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 31 毫秒
1.
Abstract— Absorption changes attributed to the triplet state of carotenoids and to primary electron donors (P-700. P-680): and fluorescence quenching at several wavelengths have been measured with a single apparatus. following flash excitation with a dye or a ruby laser. Spinach chloroplasts as well as subchloroplast particles enriched in Photosystem-1 (F1), Photosystem-2 (F1) or the light-harvesting Chl a/h (FIII) have been examined at temperatures varying between 5 and 294 K.
The triplet state of carotenoids has been identified on the basis of its difference spectrum (having a peak at 515 nm) and decay kinetics (⋍ 7 µs at low temperature; accelerated by O2 at 294 K). It is formed in all of the materials studied. The quantum yield of carotenoid triplet formation in chloroplasts increases at low temperature, but less than the fluorescence yield.
In most cases the fluorescence quenching recovers approximately with the same kinetics as the decay of the carotenoid triplets. The fluorescence recovery is, however, significantly faster for chloroplasts at 730 nm. Fluorescence quenching occurs in all types of materials. The ratio of fluorescence quenching to the concentration of carotenoid triplets varies with the material, being maximum in chloroplasts and minimum in Fm particles.
We conclude that the formation of the carotenoid triplet state is not limited to a few sites in the chloroplast and that a carotenoid triplet is a quencher of chlorophyll fluorescence. A detailed comparison of carotenoid triplets and fluorescence quenching gives some information concerning the organization of the pigments in the photosynthetic apparatus.  相似文献   

2.
The thermal decomposition behavior and the pyrolysis products of benzyl‐2,3,4,6‐tetra‐O‐acetyl‐β‐D‐glucopyranoside (BGLU) were studied with synchrotron vacuum ultraviolet (VUV) photoionization mass spectrometry at temperatures of 300, 500 and 700 °C at 0.062 Pa. Several pyrolysis products and intermediates were identified by the measurement of photoionization mass spectra at different photon energies. The results indicated that the primary decomposition reaction was the cleavage of O‐glycosidic bond of the glycoside at low temperature, proven by the discoveries of benzyloxy radical (m/z = 107) and glycon radical (m/z = 331) in mass spectra. As pyrolysis temperature increased from 300 to 700 °C, two possible pyrolytic modes were observed. This work reported an application of synchrotron VUV photoionization mass spectrometry in the study of the thermal decomposition of glycoside flavor precursor, which was expected to help understand the thermal decomposition mechanism of this type of compound. The possibility of this glycoside to be used as a flavor precursor in high temperature process was evaluated.  相似文献   

3.
A study of the Pd-containing catalyst based on manganese hexaaluminate by X-ray diffraction analysis, transmission electron microscopy, thermo-programmed reduction with hydrogen, and thermoprogrammed oxidation showed the hysteresis manifested itself in a difference between the temperature dependences of its catalytic activity in the oxidation of methane under the conditions of sample heating or cooling is related to the decomposition of PdO to metallic palladium at temperatures higher than 700–800°C and the subsequent formation of PdO nanoparticles, which are localized on the surface of metallic Pd (30–50 nm) in the form of polycrystalline films with a thickness of 2–5 nm, at temperatures lower than 600–700°C. A near-surface oxide film is formed under the conditions of cooling in oxygen-containing atmospheres, including in the presence of methane. The value of the hysteresis depends on the heat treatment temperature of the catalyst, and it reaches a maximum on the catalysts calcined at 900 and 1000°C.  相似文献   

4.
The purpose of this study is to research the thermal properties of spreads with maltitol. Thermal characteristics of spreads depend on process parameters (temperature, mixer speed rotation). Spreads are produced at different temperatures (30, 35, and 40 °C) and mixer speed rotation (1, 1.33, and 1.67 Hz). The thermogravimetric method shows the peak position and determinate the spread composition. The temperature decomposition of sucrose and maltitol is two stages (two peaks), and palm fat has a single stage decomposition (one peak). Maltitol peak is dominant for spreads containing 100 and 70 % maltitol as a sweetener. This peak is sharper than sucrose peak and the inflection point is more expressed. Shape and the position of these peaks in spreads are modified. Peaks of maltitol, palm fat, and sucrose in spreads are lower and wider because of the grinding process and the interaction between spread ingredients. Increasing the process parameters (temperature, mixer speed rotation), temperatures of these peaks are higher (closer to temperature peak of pure ingredients). The dominant parameter is mixer speed rotation. The most thermally stable spreads with any amount of maltitol are produced at a temperature of 40 °C and high mixer speed rotation (1.33 and 1.67 Hz), while the least stable maltitol spreads are produced at minimum process parameters (30 °C, 1 Hz).  相似文献   

5.
The thermal decomposition of nitromethane (NM) over the temperature range from 580 to 700 K at pressures of 4 Torr to 40 atm was analyzed. On the basis of literature data, with the use of theoretical transitional curves of the modified Kassel integral, the rate constants k of NM decomposition at the upper pressure limit were determined. The values thus obtained are in good agreement with the results of extrapolation of the high-temperature (1000–1400 K) k 1, ∞ values to lower temperatures. The reasons for which the NM decomposition rate constants differ by two orders of magnitude at low temperatures are considered. A general expression for the NM decomposition rate constant at the upper pressure limit over the 580–1400 K temperature range was determined: k 1, ∞ = (1.8 ± 0.7) × 1016 exp((?58.5 ± 2)/R T ) s?1. These data disprove the hypothesis that a nitro-nitrite rearrangement takes place during the NM decomposition at low temperatures.  相似文献   

6.
A microkinetic model is developed to study the reactivity of an O/O(2) gas mixture over a β-cristobalite (001) surface. The thermal rate constants for the relevant elementary processes are either inferred from quasiclassical trajectory calculations or using some statistical approaches, resting on a recently developed interpolated multidimensional potential energy surface based on density functional theory. The kinetic model predicts a large molecular coverage at temperatures lower than 1000 K, in contrary to a large atomic coverage at higher temperatures. The computed atomic oxygen recombination coefficient, mainly involving atomic adsorption and Eley-Rideal recombination, is small and increases with temperature in the 700-1700 K range (0.01 < γ(O) < 0.02) in good agreement with experiments. In the same temperature range, the estimated chemical energy accommodation coefficient, the main contribution to which is the atomic adsorption process is almost constant and differs from unity (0.75 < β(O) < 0.80).  相似文献   

7.
Abstract— In the reaction center of photosynthetic bacteria, with the primary ubiquinone reduced, the triplet state PR of the primary electron donor (a pair of bacteriochlorophylls named P) is PO ulated with a takes place in a few ns. We measured by flash absorption spectroscopy the influence of temperature on formation and decay kinetics of PR and 3Car in the reaction center of several strains of R. sphaeroides . The rate of triplet energy transfer, measured as the decay of PR after a flash, decreases when the temperature is lowered. Between 60 and 30 K the half-time of energy transfer becomes longer than the 3Car half-time decay (about 6 μs) and below 20 K the transfer is slower than the internal decay of PR (about 100 μs). In several cases it is clear that PR and 3Car decay independently and are not in thermal equilibrium. The singlet energy transfer from carotenoid to P occurs with a high efficiency at all temperatures.
The data can be accounted for on the basis of estimated energy levels of PR and 3Car, in the context of the equilibrium 3P ←3D where 3P is the localized triplet state of P-870 and 3D is another triplet state. A reasonable kinetic scheme leads us to estimate that 3D is 0.0025 ± 0.005 eV above 3P. 3D may thus be the state observed by Shuvalov and Parson (1981). We propose that both triplet and singlet energy transfer between P and the carotenoid occur via a bacteriochlorophyll, to which the carotenoid should be tightly coupled via exchange interaction.  相似文献   

8.
An easy alkoxide-based sol–gel method based on Ca(NO3)2·4H2O and triethyl phosphate [PO(OC2H5)3; TEP] as Ca and P precursors have been developed to synthesize nano-hydroxyapatite (HA). The structural evolution of the samples was studied using X-ray diffraction (XRD), thermal behavior, infrared analysis, and elemental analysis via scanning electron microscopy. It is noticeable that raising of the firing temperature resulted in increasing the HA content as the dominant phase at 600 and 700 °C. The phase transformation from amorphous to crystalline HA occurred at the low temperature of 400 °C, while at higher temperatures other Ca–P compounds as secondary phases transformed to HA. The crystallite size distributions and micro-strain of the HA samples produced were characterized by XRD methods with the aid of Scherrer and Williamson–Hall equations. The results of transmission electron microscopy as a complementary and reliable technique are in good agreement with those obtained from XRD. The results indicate that increasing the firing temperature caused permanent growth of mean crystallite size and a decrease in micro-strain.  相似文献   

9.
Scratch resistant sol–gel metal oxide coatings typically require a thermal post-treatment step (curing process) at temperatures between 400 and 700 °C. In this report, we demonstrate that the in situ generation of amines within sol–gel films facilitates the preparation of scratch resistant metal oxide coatings at a curing temperature of 250 °C. A selected series of carbamates was added to sol–gel formulations and included in the resulting xerogel films. During the curing process, the carbamates decomposed to liberate primary amines. These catalysed the curing process, finally leading to scratch resistant metal oxide coatings at low cure temperature.  相似文献   

10.
Au/TiO2 catalysts prepared by a deposition–precipitation process and used for CO oxidation without previous calcination exhibited high, largely temperature‐independent conversions at low temperatures, with apparent activation energies of about zero. Thermal treatments, such as He at 623 K, changed the conversion–temperature characteristics to the well‐known S‐shape, with activation energies slightly below 30 kJ mol?1. Sample characterization by XAFS and electron microscopy and a low‐temperature IR study of CO adsorption and oxidation showed that CO can be oxidized by gas‐phase O2 at 90 K already over the freeze‐dried catalyst in the initial state that contained Au exclusively in the +3 oxidation state. CO conversion after activation in the feed at 303 K is due to AuIII‐containing sites at low temperatures, while Au0 dominates conversion at higher temperatures. After thermal treatments, CO conversion in the whole investigated temperature range results from sites containing exclusively Au0.  相似文献   

11.
The dynamics of fluorescence decay and charge recombination were studied in the ether-extracted photosystem I reaction center isolated from spinach with picosecond resolution over a wide time range up to 100 ns. At all temperatures from 268 to 77 K, a slow fluorescence decay component with a 30-40 ns lifetime was detected. This component was interpreted as a delayed fluorescence emitted from the singlet excited state of the primary donor P700*, which is repopulated through charge recombination that was increased by the lack of secondary acceptor phylloquinone in the sample. Analysis of the fluorescence kinetics allowed estimation of the standard free-energy difference -DeltaG between P700* and the primary radical pair (P700(+)A0(-)) state over a wide temperature range. The values of -DeltaG were estimated to be 160/36 meV at 268/77 K, indicating its high sensitivity to temperature. A temperature-dependent -DeltaG value was also estimated in the delayed fluorescence of the isolated photosystem I in which the secondary acceptor quinone was partially prereduced by preillumination in the presence of dithionite. The results revealed that the temperature-dependent -DeltaG is a universal phenomenon common with the purple bacterial reaction centers, photosystem II and photosystem I reaction centers.  相似文献   

12.
Room-temperature Ba deposition on an oxygen-terminated theta-Al(2)O(3)/NiAl(100) ultrathin film substrate under ultrahigh vacuum (UHV) conditions is studied using X-ray photoelectron spectroscopy (XPS), Auger electron spectroscopy (AES) and temperature programmed desorption (TPD) techniques. In addition, Ba oxidation by the ions of the alumina substrate at 300 K < T < 1200 K in the absence of a gas-phase oxidizing agent is investigated. Our results indicate that at room temperature Ba grows in a layer-by-layer fashion for the first two layers, and Ba is partially oxidized. Annealing at T < 700 K results in further oxidation of the Ba species, whereas annealing at higher temperatures leads to loss of Ba from the surface via desorption and subsurface diffusion.  相似文献   

13.
The early stages of ceria growth on Rh(111) at high temperature have been investigated by low-energy electron microscopy and photoemission electron microscopy. Ceria was deposited by reactive Ce deposition at substrate temperatures between 700°C and 900°C in an oxygen ambient of 5 × 10−7 Torr. At 700°C, we observe a high nucleation density of 100-nm-sized islands. With elevated temperature, the average island size increases, and the nucleation density decreases. Triangularly shaped islands nucleate preferentially at step edges, with seemingly abrupt interfaces between Ce and Rh. At 900°C, the island edges are still straight, but during growth the islands lose their triangular form. Instead, growth along the substrate step edges becomes favorable, leading to a maze-like morphology. Atomic force microscopy reveals islands of 0.3 to 0.6-nm height, consistent with ceria islands formed by one or two trilayers (O―Ce―O) of ceria. Moreover, the second layer of the islands is also triangularly shaped, with lateral dimensions of 50 nm and similar step heights. IV-LEEM analysis leads to the conclusion that the rhodium surface is covered by a layer of reduced cerium oxide, which is partially overgrown by smaller islands of CeO2.  相似文献   

14.
Studies have been made of the secondary relaxation processes in the solid state of a number of polymers containing aromatic groups in the polymer chain. The polymers investigated include one, polystyrene, with the aromatic group in side-chain positions, and six high polymers in which phenylene rings lie in the main backbone chain. In polystyrene, wagging and torsional motions of the side chain phenyl rings give rise to a low-temperature δ-relaxation which is centered at 33°K (1.7 Hz) and which has an activation energy of about 2.3 kcal/mol. Most of the polymers with phenylene rings in the main chain exhibit a low-temperature relaxation in the temperature region from 100°–200°K. This relaxation process is centered at 159°K (0.54 Hz) in poly-p-xylylene, at 162°K (0.67 Hz) in polysulfone, and at 165°K (1.24 Hz) in poly(diancarbonate). In poly(2,6-dimethyl-p-phenylene oxide), two overlapping low-temperature relaxations are found, one in the range 125–140°K and the other near 277°K (ca. 1 Hz). The low-temperature secondary relaxation process in all of these polymers is believed to be associated with local reorientational motion of the phenylene, or substituted phenylene, rings or with combined motion of the phenylene rings and nearby chain units. For these low temperature relaxation processes, the activation energy is about 10 kcal/-mole. The temperature location of the relaxation appears to depend on the specific units to which the phenylene rings are attached and on steric and polar effects caused by substituents on the ring. In the poly-p-xylylenes the relaxation is shifted to much higher temperatures by a single Cl substitution on the ring but remains at essentially the same temperature position when dichlorosubstitution is made. The effects of water on the magnitude and temperature location of the observed low temperature relaxations, as well as the implications of the study for other polymers containing aromatic groups in their backbone chains, are discussed.  相似文献   

15.
Abstract— A chlorophyll (Chl) a solution in 3-methylpentane at 77 K exhibits an absorption spectrum with a distinct peak at 706 nm in the red-band region. The formation of the 706 nm absorbing species (S706) was reversible with respect to temperature change; no chemical change was observed. γ-Irradiation of the rigid 3-methylpentane solution at 77 K yields an absorption spectrum which can be ascribed to S706+ and S706. When carbon tetrachloride, an electron scavenger, was added to the solution, the absorption of S706+ survived, which has peaks at 850 and 956 nm. It is assumed that the S706 is hydrogen-bonded dimeric Chi a , which may be regarded as a model of P700 in photosynthesis. Cation radicals of monomeric Chi a were formed in a γ-irradiated sec -butyl chloride solution at 77 K, and an absorption spectrum with peaks at 730 and 813 nm was recorded. ESR spectra of the cation radicals of S706 and monomeric Chi a are of a similar shape but their linewidths are 7.5 and 11.0 Gauss, respectively. The linewidth narrowing observed for S706+ is clear evidence for the assumption that S706 is dimeric Chi a. Comparison was made of the absorption spectrum of S706+ with the light-induced spectrum of P700 reported earlier.  相似文献   

16.
Using an in situ pulsed laser photolysis/pulsed laser‐induced fluorescence/technique, the OH reaction kinetics of a three‐ring polycyclic aromatic hydrocarbon (PAH), phenanthrene (and its deuterated form), was investigated over the temperature range of 373–1000 K. This study represents the first examination of the OH kinetics for phenanthrene at elevated temperatures using the absolute rate technique. The phenanthrene results indicate a temperature dependence similar to its isomer anthracene, reported previously in R. Ananthula, T. Yamada, and P. H. Taylor, J Phys Chem A 2006, 100, 3559–3566, over a similar temperature range. The phenanthrene rate constants are similar to anthracene at high temperature (ca. 1000 K) and a factor of ca. 2 lower at low temperatures (373–700 K). The rate measurements were best fitted by the following two‐parameter expression of the form ATn: k1(373–1000 K) = 4.98 ± 2.96 × 10?6 * T?1.97±0.10 (in units of cm3 molecule?1 s?1, error limits ±1σ). Rate measurements with deuterated phenanthrene below 725 K were indistinguishable from the phenanthrene rate measurements, within random error limits, providing strong evidence for an OH‐addition mechanism. The effects of PAH size on their reactivity with OH radicals based on selected data over the temperature range of 243–1200 K are discussed. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 629–637, 2007  相似文献   

17.
Classical equilibrium molecular dynamics (MD) simulations have been performed to investigate the dynamical and energetic properties in hydrogen and mixed hydrogen-tetrahydrofuran sII hydrates at 30 and 200 K and 0.05 kbar, and also at intermediate temperatures, using SPC/E and TIP4P-2005 water models. The potential model is found to have a large impact on overall density, with the TIP4P-2005 systems being on average 1% more dense than their SPC/E counterparts, due to the greater guest-host interaction energy. For the lightly-filled mixed H(2)-THF system, in which there is single H(2) occupation of the small cage (1s1l), we find that the largest contribution to the interaction energy of both types of guest is the van der Waals component with the surrounding water molecules in the constituent cavities. For the more densely-filled mixed H(2)-THF system, in which there is double H(2) occupation in the small cage (2s1l), we find that there is no dominant component (i.e., van der Waals or Coulombic) in the H(2) interaction energy with the rest of the system, but for the THF molecules, the dominant contribution is again the van der Waals interaction with the surrounding cage-water molecules; again, the Coulombic component increases in importance with increasing temperature. The lightly-filled pure H(2) hydrate (1s4l) system exhibits a similar pattern vis-à-vis the H(2) interaction energy as for the lightly-filled mixed H(2)-THF system, and for the more densely-filled pure H(2) system (2s4l), there is no dominant component of interaction energy, due to the multiple occupancy of the cavities. By consideration of Kubic harmonics, there is some evidence of preferential alignment of the THF molecules, particularly at 200 K; this was found to arise at higher temperatures due to transient hydrogen bonding of the oxygen atom in THF molecules with the surrounding cage-water molecules.  相似文献   

18.
The kinetics of the reaction between γ-Al2O3 and gaseous CCl4 has been studied by isothermal TG measurements in the temperature range 700—1123 K. The reaction starts with a weight gain which can be attributed to the chemisorption of the reactive gas. The weight loss vs. time curves at relatively high temperatures can be described by the contracting cylinder equation and at relatively low temperatures by first-order kinetics. The dependence of the initial reaction rate on the CCl4 partial pressure follows the Langmuir—Hinshelwood rate expression. At 700—723 K, chemical control is thought to be predominant and an apparent activation energy of 212 kJ mole?1 is found for the chlorination process.  相似文献   

19.
Rate constants and product ion distributions for the O- and O2- reactions with O2(a 1Deltag) were measured as a function of temperature from 200 to 700 K. The measurements were made in a selected ion flow tube (SIFT) using a newly calibrated O2(a 1Deltag) emission detection scheme with a chemical singlet oxygen generator. The rate constant for the O2- reaction is approximately 7 x 10(-10) cm3 s-1 at all temperatures, approaching the Langevin collision rate constant. Electron detachment was the only product observed with O2-. The O- reaction shows a positive temperature dependence in the rate constant from 200 to 700 K. The product branching ratios show that almost all of the products at 200 K are electron detachment, with an increasing contribution from the slightly endothermic charge-transfer channel up to 700 K, accounting for 75% of the products at that temperature. The increase in the overall rate constant can be attributed to this increase in the contribution the endothermic channel. The charge-transfer product channel rate constant follows the Arrhenius form, and the detachment product channel rate constant is essentially independent of temperature with a value of approximately 6.1 x 10(-11) cm3 s-1.  相似文献   

20.
The formation of radicals from the gas-phase pyrolysis of phenol over a temperature range of 400-1000 degrees C was studied using the technique of low temperature matrix isolation electron paramagnetic resonance (LTMI EPR). Cooling the reactor effluent in a CO2 carrier gas to 77 K produces a cryogenic matrix that exhibits complex EPR spectra. However, annealing by slowly raising the matrix temperature yielded well-resolved, identifiable spectra. All annealed spectra over the temperature range of 700-1000 degrees C resulted in the generation of EPR spectra with six lines, hyperfine splitting constant approximately 6.0 G, g = 2.00430, and peak-to-peak width approximately 3 G that was readily assignable, based on comparison with the literature and theoretical calculations, as that of cyclopentadienyl radical. Annihilation procedures along with microwave power saturation experiments helped to clearly identify phenoxy radicals in the same temperature region. Conclusive identifications of cyclopentadienyl and phenoxy radicals were based on pure spectra of these radicals under the same experimental conditions generated from suitable precursors. Cyclopentadienyl is clearly the dominant radical at temperatures above 700 degrees C and is observed at temperatures as low as 400 degrees C. The low-temperature formation is attributed to heterogeneous initiation of phenol decomposition under very low pressure conditions. The high cyclopentadienyl to phenoxy ratio was consistent with the results of reaction kinetic modeling calculations using the CHEMKIN kinetic package and a phenol pyrolysis model adapted from the literature.  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