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.
Precise control of carrier concentration in both bulk and thin‐film materials is crucial for many solid‐state devices, including photovoltaic cells, superconductors, and high mobility transistors. For applications that span a wide temperature range (thermoelectric power generation being a prime example) the optimal carrier concentration varies as a function of temperature. This work presents a modified modulation doping method to engineer the temperature dependence of the carrier concentration by incorporating a nanosize secondary phase that controls the temperature‐dependent doping in the bulk matrix. This study demonstrates this technique by de‐doping the heavily defect‐doped degenerate semiconductor GeTe, thereby enhancing its average power factor by 100% at low temperatures, with no deterioration at high temperatures. This can be a general method to improve the average thermoelectric performance of many other materials.  相似文献   

2.
Effects of magnetic field (MF) at 50 Hz on chromatin conformation were studied by the method of anomalous viscosity time dependence (AVTD) in human lymphocytes from two healthy donors. MF within the peak amplitude range of 5-20 μT affected chromatin conformation. These MF effects differed significantly between studied donors, and depended on magnetic flux density and initial condensation of chromatin. While the initial state of chromatin was rather stable in one donor during one calendar year of measurements, the initial condensation varied significantly in cells from another donor. Both this variation and the MF effect depended on temperature during exposure. Despite these variations, the general rule was that MF condensed the relaxed chromatin and relaxed the condensed chromatin. Thus, in this study we show that individual effects of 50 Hz MF exposure at peak amplitudes within the range of 5-20 μT may be observed in human lymphocytes in dependence on the initial state of chromatin and temperature.  相似文献   

3.
The rate of translocation of the hydrophobic ion dipicrylamine across planar lipid membranes formed from dipalmitoyllecithin in n-decane was determined by voltage jump relaxation experiments. The activation energy of the rate constant shows a change from a positive to a negative value at about 42°C near the main phase transition temperature of this lipid. Below this temperature, the rate constant was found to increase with decreasing temperature. This anomalous behaviour extends over a temperature range of at least 10 K and may be formally interpreted as an enhanced mobility of dipicrylamine in the solid state of the membrane.  相似文献   

4.
Effect of microwave radiation with the frequency of 1000 +/- 10 MHz and specific absorption rate of 220-580 mV/g on the ferricyanide reduction by human red blood cells in the presence of methylene blue (carrier of oxidation-reduction equivalents through the membrane) was studied at different temperatures in the region of 23-34 degrees C. The temperature dependence of the ferricyanide reduction rate in Arrhenius plots shows two sharp "anomalous" sites with apparently negative activation energy at 26-27 and 29-30 degrees C. Broadness and expression of the "anomalous" sites increased with an increase of the blood storage time. The increase of the ferricyanide reduction rate under microwave irradiation was observed only in the temperature regions corresponding to the "anomalous" sites of the temperature dependence.  相似文献   

5.
The sensitivity of the method of anomalous time dependence of viscosity to changes in the conformation of DNA-protein complexes (such as nucleoide) by the action of increased temperature (33, 70 and 85 degrees C) and the combined action of temperature and Na+, Cl- ions on lysates of Escherichia coli AB1157 cells has been studied. The optimal conditions of the cell lysis was determined on the basis of the curve parameters of the anomalous time dependence of viscosity.  相似文献   

6.
P Smejtek  S Wang 《Biophysical journal》1991,59(5):1064-1073
Dipalmitoylphosphatidylcholine (DPPC) vesicles acquire negative surface charge on adsorption of negatively charged pentachlorophenolate (PCP-), and lipophilic ions tetraphenylborate (TPhB-), and dipicrylamine (DPA-). We have obtained (a) zeta-potential isotherms from the measurements of electrophoretic mobility of DPPC vesicles as a function of concentration of the adsorbing ions at different temperatures (25-42 degrees C), and (b) studied the effect of PCP- on gel-to-fluid phase transition by measuring the temperature dependence of zeta-potential at different PCP- concentrations. The zeta-potential isotherms of PCP- at 25, 32, and 34 degrees C correspond to adsorption to membrane in its gel phase. At 42 degrees C the zeta-potential isotherm corresponds to membrane in its fluid phase. These isotherms are well described by a Langmuir-Stern-Grahame adsorption model proposed by McLaughlin and Harary (1977. Biochemistry. 15:1941-1948). The zeta-potential isotherm at 37 degrees C does not follow the single-phase adsorption model. We have also observed anomalous adsorption isotherms for lipophilic ions TPhB- and DPA- at temperatures as low as 25 degrees C. These isotherms demonstrate a gel-to-fluid phase transition driven by ion adsorption to DPPC membrane during which the membrane changes from weakly to a strongly adsorbing state. The anomalous isotherm of PCP- and the temperature dependence of zeta-potential can be described by a two-phase model based on the combination of (a) Langmuir-Stern-Grahame model for each phase, (b) the coexistence of gel and fluid domains, and (c) depression of gel-to-fluid phase transition temperature by PCP-. Within the anomalous region the magnitude of zeta-potential rapidly increases concentration of adsorbing species, which was characterized in terms of a Esin-Markov coefficient. This effect can be exploited in membrane-based devices. Comments are also made on the possible effect of PCP, as an uncoupler, in energy transducing membranes.  相似文献   

7.
This study examines the relationship between the DNA binding thermodynamics and the enzymatic activity of the Klenow and Klentaq Pol I DNA polymerases from Escherichia coli and Thermus aquaticus. Both polymerases bind DNA with nanomolar affinity at temperatures down to at least 5 °C, but have lower than 1% enzymatic activity at these lower temperatures. For both polymerases it is found that the temperature of onset of significant enzymatic activity corresponds with the temperature where the enthalpy of binding (ΔHbinding) crosses zero (TH) and becomes favorable (negative). This TH/activity upshift temperature is 15 °C for Klenow and 30 °C for Klentaq. The results indicate that a negative free energy of DNA binding alone is not sufficient to proceed to catalysis, but that the enthalpic versus entropic balance of binding may be a modulator of the temperature dependence of enzymatic function. Analysis of the temperature dependence of the catalytic activity of Klentaq polymerase using expanded Eyring theory yields thermodynamic patterns for ΔG, ΔH, and TΔS that are highly analogous to those commonly observed for direct DNA binding. Eyring analysis also finds a significant ΔCp of formation of the activated complex, which in turn indicates that the temperature of maximal activity, after which incorporation rate slows with increasing temperature, will correspond with the temperature where the activation enthalpy (ΔH) switches from positive to negative.  相似文献   

8.
The kinetics of the metarhodopsin I–II reaction have been measured over a wide range of temperatures (1–37°C) and pH values (4.5–8) with suspensions containing fragments of bovine rod outer segments. It was found that for all conditions the occurrence of metarhodopsin II could be described by two independent first-order processes. The fast component: slow component amplitude ratio depends upon pH and temperature.The kinetics of the lumi-metarhodopsin I reaction show the same pH dependence for the fast component: slow component amplitude ratio as the one observed for the metarhodopsin II signals.All the results observed could be described with the assumption that rhodopsin itself exists in two conformational states before bleaching which are in a pH and temperature-dependent equilibrium. This hypothesis is confirmed by its ability to explain some apparently anomalous observations in the literature.  相似文献   

9.
10.
The thermodynamics of melting of bacteriophage T4 lysozyme and four of its mutants have been measured by van't Hoff methods. The effect of pH has been explored and utilized to obtain the dependence of the enthalpy on temperature as suggested by Privalov and co-workers. The enthalpy change is a steep linear function of temperature. ΔCp is large and constant within experimental error. Changes in ΔHu are as large as 30% for a single point mutation. Changes in enthalpy are largely compensated by changes in entropy. Changes in stability, as measured by the free energy of unfolding, are smaller than those of ΔH, but are very large in a relative sense, since ΔG is very much smaller than ΔH. Origins of the destabilization caused by mutations are discussed.  相似文献   

11.
The exchange kinetics of the slowest exchanging BPTI beta-sheet protons are complex compared to model peptides; the activation energy, E alpha, and the pH dependence are temperature dependent. We have measured the exchange kinetics in the range pH 1--11, 33--71 degrees C, particularly the temperature dependence. The data are fit to a model in which exchange of each proton is determined by two discrete dynamical processes, one with E alpha approximately 65 kcal/mol and less than first order dependence on catalyst ion, and one with E alpha 20--30 kcal/mol and approaching first order in catalyst ion. The low activation energy process is the mechanism of interest in the native conformation of globular proteins and involves low energy, small amplitude fluctuations; the high activation energy process involves major unfolding. The model is simple, has a precedent in the hydrogen exchange literature, and explains quantitatively the complex feature of the exchange kinetics of single protons in BPTI, including the following. For the slowest exchanging protons, in the range 36 degrees--68 degrees C, E alpha is approximately 65 kcal/mol at pH approximately 4, 20--30 kcal/mol at pH greater than 10, and rises to approximately 65 kcal/mol with increasing temperature at pH 6--10; the Arrhenius plots converge around 70 degrees C; the pH of minimum rate, pHmin, is greater than 1 pH unit higher at 68 degrees C than for model compounds; and at high pH, the pH-rate profiles shift to steeper slope; the exchange rates around pHmin are correlated to the thermal unfolding temperature in BPTI derivatives (Wagner and Wüthrich, 1979, J. Mol. Biol. 130:31). For the more rapidly exchanging protons in BPTI the model accounts for the observation of normal pHmin and E alpha of 20--30 kcal/mol at all pH's. The important results of our analysis are (a) rates for exchange from the folded state of proteins are not correlated to thermal lability, as proposed by Wuthrich et al. (1979, J. Mol. Biol. 134:75); (b) the unfolding rate for the BPTI cooperative thermal transition is equal to the observed exchange rates of the slowest exchanging protons between pH 8.4--9.6, 51 degrees C; (c) the rates for exchange of single protons from folded BPTI are consistent with our previous hydrogen-tritium exchange results and with a penetration model of the dynamic processes limiting hydrogen exchange.  相似文献   

12.
The kinetics of the metarhodopsin I-II reaction have been measured over a wide range of temperatures (1-37C ) and pH values (4.5-8) with suspensions containing fragments of bovine rod outer segments. It was found that for all conditions the occurrence of metarhodopsin II could be described by two independent first-order processes. The fast component: slow component amplitude ratio depends upon pH and temperature. The kinetics of the lumi-metarhodopsin I reaction show the same pH dependence for the fast component: slow component amplitude ratio as the one observed for the metarhodopsin II signals. All the results observed could be described with the assumption that rhodopsin itself exists in two conformational states before bleaching which are in a pH and temperature-dependent equilibrium. This hypothesis is confirmed by its ability to explain some apparently anomalous observations in the literature.  相似文献   

13.
A new theory for the electron transfer by the non-adiabatic process is formulated taking into account the origin shift and the frequency change of the vibration. The resultant formulas are quite similar to those of Jortner (Jortner, J. (1976) J. Chem. Phys. 64, 4860–4867) except that the free energy gap ΔG is used instead of the energy gap ΔE. By applying this theory to the photosynthetic electron transfer, the role of the remarkable temperature dependence of the electron transfer from cytochrome to P+ in Chromatium vinosum and the experimental data were reproduced very well using a small value of the coupling strength in contrast with the previous theory. This implies that proteins play a role to exclude many of the solvent molecules from the region of the electron transfer reaction between the donor and acceptor molecules. The negative activation process in the back electron transfer from Q?A to P+, the very slow back electron transfer from I? to P+ and the solvent isotope effect on the cytochrome oxidation are also successfully explained by this new theory. It is shown that even a qualitative conclusion as to the molecular parameters obtained from the temperature dependence of the electron transfer is different between the present theory and that of Jortner.  相似文献   

14.
The hypothesis that changes in the IP amplitude of the fluorescence transient OJIP reflect changes in leaf photosystem I (PSI) content was tested using mineral-deficient sugar beet plants. Young sugar beet plants (Beta vulgaris) were grown hydroponically on nutrient solutions containing either 1 mM or no Mg(2+) and 2.1 μM to 1.88 mM SO(4)(2-) for 4 weeks. During this period two leaf pairs were followed: the already developed second leaf pair and the third leaf pair that was budding at the start of the treatment. The IP amplitude [ΔF(IP) (fluorescence amplitude of the I-to-P-rise) and its relative contribution to the fluorescence rise: ΔV(IP) (amplitude of the relative variable fluorescence of the I-to-P-rise = relative contribution of the I-to-P-rise to the OJIP-rise)] and the amplitude of the transmission change at 820 nm (difference between all plastocyanin and the primary electron donor of photosystems I oxidized and reduced, respectively) relative to the total transmission signal (ΔI(max) /I(tot)) were determined as a function of the treatment time. Correlating the transmission and the two fluorescence parameters yielded approximately linear relationships in both cases. For the least severely affected leaves the parameter ΔV(IP) correlated considerably better with ΔI(max) /I(tot) than ΔF(IP) indicating that it is the ratio PSII:PSI that counts. To show that the relationship also holds for other plants and treatments, data from salt- and drought-stressed plants of barley, chickpea and pea are shown. The relationship between ΔV(IP) and PSI content was confirmed by western blot analysis using an antibody against psaD. The good correlations between ΔI(max) /I(tot) and ΔF(IP) and ΔV(IP) , respectively, suggest that changes in the IP amplitude can be used as semi-quantitative indicators for (relative) changes in the PSI content of the leaf.  相似文献   

15.
We analyzed spectra of light and heavy water at temperatures from 4 up to 50°C in a frequency range of 0.15 to 6.5 THz. It was shown that the amplitude of high-frequency relaxation absorption band with its maximum at 0.5 THz extends with increasing temperature and this temperature dependence for light water has a marked feature at 35–40°C as a sharp growth. This fact is noteworthy because this range corresponds to physiological values of a body temperature of the warm-blooded organisms. At the same time, the analogous temperature dependence for heavy water in the considered temperature range lacks this particular feature. Thus, the water with its properties differs significantly not only from other fluids, but also from its own isotopologues.  相似文献   

16.
The temperature dependence of the properties of unitary currents in cultured rat ventricular myocytes has been studied. Currents flowing through an ATP-dependent K+ channel were recorded from inside-out patches with the bath temperature varied from 10 degrees to 30 degrees C. The channel conductance was 56 pS at room temperature (22 degrees C), and the amplitudes of unitary currents and the channel conductance exhibited a relatively weak (Q10 from 1.4 to 1.6) dependence on temperature. The temperature dependence of channel mean open times was biphasic with the low temperature (10-20 degrees C) range showing a relatively stronger temperature dependence (Q10 of 2.3) than the high temperature (20-30 degrees C) range (Q10 of 1.6). The activation energies for the two regions were determined from an Arrhenius plot with the activation energy, corresponding to the lower temperature range, near 16 kcal/mol. Thermodynamic analysis, using transition rate theory, indicated that the formation of a transition state prior to channel closure to be associated with a positive entropy component for the high Q10 region.  相似文献   

17.
To determine the role of photon energy on charge generation in bulk heterojunction solar cells, the bias voltage dependence of photocurrent for excitation with photon energies below and above the optical band gap is investigated in two structurally related polymer solar cells. Charges generated in (poly[2,6‐(4,4‐bis(2‐ethylhexyl)‐4H‐cyclopenta[2,1‐b;3,4‐b′′]dithiophene)‐alt‐4,7‐(2,1,3‐benzothia­diazole)] (C‐PCPDTBT):[6,6]‐phenyl‐C61‐butyric acid methyl ester (PCBM) solar cells via excitation of the low‐energy charge transfer (CT) state, situated below the optical band gap, need more voltage to be extracted than charges generated with excitation above the optical band gap. This indicates a lower effective binding energy of the photogenerated electrons and holes when the excitation is above the optical band gap than when excitation is to the bottom of the CT state. In blends of PCBM with the silicon‐analogue, poly[(4,4‐bis(2‐ethylhexyl)dithieno[3,2‐b:2,3d]silole)‐2,6‐diyl‐alt‐(2,1,3‐benzothiadiazole)‐4,7‐diyl] (Si‐PCPDTBT), there is no effect of the photon energy on the electric field dependence of the dissociation efficiency of the CT state. C‐PCPDTBT and Si‐PCPDTBT have very similar electronic properties, but their blends with PCBM differ in the nanoscale phase separation. The morphology is coarser and more crystalline in Si‐PCPDTBT:PCBM blends. The results demonstrate that the nanomorphological properties of the bulk heterojunction are important for determining the effective binding energy in the generation of free charges at the heterojunction.  相似文献   

18.
During the synthesis of integral membrane proteins (IMPs), the hydrophobic amino acids of the polypeptide sequence are partitioned mostly into the membrane interior and hydrophilic amino acids mostly into the aqueous exterior. Using a many-body statistical mechanics model, we analyze the minimum free energy state of polypeptide sequences partitioned into α-helical transmembrane (TM) segments and the role of thermal fluctuations. Results suggest that IMP TM segment partitioning shares important features with general theories of protein folding. For random polypeptide sequences, the minimum free energy state at room temperature is characterized by fluctuations in the number of TM segments with very long relaxation times. Moreover, simple assembly scenarios do not produce a unique number of TM segments due to jamming phenomena. On the other hand, for polypeptide sequences corresponding to actual IMPs, the minimum free energy structure with the wild-type number of segments is free of number fluctuations due to an anomalously large gap in the energy spectrum. Now, simple assembly scenarios do reproduce the minimum free energy state without jamming. Finally, we find a threshold number of random point mutations where the size of the anomalous gap is reduced to the point that the wild-type ground state is destabilized and number fluctuations reappear.  相似文献   

19.
The bacterial flagellar motor can rotate either clockwise (CW) or counterclockwise (CCW). Three flagellar proteins, FliG, FliM, and FliN, are required for rapid switching between the CW and CCW directions. Switching is achieved by a conformational change in FliG induced by the binding of a chemotaxis signaling protein, phospho-CheY, to FliM and FliN. FliG consists of three domains, FliG(N), FliG(M), and FliG(C), and forms a ring on the cytoplasmic face of the MS ring of the flagellar basal body. Crystal structures have been reported for the FliG(MC) domains of Thermotoga maritima, which consist of the FliG(M) and FliG(C) domains and a helix E that connects these two domains, and full-length FliG of Aquifex aeolicus. However, the basis for the switching mechanism is based only on previously obtained genetic data and is hence rather indirect. We characterized a CW-biased mutant (fliG(ΔPAA)) of Salmonella enterica by direct observation of rotation of a single motor at high temporal and spatial resolution. We also determined the crystal structure of the FliG(MC) domains of an equivalent deletion mutant variant of T. maritima (fliG(ΔPEV)). The FliG(ΔPAA) motor produced torque at wild-type levels under a wide range of external load conditions. The wild-type motors rotated exclusively in the CCW direction under our experimental conditions, whereas the mutant motors rotated only in the CW direction. This result suggests that wild-type FliG is more stable in the CCW state than in the CW state, whereas FliG(ΔPAA) is more stable in the CW state than in the CCW state. The structure of the TM-FliG(MC)(ΔPEV) revealed that extremely CW-biased rotation was caused by a conformational change in helix E. Although the arrangement of FliG(C) relative to FliG(M) in a single molecule was different among the three crystals, a conserved FliG(M)-FliG(C) unit was observed in all three of them. We suggest that the conserved FliG(M)-FliG(C) unit is the basic functional element in the rotor ring and that the PAA deletion induces a conformational change in a hinge-loop between FliG(M) and helix E to achieve the CW state of the FliG ring. We also propose a novel model for the arrangement of FliG subunits within the motor. The model is in agreement with the previous mutational and cross-linking experiments and explains the cooperative switching mechanism of the flagellar motor.  相似文献   

20.
Multi-proton spin-echo images were collected from cold-acclimated winter wheat crowns (Triticum aestivum L.) cv. Cappelle Desprez at 400 MHz between 4 and ?4 °C. Water proton relaxation by the spin-spin (T2) mechanism from individual voxels in image slices was found to be mono-exponential. The temperature dependence of these relaxation rates was found to obey Arrhenius or absolute rate theory expressions relating temperature, activation energies and relaxation rates, Images whose contrast is proportional to the Arrhenius activation energy (Ea), Gibb's free energy of activation (ΔG?), and the entropy of activation (ΔS?) for water relaxation on a voxel basis were constructed by post-image processing. These new images exhibit contrast based on activation energies rather than rules of proton relaxation. The temperature dependence of water proton T2 relaxation rates permits prediction of changes in the physical state of water in this tissue over modest temperature ranges. A simple model is proposed to predict the freezing temperature kof various tissue in wheat crowns. The average Ea and ΔH? for water proton T2 relaxation over the above temperature range in winter wheat tissue were ?6.4 ± 14.8 and ?8.6 ± 14.8kj mol?1, respectively. This barrier is considerably lower than the Ea for proton translation in ice at 0°C, which is reported to be between 46.0 and 56.5 kj mol?1  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