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.
Monte Carlo simulations of flexible two-dimensional model membranes embedded in three space dimensions are reported. We explain in detail the techniques how to simulate fluid open membranes and fluid closed membranes (vesicles). It is shown that polymerized open membranes are rough and flat. Accordingly, the two larger components of the inertia tensor are proportional to the number of monomers of the surface, λ3 ≈ λ2N, whereas the smallest λ1N 0.65. Polymerized vesicles are isotropic and their mean square radius of gyration is R 2 ~ λ k N. In contrast, fluid membranes and vesicles exhibit crumpled shapes with λ k N 0.8 for k = 1,2,3. A monomer on a fluid surface exhibits a time-dependent mean squared displacement of r 2 (t) ~ t 0.8.  相似文献   

2.
Actin is an abundant protein that constitutes a main component of the eukaryotic cytoskeleton. Its polymerization and depolymerization are regulated by a variety of actin-binding proteins. Their functions range from nucleation of actin polymerization to sequestering G-actin in 1∶1 complexes. The kinetics of forming these complexes, with rate constants varying at least three orders of magnitude, is critical to the distinct regulatory functions. Previously we have developed a transient-complex theory for computing protein association mechanisms and association rate constants. The transient complex refers to an intermediate in which the two associating proteins have near-native separation and relative orientation but have yet to form short-range specific interactions of the native complex. The association rate constant is predicted as k a = k a0 , where k a0 is the basal rate constant for reaching the transient complex by free diffusion, and the Boltzmann factor captures the bias of long-range electrostatic interactions. Here we applied the transient-complex theory to study the association kinetics of seven actin-binding proteins with G-actin. These proteins exhibit three classes of association mechanisms, due to their different molecular shapes and flexibility. The 1000-fold k a variations among them can mostly be attributed to disparate electrostatic contributions. The basal rate constants also showed variations, resulting from the different shapes and sizes of the interfaces formed by the seven actin-binding proteins with G-actin. This study demonstrates the various ways that actin-binding proteins use physical properties to tune their association mechanisms and rate constants to suit distinct regulatory functions.  相似文献   

3.
The kinetics of association and dissociation of Escherichia coli 30 S and 50 S ribosomal subunits appear to fit the simple scheme
over a wide range of Mg2+ and ribosome concentrations, for the preparations studied (which have a sharp [Mg2+]-dependence on the equilibrium degree of association, e.g. 10% to 90% for 1.5 mm to 3.5 mm). Both rate constants depend strongly upon magnesium ion concentration (k2 goes from 0.04 × 106 to 21 × 106m−1 s−1 as [Mg2+] goes from 1.5 mm to 8 mm; k1 goes from 150 to 2 s−1 in the interval 1.0 to 3.0 mm), but k1 may level off above 3.0 mm and k2 increases slowly at high [Mg2+]. (The highest rate may not be far from the diffusion-controlled limit.) The primary effect of Mg2+, as calculated from the rather large changes in binding as a function of [Mg2+], is to decrease the contribution of electrostatic repulsion to the free energy of activation; specific, or class-specific, interactions of di- and multivalent cations with unknown ribosomal substructures may modulate this effect.  相似文献   

4.
《Inorganica chimica acta》1986,121(2):175-183
Chloride anation of trans-Pt(CN)4ClOH2 has been studied with and without Pt(CN)42− present at 25.0°C by use of stopped-flow and conventional spectrophotometry and a 1.00 M perchlorate medium. The rate law in the absence of Pt(CN)42− is Rate=(p1 + p2 [H+] ) [Cl]2 [complex]/(1 + q [Cl]) with p1=(3.0 ± 0.1) × 10−5 M−2s−1, p2=(3.6 ± 0.1) × 10−5 M−3 s−1 and q=(0.62 ± 0.02) M−1. It is compatible with a chloride assistance via an intermediate of the type Cl-Cl-Pt(CN)4···OH22−, in which the reactivity of the aqua ligand is enhanced due to a partial reduction of the platinum. This mechanism of halide assistance is in principle the same as the modified reductive elimination oxidative addition (REOA) mechanism proposed by Poë, in which the intermediate is not split into free halogen, platinum(II) and water, and in which electron transfer not necessarily involves complete reduction to platinum(II). To avoid confusion with complete reductive eliminations, reactions without split of the intermediates are here termed halide-assisted reactions. The pH-dependence indicates acid catalysis via a protonated intermediate ClClPt(CN)4···OH3.The Pt(CN)42−accelerated path has the rate law Rate=
[Cl-] [Pt(CN)42−] [complex] where k=(39.9±0.5) M−2 s−1 and Ka=(4.0±0.2)10−2 M is the protolysis constant of trans-Pt(CN)4ClOH2−.Reaction between PtCl5OH2 and chloride is accelerated by Pt(CN)42− and gives PtCl62− as the reaction product. The rate law is Rate=k [Cl] [Pt(CN)42−] [PtCl5OH2] with k=(5.6 ± 0.2)10−3 M−2 s−1 at 35.0°C and for a 1.50 M perchlorate acid medium. The reaction takes place without central ion exchange. Alternative mechanisms with two consecutive central ion exchanges can be excluded. The role of Pt(CN)42− in this reaction is very similar to that of the assisting halide in the halide assisted anations. [p ]Reaction between trans-Pt(CN)4ClOH2 and PtCl42− gives Pt(CN)42− and PtCl5OH2 as products and has the rate law Rate=k[PtCl42−] [trans-Pt(CN)4ClOH2] with k=(3.32 ± 0.02) M−1 s−1 at 25 °C for a 1.00 M perchloric acid medium. The formation of an aqua complex as the primary reaction product and the rate independent of [Cl] shows that formation of a bridged intermediate of the type Pt(II)Cl4ClPt(IV)(CN)4OH23− is formed in the initial reaction step, not five-coordinated PtCl53−.  相似文献   

5.
To determine the most favorable conditions for the production of ethanol by Pachysolen tannophilus, this yeast was grown in batch cultures with various initial concentrations of two of the constituents of the culture medium: d-xylose (so), ranging from 1 g·l−1 to 200 g·l−1, and yeast extract (lo), ranging from 0 g·l−1 to 8 g·l−1. The most favorable conditions proved to be initial concentrations of So=25 g·l−1 and lo=4 g·l−1, which gave a maximum specific growth rate of 0.26 h−1, biomass productivity of 0.023 g·l−1·h−1, overall biomass yield of 0.094 g·g−1, specific xylose-uptake rate (qs) of 0.3 g·g−1·h−1 (for t=50 h), specific ethanol-production rate (qE) of 0.065 g·g−1·h−1 and overall ethanol yield of 0.34 g·g−1; qs values decreased after the exponential growth phase while qE remained practically constant.  相似文献   

6.
We present a mechanism for the selectivity of covalent/electrostatic binding of the Cr(III) ion to collagen, mediated by the kosmotropicity of the anions. Although a change in the long-range ordered structure of collagen is observed after covalent binding (Cr(III)-OOC) in the presence of SO42− at pH 4.5, the νsym(COO) band remains intense, suggesting a relatively lower propensity for the Cr(III) to bind covalently instead of electrostatically through Cr(H2O)63+. Replacing SO42− with Cl reduces the kosmotropic effect which further favors the electrostatic binding of Cr(III) to collagen. Our findings allow a greater understanding of mechanism-specific metal binding in the collagen molecule. We also report for the first time, surface-enhanced Raman spectroscopy to analyze binding mechanisms in collagen, suggesting a novel way to study chemical modifications in collagen-based biomaterials.  相似文献   

7.
Surflex-Dock was applied to study interactions between 30 thiourea analogs and neuraminidase (NA). The docking results showed that hydrogen bonding and electrostatic interactions were highly correlated with the activities of neuraminidase inhibitors (NIs), followed by hydrophobic and steric factors. Moreover, there was a strong correlation between the predicted binding affinity (total score) and experimental pIC50 (correlation coefficient r = 0.870; P < 0.0001). A three dimensional holographic vector of atomic interaction field (3D-HoVAIF) was employed to construct a QSAR model. The r 2, q 2 and r 2 test values of the optimal QSAR model were 0.849, 0.724 and 0.689, respectively. From the QSAR model, it could be seen that electrostatic, hydrophobic and steric interactions were closely related to inhibitory activity, which was consistent with the docking results. Based on the docking and QSAR results, five new compounds with high predicted activities were designed.  相似文献   

8.
Abstract

The noncovalent interactions of phytate (Phy6-) with biogenic amines were studied potentiometrically in aqueous solution, at t= 25°C. Several species are formed in the different H+-Phy6--amine (A) systems, which have the general formula Ap(Phy)Hq(12-q)-, with p ≤ 3 and 6 ≤ q ≤ 10. The stability of these species is strictly dependent on the charges involved in the formation equilibria. For the equilibrium pHiAi+ + Hj(Phy)(12-j)- = Ap(Phy)Hq(12q)-, (q = pi + j)we found the relationship logK= aζ (ζ is the charge product of reactants), where a= 0.35(0.03, valid for all the amines; this roughly indicates an average free energy contribution per bond -ΔG0 = 4.0 ± 0.2 kJ mol-1. A slightly more sophisticated equation is also proposed for predicting the stability of these species. Owing to the quite high (partially protonated) phytate charge, the stability of Ap(Phy)Hq(12-q)- species is quite high, making phytate a strong amine sequestering agent in a wide pH range.  相似文献   

9.
《Inorganica chimica acta》1988,150(1):81-100
The (NH3)5CoOC(NH2)23+ ion is consumed in water according to the rate law k(obs.) = k1 + k2[OH], where k1 = 4.0 × 10−5 s−1 and k2 = 14.2 M−1 s−1 (0–0.1 M [OH];μ = 1.1 M, NaClO4, 25 °C). A hitherto unrecognized intramolecular O- to N- linkage isomerization reaction has been detected. In strongly acid solution only aquation to (NH3)5CoOH23+ is observed, but in 0.1–1.0 M [OH], 7% of the directly formed products is the urea-N complex (NH3)5CoNHCONH22+ which has been isolated. In the neutral pH region a much greater proportion (25%) of the products is the urea-N species. These results are interpreted in terms of an urea-O to urea-N linkage isomerization reaction competing with hydrolysis for both spontaneous (k1) and base-catalyzed (k2) pathways; the rearrangement is not observed in strongly acidic solution (pH ⩽ 1) because the protonated N-bonded isomer (pKa ≈ 3) is unstable with respect to the O-bonded form. The appearance of the isomerization pathway as the pH is raised in the 0–6 region is commensurate with a rate increase which cannot be attributed to a contribution from the base catalysis term k2[OH]. It is argued that this observation establishes, for the spontaneous pathway, that hydrolysis and linkage isomerization are separate reaction pathways — there is no common intermediate. The product distribution and rate data lead to the complete rate law, k(obs.) = k1 + k2[OH] = (ks + kON) + (kOH + kON) [OH] for the reactions of the O-bonded isomers, where ks, kOH are the specific rates for hydrolysis, and kON, kON are the specific rates for O- to N-linkage isomerization, by spontaneous and base-catalyzed pathways respectively; kON = 1.3 × 10−5 s−1 and kON = 1.1 M−1 s−1 (μ = 1.0 M, NaClO4, 25 °C). The O- to N- linkage isomerization has been observed also for complexes of N-methylurea, N,N-dimethylurea and N-phenylurea, but not for the N,N′-dimethylurea species. There is an approximately statistical relationship among the data for −NH2 capture (versus H2O), while −NHR and −NR2 do not compete with water as nucleophiles for Co(III) in either the spontaneous or base-catalyzed hydrolysis processes. For each urea-O complex, O- to N-isomerization is a more significant parallel reaction in the spontaneous as opposed to the base-catalyzed pathway. This is interpreted as being indicative of more associative character in the spontaneous route to products, a conclusion supported by other evidence. Some activation parameter data have been recorded and the effect of the N-substitution on the rates of solvolysis (H2O, Me2SO) is discussed. The urea-N complexes have been isolated as their deprotonated forms, [(NH3)5CoNHCONRR′](ClO4)2·xH2O (R,R′ = H, CH3). They are kinetically inert in neutral to basic solution but in acid they protonate (H2O, pKa 2–3; μ = 1.0 M, 25 °C) and then isomerize rapidly back to their O-bonded forms. Some solvolysis accompanies this N- to O-rearrangement in H2O and Me2SO. Specific rates and activation parameters are reported. The kinetic data follow a rate law of the form kNO(obs.) = (k + kNO)[H+]/(Ka + [H+]) and the active species in the reaction is the protonated form; k, kNO are the specific rates for hydrolysis and isomerization, respectively. Proton NMR data establish that the site of protonation (in Me2SO) is the cobalt-bound nitrogen atom. For the unsubstituted urea species (NH3)5CoNH2CONH23+, diastereotopic exo-NH2 protons arising from restricted rotation about the CN bond are observed. The relevance to the mechanism of the linkage isomerization process is considered. 13C and 1H NMR and electronic absorption spectral data are presented, and distinctions between linkage isomers and the solution structures (electronic and conformational) are discussed. The urea-N/urea-O complex equilibrium is governed by the relation KNO(obs.) = KNO[H+]/[H+](Ka), where KNO is the equilibrium constant = [(NH35Co(urea-O)3+]/[(NH3)5Co(urea-N)3+]. Values for KNO(=kNO/kON = 260 and pKa ≈ 3 for the NH2CONH2 system are consistent with the stability of the N-isomer in feebly acidic to basic solution (e.g. pH 6, KNO(obs.) = 2.6 × 10−2) and instability in acid solution (e.g. pH 1, KNO(obs.) = 240). The equilibrium data for this and other urea complexes of (NH3)5Co(III) are contrasted with the result for the analogous Rh(III)NH2CONH2 system KNO ≈ 1).  相似文献   

10.
An H-2-specific monoclonal antibody (mAb Q-1) was obtained from B10.Q (H-2 q) mice injected with syngeneic Sendai virus-coated cells. The IgM monoclonal antibody recognizes the public determinant H-2.25 shared by H-2 k (K k) and H-2 r haplotypes and cross-reacts with H-2d, H-2s, H-2p, and H-2q cells, the latter being the haplotype of the challenged B-cell donor. The binding of mAb Q-1 to H-2d, H-2s, H-2q, and H-2p cells was lower than to H-2k and H-2r and of decreasing affinity but could be clearly distinguished from the negative reactions with H-2b and H-2f cells. MAb Q-1 distinguishes between Sendai virus-coated and uncoated lymphocytes only cells with low-affinity binding. On virus-coated or infected (H-2p, H-2q, H-2d, H-2s) cells lysis was stronger than on normal lymphocytes. We interpret the enhanced lysis of Sendai virus-positive cells by mAb Q-1 to be due to recognition of a modified exposure of public H-2 determinants induced by Sendai virus.On leave from The Institute of Immunology and Experimental Therapy, Wroclaw, Poland  相似文献   

11.
This study investigates Ga-doped n-type PbTe thermoelectric materials and the dynamic phase conversion process of the second phases via Cu2Se alloying. Introducing Cu2Se enhances its electrical transport properties while reducing its lattice thermal conductivity (κlat) via weak electron–phonon coupling. Cu2Te and CuGa(Te/Se)2 (tetragonal phase) nanocrystals precipitate during the alloying process, resulting in Te vacancies and interstitial Cu in the PbTe matrix. At room temperature, Te vacancies and interstitial Cu atoms serve as n-type dopants, increasing the carrier concentration and electrical conductivity from ≈1.18 × 1019 cm−3 and ≈1870 S cm−1 to ≈2.26 × 1019 cm−3 and ≈3029 S cm−1, respectively. With increasing temperature, the sample exhibits a dynamic change in Cu2Te content and the generation of a new phase of CuGa(Te/Se)2 (cubic phase), strengthening the phonon scattering and obtaining an ultralow κlat. Pb0.975Ga0.025Te-3%CuSe exhibits a maximum figure of merit of ≈1.63 at 823 K, making it promising for intermediate-temperature device applications.  相似文献   

12.
Macrobrachium lamarrei (H. Milne-Edwards, 1837) is a well-known freshwater prawn species of Bangladesh. The aim of the research is to explore various reproductive aspects (size at sexual maturity, reproductive period and fecundity) of M. lamarrei in the Ganges River, Bangladesh through October 2012 to September 2013. We also study the environmental parameters and their impact on reproduction of M. lamarrei. A total 391 (ovigerous = 141, non-ovigerous = 250) female specimens were collected using Drag net. The TL50 (the TL at which 50% of individuals become mature) was calculated by a logistic equation as 5.20 cm. Based on the availability of ovigerous females the spawning season was February-November with the peak June-July. Further, 50% and 90% ovigerous females were observed when Fulton’s condition factor (KF) was 0.85 and 1.03, respectively. The total fecundity (FT) was ranged from 65 to 370 where TL was 4.20–6.40 cm and BW was 0.84–2.50 g. Fecundity was found to be highly correlated with TL (r2 ≥ 0.96, rs = 0.96, p < 0.0001) and BW (r2 ≥ 0.88, rs = 0.93, p < 0.0001). Temperature (rp = 0.82, p = 0.009), dissolved oxygen (DO) (rp = −0.83, p = 0.0007), pH (rp = 0.80, p = 0.0014) and total alkalinity (rp = −0.87, p = 0.0002), were highly correlated with ovigerous females. The average temperature on peak spawning season was 32 °C. Also, the spawning period connected with the peak rainfall and showed a notable relation between rainfall and ovigerous females. In addition, exploration of long data series pointed that yearly average air temperature is rising by 0.029 °C yr−1, whereas yearly average rainfall is falling by 2.96 mm yr−1. Therefore, the result will be helpful for the sustainable management and conservation of M. lamarrei through fixed permissible mesh size and establishment of a ban period in the Ganges River, Bangladesh and adjoining ecosystems.  相似文献   

13.
Factors affecting the rates of plasmid transfer were investigated using Escherichia coli LC102 bearing a conjugative plasmid R100-1 and E. coli DH1. The rate constant of transconjugant increase, kti, was used for presenting the degree of plasmid transmissibility instead of the plasmid transfer efficiency (pte). The rate constant was defined as the specific rate of transconjugant increase (srti, the number of transconjugants per donor per h) divided by the recipient cell concentration. The kti values ranged between 10−10 and 10−15 ml cells−1 h−1, when estimated under various conditions. Moderate liquid agitation had a favorable effect on ktf but agitation rates higher than 33 s−1 (intergrated shear force) greatly decreased the value of kti. The transconjugant-forming activity of the cells growing in continuous culture did not significantly change with the dilution rate, except those growing at dilution rates less than 0.1 h−1. The rate constant kti at temperatures of 10–15°C was as low as the detection limit (10−15 ml cells−1 h−1).  相似文献   

14.
《BBA》1985,809(3):311-319
The association of the site of photosynthetic water oxidation with Cl and other activating anions was analyzed with Photosystem II-enriched thylakoid particles prepared by Triton X-100 treatment. On the basis of the experimental evidence it is proposed that, regardless of the presence of the extrinsic 18 and 23 kDa polypeptides, the reactivation of Cl-depleted particles by added anions is contingent upon the protonation of membrane-bound buffering groups with an apparent pKa of approx. 6. The rate of dissociation of the anion from the site of water oxidation followed the order NO3 . ClO4 . Br ≈ Cl. A model is developed that takes into account the known requirement of more than 1 anion/center for maximum activity. It assumes a finite and anion-dependent capacity of the water-oxidizing site for the activating anions, and that the rate of photosynthetic oxygen evolution is proportional to the number of anions at the water-oxidizing enzyme. In accord with suggestions made earlier, and in agreement with their further elaboration by Coleman and Govindjee (Proc. 16th FEBS Congress, Moscow, VNU Science Press, Utrecht), it is proposed that the Cl requirement of Photosystem II is linked to protonation-deprotonation events associated with photosynthetic water oxidation.  相似文献   

15.
The stereodynamics for H++HD and its isotopic variant D++HD were studied with a quasi-classical trajectory (QCT) method at a collision energy of 0.7 eV on the ground 11A′ potential energy surface (PES). The polarization-dependent differential cross-sections (PDDCSs) in the center-of-mass frame are presented here. Furthermore, the distribution of the angle between k and j′, p(θ r ) and the distribution of the dihedral angle p(ϕ r ) were calculated and are discussed. The results indicate that isotopic substitution exerts substantial effects on the differential cross-section and the product’s rotational polarization.  相似文献   

16.
The kinetics of malonate replacement in bis- (malonato)oxovanadate(IV), [VO(mal)2H2O]2−(hereafter water molecule will be omitted), by oxalate has been studied by the stopped-flow method. The reaction was found to consist of two consecutive steps (k1 and k2: first-order rate constants) passing through a mixed ligand complex, [VO(mal)(ox)]2−. The rates for each step depended linearly on the concentrations of free oxalate species, Hox and ox2−. The second-order rate constants for the replacement by ox2− were much larger in the k1 step than in the k2 step and the activation parameters were determined as follows: ΔH= 43.5 ± 5.6 kJ mol−1, ΔS±-53 ± 19 J K−1 mol−1 and ΔH≠= 43.6 ± 0.5 kJ mol−1, δS≠ = -62 ± 2 J K−l mol−1 for the k1 and k2 steps, respectively. The volume of activation was determined to be -0.65 ± 0.75 cm3 mol−1 at 20.2 °C by the high-pressure stopped-flow method for the apparent rate constants.  相似文献   

17.
In normalyears, eggs and prolarvae of the plaice (Pleuronectes platessa L.) in the southern North Sea develop within the temperature range 6.0–8.5 °C, although the water may at times be some degrees colder or warmer than this. The effect of temperature, t °C, on the embryonic development time, D days, has been investigated within the tolerated range 2.8–10.5 °C. Various models to express the observed curvilinear relationship between t and D have been considered, that giving the closest fit to the data being (tt0)(DD0) = k or D = k(t−t0)+D0. A method is given for the calculation of constants k, D0, and t0. The relationship may also be expressed by the equation D = a(tt0)b where a and b are constants, but t0 must in this case be found by iteration. At investigated temperatures in the range 4.1–10.5 °C the smallest eggs in a batch from a single source hatched first. Within the tolerated range, hatching prolarvae were substantially smaller at 10.5 °C than at the other temperatures. During the period of prolarval yolk utilization, growth is slower at the high temperatures, so that median temperatures of 6.5–8.0 °C are most efficient in terms of the relationship between growth in length and yolk utilization. Toward the end of the yolk-sac phase, the rate of yolk utilization declines unless a suitable external food source (e.g., Artemia nauplii) is provided.  相似文献   

18.
Salts inhibit the activity of sweet almond β-glucosidase. For cations (Cl salts) the effectiveness follows the series: Cu+2, Fe+2 > Zn+2 > Li+ > Ca+2 > Mg+2 > Cs+ > NH4+ > Rb+ > K+ > Na+ and for anions (Na+ salts) the series is: I > ClO4 > SCN > Br  NO3 > Cl  OAc > F  SO4 2. The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of kcat/Km for the β-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pKas and not to an effect on the pH-independent second-order rate constant, (kcat/Km)lim. For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pKa1 increases (from 3.7 to 4.9) and the apparent pKa2 decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (kcat/Km)lim (= 9 × 104 M 1 s 1) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pKas when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pKa shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pKa of 3.7 and dissociated from the group with a pKa of 7.2) is very small (≈ 0.03%). At higher salt concentrations, where the two pKas become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (kcat/Km)lim, of the monoprotonated species. For other enzymes which may show such salt-induced pKa shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

19.
The plant-to-plant movement of insects in one of the factors determining the distribution of individuals in insect populations. In this report the movement of barley aphids was analyzed by a statistical model. The model is represented as the convolution of three probability functions:
  1. The probability that s individuals are found on a plant at time t0:Q(s);
  2. The probability that i individuals leave the plant and remain on the ground from time t0 to t1:sCipiqs−i and p+q=1, where p and q are the proportions of individuals which do not leave a plant and which leave it once or more, respectively;
  3. The probability that j individuals climb a plant between time t0 to t1 and stay there at time t1:e−λλj!, where λ is the mean of the individuals.
The probability that l individuals are located on a plant at time t1 is represented by the following equation It was shown by simple experiments that the experimental populations were well fitted to the model.  相似文献   

20.
Changes in the rate and intensity of oxygen consumption during individual ontogeny of 14 specimens of Lymnaea stagnalis in the period from the 10th week after emergence until death was investigated in aquaculture. It was demonstrated that the rate of oxygen consumption increased and the intensity of this process decreased during the whole period of observations. Alterations of these parameters were accompanied with permanent oscillations of their meanings. The correlation between intensity of oxygen consumption (q) and age (t) can be described with the equation q = q st/(1-exp(−k g (t+t 0))). The values of coefficients of this equation do not differ significantly between individuals and, on average, comprise k g = 0.0696 ± 0.0072 weeks−1; q st = 60.4 ± 2.6 mcl O2/(h · g); t 0 = −3.0 ± 0.7 weeks. The dependence of the rate of oxygen consumption (Q, μl O2/h) on body weight (M, g) for all data is significantly described with the allometric equation Q = 0.369M 0.779.  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