bulky substituents X3, neighboring the donor oxygen.With the standard substrate trans-[Ptpy2Cl2] some new npt0 values were determined, namely for N, N′- dimethylthiourea (npt0 = 7.02), N, N′ -diphenylthlourea (npt0 = 7.19), N, N, N′, N′-tetramethylthiourea (npt0 = 6.05) and for the pseudo-halide dicyanoamide ion, N(CN)2- (npt0= 3.05). The npt0 value for the pseudo-halide tricyanomethanide, ion, C(CN)3-, was estimated to be 3.03.  相似文献   

17.
In order to investigate the relationship existing between the electronic structures of collagen and its biochemical functions in vivo, the semiempirical CNDO/2 SCF MO calculations were carried out on several model polymers of collagen, (Gly-Pro)n, (Gly-Hyp)n, (Ala-Pro)n, (Ala-Hyp)n, (Gly-Pro-Gly)n, (Gly-Hyp-Gly)n, (Gly-Pro-Pro)n and (Gly-Pro-Hyp)n. Geometries of the skeleton of these polymers were assumed to be the same as those of poly(l-proline) I (cis) and II (trans) and the calculations were performed only on infinite polymers in a single chain. The results show that the cis form is always more stable than the trans form for all the polymers treated. This energy difference between the cis and trans forms depends, for example, on the kind of amino acid residue, Gly or Ala, but this could not be seen in the Pro or Hyp residue. The flexibility or mobility of the collagen structure was explained using the energy difference between the cis and trans forms of the polymers, i.e. the cis-trans conversion of the collagen was discussed in connection with the energy difference. The reason why the collagen has the constitution of (Gly-Pro-Hyp)n is briefly discussed.  相似文献   

18.
Treatment of methyl 4,6-O-benzylidene-α-D-mannopyranoside with dichloromethylenedimethylammonium chloride gave methyl 4,6-O-benzylidene-3-chloro-3-deoxy-2-(N,N-dimethylcarbamoyl)-α-D-altropyranoside and methyl 4,6-O-benzy]idene-2-chloro-2-deoxy-3-(N,N-dimethylcarbamoyl)-α-D-glucopyranoside. Methyl 4,6-O-benzylidene-α-D-allopyranoside gave under analogous conditions the corresponding 2-chloro-3-(N,N-dimethylcarbamoyl)-α-D-altrose and 3-chloro-2-(N,N-dimethylcarbamoyl)-α-D-glucose derivatives. Methyl 5-O-benzyl-α,β-D-ribofuranoside and methyl 5-O-methyl-β-D-ribofuranoside gave only the corresponding methyl 3-chloro-2-(N,N-dimethylcarbamoyl)-α-D-xylofuranoside derivatives.  相似文献   

19.
For the Davydov Hamiltonian several special cases are known which can be solved analytically. Starting from these cases we show that the initial state for a simulation using Davydovs |D1> approximation has to be constructed from a given set of initial lattice displacements and momenta in form of a coherent state with its amplitudes independent of the lattices site, corresponding to Davydovs |D2> approximation. In the |D1> ansatz the coherent state amplitudes are site dependent. The site dependences evolve from this initial state exclusively via the equations of motion. Starting the |D1> simulation from an ansatz with site dependent coherent state amplitudes leads to an evolution which is different from the analytical solutions for the special cases. Further we show that simple construction of such initial states from the expressions for displacements and momenta as functions of the amplitudes leads to results which are inconsistent with the expressions for the lattice energy. The site-dependence of coherent state amplitudes can only evolve through the exciton-phonon interactions and cannot be introduced already in the initial state. Thus also in applications of the |D1> ansatz to polyacetylene always |D2> type initial states have to be used in contrast to our previous suggestion [W. Förner, J. Phys.: Condens. Matter 1994, 6, 9089-9151, on p. 9105]. Further we expand the known exact solutions in Taylor serieses in time and compare expectation values in different orders with the exact results. We find that for an approximation up to third order in time (for the wave function) norm and total energy, as well as displacements and momenta are reasonably correct for a time up to 0.12-0.14 ps, depending somewhat on the coupling strengh for the transportless case. For the oscillator system in the decoupled case the norm is correct up to 0.6-0.8 ps, while the expectation values of the number operators for different sites are reasonably correct up to roughly 0.6 ps, when calculated from the third order wave function. The most important result for the purpose to use such expansions for controlling the validity of ansatz states is, however, that the accuracy of S(t) and H(t) (constant in time, exact values known in all cases) is obviously a general indicator for the time region in which a given expansion yields reliable values also for the other, physically more interesting expectation values.  相似文献   

20.
A theoretical investigation of the ring-opening polymerization (ROP) mechanism of ε-caprolactone (CL) with tin(II) alkoxide, Sn(OR)2 initiators (R?=?n-C4H9, i-C4H9, t-C4H9, n-C6H13, n-C8H17) was studied. The density functional theory at B3LYP level was used to perform the modeled reactions. A coordination-insertion mechanism was found to occur via two transition states. Starting with a coordination of CL onto tin center led to a nucleophilic addition of the carbonyl group of CL, followed by the exchange of alkoxide ligand. The CL ring opening was completed through classical acyl-oxygen bond cleavage. The reaction barrier heights of ε-caprolactone with different initiators were calculated using potential energy profiles. The reaction of ε-caprolactone with Sn(OR)2 having R?=?n-C4H9 has the least value of barrier height compared to other reactions. The rate constants for each reaction were calculated using the transition state theory with TheRATE program. The rate constants are in good agreement with available experimental data.  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