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 640 毫秒
1.
Relaxation Young's and shear moduli of bovine bone and bone collagen were investigated. It was found that each relaxation process observed had two stages, which were referred to as process I and process II in order of time. Process II was described by a simple exponential decay while process I was not. The Kohlrausch-Williams-Watts (KWW) function, ψ(t) = exp[t1)B] (0 < B < 1), was found to be suitable to describe process I. The normalized relaxation modulus, Mr(t), was expressed by the combination of the simple exponential type relaxation function and the KWW function
Mr=A1exp[−(t1)B]+A2exp[(t1)](0<B1)
On the basis of this equation, the relaxation mechanism in bone and bone collagen was identified. According to the model proposed for the KWW relaxation function, the stress relaxation process in bone was considered to be governed by viscoelastic properties of matrix collagen fiber. The model for the KWW relaxation function requires the disordered glassy structure of collagen fiber, which is consistent with the results of the structural investigations.  相似文献   

2.
The process of physical degradation by means of the ultrasonic action towards chitosans with mole fraction of 2-acetamido-2-deoxy-β- -glucopyranose units (the degree of N-acetylation, FA) in the range of 0.10≤FA≤0.28, and the weight average polymerisation degree in the range of has been investigated. The decrease of as well as changes in the distribution of the degree of polymerisation (P) has been determined as a function of time, FA, temperature, concentration of chitosan solution and concentration of acetic acid in the solution. The use of low-power ultrasound emitter allowed to establish that in the case of chitosan (binary heteropolysaccharide) the general rate parameter (k) increased with FA. This can be explained by the relatively stronger aggregation of macromolecules with higher FA, which results in size increase of macromolecular individuals and hence in their higher susceptibility to ultrasonic action. It was also observed that k decreased with chitosan concentration and temperature. The value of limiting degree of polimerisation (xe) was found to be influenced by structural parameters of chitosan chains (FA, aggregation). The increase of acetic acid concentration caused the increase in the k value, what indicated accelerating effect of ultrasound towards acidic hydrolysis of chitosan. The shape of the P curve of sonochemically degraded chitosans are in good correlation with the mid-point breakage concept of degradation accepted in sonochemical degradation of polymers.  相似文献   

3.
These data describe improved modulation of discharge rates (rate coding) of first dorsal interosseous motor units throughout the acquisition of a complex force-matching skill involving isometric index finger abduction. In each of 15 consecutive trials, subjects attempted to match their force to a trajectory consisting of the sum of two sine waves (0.15 and 0.5 Hz) and random oscillations (overall mean force level ˜20% MVC). Reductions in root-mean-square (RMS) error of each subject’s force relative to the trajectory indicated substantial improvements in force-matching ability (F=33.8, p<0.001). With the acquisition of this new skill, there was increased amplitude modulation of muscular force near both dominant frequencies of the force-matching trajectory (F=10.6, p=0.008). The standard deviation and coefficient of variation of motor unit inter-spike intervals both decreased with improved performance indicating a general reduction in the amplitude of firing rate modulations (SD: F=18.69, p=0.001; CV: F=43.6, p<0.001). After skill acquisition, there was decreased firing rate modulation outside of the two dominant frequencies and increased amplitude of firing rate modulation at the higher of the two dominant frequencies (0.5 Hz, F=8.23, p=0.015). These findings indicate that improved precision of rate coding was a contributor to the acquisition of the new force-matching task. That the change in rate coding was frequency dependent suggests that factors other than frequency coding may contribute to the improved force matching at 0.15 Hz.  相似文献   

4.

1. 1.|In 15 conscious Pekin ducks, 40 “warm sensitive” hypothalamic neurons were identified according to their discharge rates at 40°C Thy (F40), local temperature coefficients (Δ/ΔT) and Q10.

2. 2.|Q10 and either F40 or ΔFT were little or not related.

3. 3.|A positive correlation between F40 and ΔFT was observed which was particularly close (r = 0.94 and 0.96) when the neurons were classified according to their Q10 of <2 and >2.

4. 4.|The results suggest that neurons with positive temperature coefficients in the duck's hypothalamus mostly exhibit linear to exponential temperature-discharge relationships.

5. 5.|This is an contrast to observations on mammalian hypothalamic thermosensitive neurons and may relate to the absence of the thermosensory function in the duck's rostral brainstem.

Author Keywords: Neuronal thermosensitivity; hypothalamic thermosensory function; Temperature and synaptic transmission; avian thermoregulation; mammalian thermoregulation  相似文献   


5.
This study uses chlorophyll a fluorescence to examine the effect of environmentally relevant (1–4 h) exposures of thermal stress (35–45 °C) on seagrass photosynthetic yield in seven tropical species of seagrasses. Acute response of each tropical seagrass species to thermal stress was characterised, and the capacity of each species to tolerate and recover from thermal stress was assessed. Two fundamental characteristics of heat stress were observed. The first effect was a decrease in photosynthetic yield (Fv / Fm) characterised by reductions in F and Fm′. The dramatic decline in Fv / Fm ratio, due to chronic inhibition of photosynthesis, indicates an intolerance of Halophila ovalis, Zostera capricorni and Syringodium isoetifolium to ecologically relevant exposures of thermal stress and structural alterations to the PhotoSystem II (PSII) reaction centres. The decline in Fm′ represents heat-induced photoinhibition related to closure of PSII reaction centres and chloroplast dysfunction. The key finding was that Cymodocea rotundata, Cymodocea serrulata, Halodule uninervis and Thalassia hemprichii were more tolerant to thermal stress than H. ovalis, Z. capricorni and S. isoetifolium. After 3 days of 4 h temperature treatments ranging from 25 to 40 °C, C. rotundata, C. serrulata and H. uninervis demonstrated a wide tolerance to temperature with no detrimental effect on Fv / Fm′ qN or qP responses. These three species are restricted to subtropical and tropical waters and their tolerance to seawater temperatures up to 40 °C is likely to be an adaptive response to high temperatures commonly occurring at low tides and peak solar irradiance. The results of temperature experiments suggest that the photosynthetic condition of all seagrass species tested are likely to suffer irreparable effects from short-term or episodic changes in seawater temperatures as high as 40–45 °C. Acute stress responses of seagrasses to elevated seawater temperatures are consistent with observed reductions in above-ground biomass during a recent El Niño event.  相似文献   

6.
Trans-dihydroxo-[tetrakis(2,6-dichlorophenyl)porphinato]ruthenium(IV) ([Ru(OH)2(TDCPP)]) was prepared by meta-chloroperbenzoic acid oxidation of [Ru(CO)(TDCPP)] in dichloromethane-toluene, and its crystal structure is reported. Crystal data for [Ru(OH)2(TDCPP)]·2toluene:C44H22N4O2Cl8Ru·2C7H8, orthorhombic, space group Pbca a = 13.149(1), B = 19.893(2), C = 21.093(2)Å, U = 55.17.3(2) Å3, Z = 4. The short axial Ru---O bond distance, 1.790(7) Å, is in the range expected for a double Ru(IV)-oxygen bond. Both hydroxo ligands are approximately located in the mean plane of two opposite dichlorophenyl groups. Full-matrix least-squares refinement of positional and thermal parameters, using 2368 unique reflections with F > 2.5 σ (F) led to R(F) = 0.063; Rw = 0.066.  相似文献   

7.
The degree of fluoresence polarization, P, of unoriented and magnetically oriented spinach chloroplasts as a function of excitation (400–680 nm) and emission wavelengths (675–750 nm) is reported. For unoriented chloroplasts P can be divided into two contributions, PIN and PAN. The latter arises from the optical anisotropy of the membranes which is due to the orientation with respect to the membrane plane of pigment molecules in vivo. The intrinsic polarization PIN, which reflects the energy transfer between different pigment molecules and their degree of mutual orientation, can be measured unambiguously only if (1) oriented membranes are used and the fluorescence is viewed along a direction normal to the membrane planes, and (2) the excitation is confined to the Qy (≈ 660−680 nm) absorption band of chlorophyll in vivo. With 670–680 nm excitation, values of P using unoriented chloroplasts can be as high as +14%, mostly reflecting the orientational anisotropy of the pigments. Using oriented chloroplasts, PIN is shown to be +5±1%. The excitation wavelength dependence studies of PIN indicate that the carotenoid and chlorophyll Qy transition moments tend to be partially oriented with respect to each other on a local level (within a given photosynthetic unit or its immediate neighbors).  相似文献   

8.
The viscosity in the low shear rate Newtonian domain of three biopolymers, locust bean gum, guar gum and xanthan gum was studied as a function of temperature and of polymer concentration in various aqueous solvents. The intrinsic viscosities [η]o of both galactomannans are not modified in the presence of 10 or 40% sucrose. In this case, a master curve relating the Newtonian specific viscosity (ηsp)o, to the reduced concentration c[η]o is obtained and allows (in good agreement with theoretical conjectures), two critical concentrations C* and C** to be defined, from which the value of the expansion coefficient may be estimated. For xanthan, as expected for a polyelectrolyte, [η]o depends strongly on salt concentration and on added sucrose and the results did not obey the above-mentioned master curve. However, it is shown that (ηsp)o depends only on xanthan concentration whenC > C**, and then it is assumed that chain dimensions have attained their unperturbed values whatever the solvent. Considering that both types of chains, random coils (galactomannans) and semi-rigid (xanthan) should give the same (ηsp)o-C[η]o master curve for C > C** when [η]o is replaced by its unperturbed counterpart [η]θ, a method for estimating [η]θ for the xanthan sample is proposed. In conclusion, the numerous exceptions to the widely accepted (ηsp)o vs C[η]o “universal” behaviour are mainly ascribed to significant differences in expansion coefficient values which depend on both the polymer and the solvent.  相似文献   

9.
B. Bouges-Bocquet 《BBA》1973,292(3):772-785

1. 1. By varying the redox potential of a chloroplast suspension, we obtained new evidence for an equilibrium between states S0 and S1 in the model of Kok, B., Forbush, B. and McGloin, N. (1970, Photochem. Photobiol. 11, 457–475). The mid-point potential of the S0 to S1 couple is close to that for the pool of the electron acceptor of System II, A to A.

2. 2. The limiting steps between two consecutive photoreactions of System II in Chlorella and spinach chloroplasts, have been studied.

2.1. (a) The limiting step from S1 to S2 (noted γ1t)) is not exponential. Its temperature coefficient becomes greater as the reaction proceeds. The shape of the kinetics is an intrinsic property of each center. Chloroplasts fixed with 2% glutaraldehyde, show simple first order kinetics.

2.2. (b) The limiting step from S0 to S10t)) exhibits the same characteristics as γ1t)).

2.3. (c) The limiting step from S2 to S32t)) shows sigmoidal kinetics; two reactions are involved. One of the reactions exhibits the same properties as γ0t) and γ1t).

2.4. (d) The limiting step from S3 to S03t)) is a first order reaction, two times slower than the other transitions. This reaction is interpretated in terms of oxygen release.

3. 3. We also studied the limiting steps in the presence of low concentrations (50 μM) of hydroxylamine. The results favor the binding of two molecules of hydroxylamine to every photochemical center.

Abbreviations: DCIP, dichlorophenolindophenol  相似文献   


10.
通过转基因烟草(Nicotiana tabacum)验证天山雪莲(Saussurea involucrata) Δ9硬脂酰-ACP脱饱和酶基因SiSAD与拟南芥(Arabidopsis thaliana)中同源基因AtFAB2的抗寒性功能。利用农杆菌介导法将植物表达载体PSiSAD:AtFAB2PSiSAD:SiSAD导入烟草, 然后将2种转基因和野生型烟草分别置于20°C、10°C、5°C、0°C及-2°C下处理2小时, 检测其相对电导率、丙二醛(MDA)含量、叶绿素荧光参数(Fv/Fm)及脂肪酸含量。将-2°C处理2小时后的植株置于25°C培养1周进行生长恢复实验。结果表明, 生长恢复实验中转SiSAD基因烟草的恢复效果显著优于转AtFAB2基因和野生型烟草。在0°C和-2°C处理2小时后, 转SiSADAtFAB2基因型和野生型烟草的相对电导率和丙二醛含量呈现显著递增趋势; 转SiSADAtFAB2基因型烟草的Fv/Fm显著高于野生型烟草, 其中, 转SiSAD基因烟草的Fv/Fm显著高于转AtFAB2基因烟草。转AtFAB2基因型和野生型烟草的油酸(C18:1)含量随着温度的降低逐渐升高后降低并在0°C时达到最高值; 而转SiSAD基因型烟草C18:1含量持续升高, 并在-2°C时达到最高值, 其含量分别是转AtFAB2基因型和野生型烟草的1.58倍和1.7倍。以上结果表明, 天山雪莲Δ9硬脂酰-ACP脱饱和酶基因SiSAD与拟南芥中同源基因AtFAB2均可以显著增强非低温驯化烟草的抗寒性, 但是SiSAD基因效果显著优于AtFAB2。  相似文献   

11.
0-group Carcinus maenas (L.) was investigated from June 1975 to September 1976 on a shallow sandy bottom at Kvarnbukten Bay, Gullmar Fjord (58° 15′N: 11°28′E), Sweden, at an average salinity of 25% and a range of monthly mean temperatures of −0.3 to 197. °C.

The new year-class settles from August to early September at a carapace breath of 2 to 3 mm and a calorific content of 32 cal. The distribution is restricted to clusters of the mussel Mytilus edulis L. Depth, type of substratum, and patches of the eel-grass Zostera marina L. are of no importance for their spatial distribution. There is no migration to deeper water in the autumn. The carapace breadth is ≈ 9.5 mm after one year of benthic life. Sexual maturity is reached after two years. Growth occurs at temperatures above 10 °C, i.e., from August to October and from May to July. During the first year of benthic life the animals moult 7 times. The 0-group seems to be micro-carnivores feeding on the sediment meiofauna.

The individual energy budget for the first year of benthic life is: consumption (Cc) 905 cal., production (P1c) 236 cal., cast carapaces (P2c) 153 cal., respiration (Rc) 404 cal., and rejectiction (Fc) 112 cal. The assimilation efficiency (Uc−1) is 88%, the gross growth efficiency (K1c) 43%, and the net growth efficiency (K2c) 49%.

At Kvarnbukten Bay there are large variations in size between the separate year-classes. The energy content of the food consumed by the 1975/76 cohort was used as follows: 4% was stored in living biomass after one year, 36% was released to other trophic levels as dead animals and cast carapaces, 13% rejected as faeces, and 47% was lost through respiration.  相似文献   


12.
M. Kitajima  W.L. Butler 《BBA》1975,408(3):297-305
The parameters listed in the title were determined within the context of a model for the photochemical apparatus of photosynthesis.

The fluorescence of variable yield at 750 nm at −196 °C is due to energy transfer from Photosystem II to Photosystem I. Fluorescence excitation spectra were measured at −196 °C at the minimum, FO, level and the maximum, FM, level of the emission at 750 nm. The difference spectrum, FMFO, which represents the excitation spectrum for FV is presented as a pure Photosystem II excitation spectrum. This spectrum shows a maximum at 677 nm, attributable to the antenna chlorophyll a of Photosystem II units, with a shoulder at 670 nm and a smaller maximum at 650 nm, presumably due to chlorophyll a and chlorophyll b of the light-harvesting chlorophyll complex.

Fluorescence at the FO level at 750 nm can be considered in two parts; one part due to the fraction of absorbed quanta, , which excites Photosystem I more-or-less directly and another part due to energy transfer from Photosystem II to Photosystem I. The latter contribution can be estimated from the ratio of FO/FV measured at 692 nm and the extent of FV at 750 nm. According to this procedure the excitation spectrum of Photosystem I at −196 °C was determined by subtracting 1/3 of the excitation spectrum of FV at 750 nm from the excitation spectrum of FO at 750 nm. The spectrum shows a relatively sharp maximum at 681 nm due to the antenna chlorophyll a of Photosystem I units with probably some energy transfer from the light-harvesting chlorophyll complex.

The wavelength dependence of was determined from fluorescence measurements at 692 and 750 nm at −196 °C. is constant to within a few percent from 400 to 680 nm, the maximum deviation being at 515 nm where shows a broad maximum increasing from 0.30 to 0.34. At wavelengths between 680 and 700 nm, increases to unity as Photosystem I becomes the dominant absorber in the photochemical apparatus.  相似文献   


13.
The phosphinoalkenes Ph2P(CH2)nCH=CH2 (n= 1, 2, 3) and phosphinoalkynes Ph2P(CH2)n C≡CR (R = H, N = 2, 3; R = CH3, N = 1) have been prepared and reacted with the dirhodium complex (η−C5H5)2Rh2(μ−CO) (μ−η2−CF3C2CF3). Six new complexes of the type (ν−C5H5)2(Rh2(CO) (μ−η11−CF3C2CF3)L, where L is a P-coordinated phosphinoalkene, or phosphinoalkyne have been isolated and fully characterized; the carbonyl and phosphine ligands are predominantly trans on the Rh---Rh bond, but there is spectroscopic evidence that a small amount of the cis-isomer is formed also. Treatment of the dirhodium-phosphinoalkene complexes with (η−CH3C5H4)Mn(CO)2thf resulted in coordination of the manganese to the alkene function. The Rh2---Mn complex [(η−C5H5)2Rh2(CO) (μ−η11−CF3C2CF3) {Ph2P(CH2)3CH=CH2} (η−CH3C5H4)Mn(CO)2] was fully characterized. Simi treatment of the dirhodium-phosphinoalkyne complexes with Co2(CO)8 resulted in the coordination of Co2(CO)6 to the alkyne function. The Rh2---Co2 complex [(η−C5H5)2Rh2(CO) (μ−η11−CF3C2CF3) {Ph2PCH2C≡CCH3}Co2(CO)2], C37H25Co2F6O7PRh2, was fully characteriz spectroscopically, and the molecular structure of this complex was determined by a single crystal X-ray diffraction study. It is triclinic, space group (Ci1, No. 2) with a = 18.454(6), B = 11.418(3), C = 10.124(3) Å, = 112.16(2), β = 102.34(3), γ = 91.62(3)°, Z = 2. Conventional R on |F| was 0.052 fo observed (I > 3σ(I)) reflections. The Rh2 and Co2 parts of the molecule are distinct, the carbonyl and phosphine are mutually trans on the Rh---Rh bond, and the orientations of the alkynes are parallel for Rh2 and perpendicular for Co2. Attempts to induce Rh2Co2 cluster formation were unsuccessful.  相似文献   

14.
Hydroxylated 2,19-methylene-bridged androstenediones were designed as potential mimics of enzyme oxidized intermediates of androstenedione. These compounds exhibited competitive inhibition with low micromolar affinities for aromatase. These inhibitory constants (Ki values) were 10 times greater than the 2,19-methylene-bridged androstenedione constant (Ki = 35–70 nM). However, expansion of the 2,19-carbon bridge to ethylene increased aromatase affinity by 10-fold (Ki = 2 nM). Substitution pf a methylene group with oxygen and sulfur in this expanded bridge resulted in Ki values of 7 and 20 nM, respectively. When the substituent was an NH group, the apparent inhibitory kinetics changed from competitive to uncompetitive. All of these analogs exhibited time-dependent inhibition of aromatase activity following preincubation of the inhibitor with human placental microsomes prior to measuring residual enzyme activity. Part of this inhibition was NADPH cofactor-dependent for the 2,19-methyleneoxy- but not for the 2,19-ethylene-bridged androstenedione. The time-dependent inhibition for these four analogs was very rapid since they exhibited τ50 values, the t1/2 for enzyme inhibition at infinite inhibitor concentration, of 1 to 3 min. These A-ring-bridged androstenedione analogs represent a novel series of potent steroidal aromatase inhibitors. The restrained A-ring bridge containing CH2, O, S, or NH could effectively coordinate with the heme of the P450 aromatase to allow the tight-binding affinities reflected by their nanomolar Ki values.  相似文献   

15.
The in vitro metabolism of cortisol in human liver fractions is highly complex and variable. Cytosolic metabolism proceeds predominantly via A-ring reduction (to give 3,5β-tetrahydrocortisol; 3,5β-THF), while microsomal incubations generate upto 7 metabolites, including 6β-hydroxycortisol (6β-OHF), and 6β-hydroxycortisone (6β-OHE), products of the cytochrome P450 (CYP) 3A subfamily. The aim of the present study was, therefore, to examine two of the main enzymes involved in cortisol metabolism, namely, microsomal 6β-hydroxylase and cytosolic 4-ene-reductase. In particular, we wished to assess the substrate specificity of these enzymes and identify compounds with inhibitory potential. Incubations for 30 min containing [3H]cortisol, potential inhibitors, microsomal or cytosolic protein (3 mg), and co-factors were followed by radiometric HPLC analysis. The Km value for 6β-OHF and 6β-OHE formation was 15.2 ± 2.1 μM (mean ± SD; n = 4) and the Vmax value 6.43 ± 0.45 pmol/min/mg microsomal protein. The most potent inhibitor of cortisol 6β-hydroxylase was ketoconazole (Ki = 0.9 ± 0.4 μM; N = 4), followed by gestodene (Ki = 5.6 ± 0.6 μM) and cyclosporine (Ki = 6.8 ± 1.4 μM). Both betamethasone and dexamethasone produced some inhibition (Ki = 31.3 and 54.5 μ, respectively). However, substrates for CYP2C (tolbutamide), CYP2D (quinidine), and CYP1A (theophylline) were essentially non-inhibitory. The Km value for cortisol 4-ene-reductase was 26.5 ± 11.2 μM (n = 4) and the Vmax value 107.7 ± 46.0 pmol/min/mg cytosolic protein. The most potent inhibitors were androstendione (Ki = 17.8 ± 3.3 μM) and gestodene (Ki = 23.8 ± 3.8 μM). Although both compounds have identical A-rings to cortisol, and undergo reduction, inhibition was non-competitive.  相似文献   

16.
卢森堡  陈云明  唐亚坤  吴旭  温杰 《生态学杂志》2017,28(11):3469-3478
以黄土丘陵区油松-沙棘混交林为研究对象,运用热扩散式探针(TDP)于2015年6—10月对油松和沙棘的树干液流密度(Fd)进行连续观测,同步测定了光合有效辐射(PAR)、水汽压亏缺(VPD)和土壤水分(SWC)等环境因子,分析两树种对降雨利用的差异.采用Threshold-delay 模型、多元回归分析和偏相关分析方法,研究两树种Fd对降雨的响应过程,并确定环境因子对Fd的影响.结果表明: 随着降雨量递增,两树种Fd的最大变化量都先上升后降低;其中0~1 mm降雨范围内,油松Fd(-16.3%)和沙棘Fd(-6.3%)都明显降低;1~5 mm降雨范围内,油松Fd(-0.4%)降低而沙棘Fd(9.0%)明显升高.油松和沙棘Fd对降雨响应的最小降雨阈值(RL)分别为6.4和1.9 mm,滞后时间(τ)为1.96和1.67 d.降雨前油松Fd峰值集中在12:00—12:30(70%),沙棘Fd峰值分别集中在10:30—12:00(48%)和16:00—16:30(30%);降雨后油松Fd峰值集中在11:00—13:00(40%),沙棘Fd峰值分别集中在12:00—13:00(52%)和16:30—17:00(24%).降雨前影响油松和沙棘Fd的环境因子大小顺序为PAR>VPD;降雨后影响油松Fd的环境因子大小顺序为PAR>VPD>0~20 cm SWC(SWC0~20),影响沙棘Fd的环境因子大小顺序为SWC0~20>PAR>VPD.油松-沙棘混交林对水分利用的稳定性较高.  相似文献   

17.
Conductance and relaxations of gelatin films in glassy and rubbery states   总被引:1,自引:0,他引:1  
The dielectric constant, ′, and the dielectric loss, ″, for gelatin films were measured in the glassy and rubbery states over a frequency range from 20 Hz to 10 MHz; ′ and ″ were transformed into M* formalism (M*=1/(′−i″)=M′+iM″; i, the imaginary unit). The peak of ″ was masked probably due to dc conduction, but the peak of M″, e.g. the conductivity relaxation, for the gelatin used was observed. By fitting the M″ data to the Havriliak–Negami type equation, the relaxation time, τHN, was evaluated. The value of the activation energy, Eτ, evaluated from an Arrhenius plot of 1/τHN, agreed well with that of Eσ evaluated from the DC conductivity σ0 both in the glassy and rubbery states, indicating that the conductivity relaxation observed for the gelatin films was ascribed to ionic conduction. The value of the activation energy in the glassy state was larger than that in the rubbery state.  相似文献   

18.
为了比较光系统II实际光化学量子效率(ΦPSII)对光的响应机理模型(简称机理模型)、负指数模型和指数模型的优缺点, 用LI-6400-40B光合作用测定仪控制CO2浓度和温度, 测量了剑叶金鸡菊(Coreopsis lanceolata)、黄荆(Vitex negundo)和大狼杷草(Bidens frondosa)的电子传递速率(ETR)对光的响应曲线(ETR-I)和ΦPSII对光的响应曲线(ΦPSII-I), 然后用这3个模型分别拟合了这些数据。拟合结果表明: 3个模型都可以较好地拟合这3种植物的ETR-I的响应数据和ΦPSII-I的响应数据, 但由指数模型拟合ETR-IΦPSII-I的响应数据得到相应的饱和光强(PARsat)和光系统II最大光能利用效率(Fv/Fm)之间存在显著差异, 且估算的饱和光强远低于实测值。由机理模型可知, ΦPSII不仅与光强的函数有关, 还与植物的内禀特性有关, 即与天线色素分子的本征光能吸收截面、激子的传递效率、能级的简并度、光化学反应常数、热耗散常数和处于最低激发态的平均寿命等参数有关。此外, 由机理模型还可知, ΦPSII随光强的增加而下降的原因是捕光色素分子的有效光能吸收截面随光强增加而降低。  相似文献   

19.
Esterification of lysophosphatidylcholine (LPC) with conjugated linoleic acid (CLA) was carried out using porcine pancreatic phospholipase A2 (PLA2). PLA2 only slightly synthesized phosphatidylcholine containing CLA (CLA-PC) at 2.6% by the addition of water. Addition of formamide in place of water markedly increased the yield of CLA-PC. In addition, synthesis of CLA-PC by PLA2 was affected by the amount of substrate CLA and PLA2 in the reaction system. Under optimal reaction conditions using 11 mg LPC, 18 mg CLA, 550 mg glycerol, 50 μL formamide, 3.3 × 104 U PLA2, and 0.3 μmol CaCl2 at 37 °C for 6 h, the reaction yield of CLA-PC reached 65 mol%. Furthermore, addition of protein such as albumin and casein suppressed the decrease of CLA-PC yield after 6 h. PLA2 exhibited the highest activity for the 10t,12c-CLA isomer among four CLA isomers (9c,11t-CLA, 9c,11c-CLA, 9t,11t-CLA and 10t,12c-CLA), whereas that for 9c,11c-CLA was the lowest. These results showed that the present esterification system for LPC and CLA by PLA2 is effective for producing CLA-PC.  相似文献   

20.
Extracellular action potentials produced by a muscle fibre of finite length were calculated for recordings at the skin surface. The sensitivity of power spectra to variations in propagation velocity (ν) and intracellular action potential (IAP) duration (Tin) was studied theoretically. The magnitude and distribution of the spectral power of muscle fibre potentials depend on the electrode longitudinal position. The relative shifts of the spectra in dB induced by variation in ν or Tin hardly depend on the longitudinal position of the electrode. A variation in ν affects only the power spectrum positive slope and the initial part of the high-frequency roll-off and a variation in Tin affects only the remaining part of the high-frequency roll-off. The total spectral amplitude is practically non-sensitive to variations in the wavelength, b = ν.Tin. The total power is sensitive to variations in ν, Tin as well as in b, and its relative changes depend on the electrode longitudinal position. The whole power spectrum is shifted along the frequency axis and mode (Fmax), median (Fmed) and mean (Fmean) frequencies have practically equal percentage changes only when ν and Tin vary jointly in such a way that the product ν.Tin keeps unchanged.  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