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.
The changes of decay rates of radionuclide 111in (electron capture) and 32p (β decay) induced by external mechanic motion are studied. The results indicate that, in the external circular rotation in clockwise and anticlockwise centrifuge on Northern Hemisphere (radius 8 cm, 2000 r/min), the half life of 111in compared with the referred (2.83 d) is decreased at 2.83% and increased at 1.77%, respectively; the half life of 32P compared with the referred (14.29 d) is decreased at 3.78% and increased at 1.75%, respectively. When the clockwise and anticlockwise rotations increase to 4000 r/min, the half life of 111in is decreased at 11.31% and increased at 6.36%, respectively; the half life of 32P is decreased at 10.08% and increased at 4.34%, respectively. When the circular rotation is removed, the decay rates of 111in and 32P return back to the referred, respectively. It is found that the external circular rotations in clockwise and anticlockwise centrifuge selectively increased and decreased the decay rates of 111in and 32p, respectively, and the effects are strongly dependent on the strength of circular rotation. It is suggested that these effects may be caused by the chiral interaction.  相似文献   

2.
The influence of mechanic motion on the decay rate of 32P was studied by means of liquid scintillation counting (LSC). The results indicate that, on the Northern Hemisphere, the half life of 32P in anticlockwise circular centrifuge rotation (radius 10 cm, 4000 r/min) equals the one in natural conditions within 0.6 percent uncertainty.  相似文献   

3.
The influence of mechanic motion on the decay rate of 32P was studied by means of liquid scintillation counting (LSC). The results indicate that, on the Northern Hemisphere, the half life of 32P in anticlockwise circular centrifuge rotation (radius 10 cm, 4000 r/min) equals the one in natural conditions within 0.6 percent uncertainty.  相似文献   

4.
A polar electro-optic response is observed in droplets of an achiral nematic liquid crystal in coexistence with the isotropic phase. Between crossed polarizers each pancake-shaped droplet shows extinction brushes in the form of a centred cross aligned with the polarizer axes. An applied electric field E induces a rotation of the crosses about the field direction, with about half the droplets switching clockwise and the other half anticlockwise. The sense of rotation in each droplet changes when E is reversed. We propose that a twisted bipolar director structure is stabilized in the droplets by a relatively large splay elastic constant and tangential boundary conditions. The molecules twist along the diameter of the droplets, perpendicular to the applied field, which results in a linear rotation of the director by the inverse flexoelectric effect. Since the molecules are achiral, the handedness of the twist, and hence the sense of the switching, in any droplet is arbitrary.  相似文献   

5.
《Liquid crystals》1999,26(10):1555-1561
A polar electro-optic response is observed in droplets of an achiral nematic liquid crystal in coexistence with the isotropic phase. Between crossed polarizers each pancake-shaped droplet shows extinction brushes in the form of a centred cross aligned with the polarizer axes. An applied electric field E induces a rotation of the crosses about the field direction, with about half the droplets switching clockwise and the other half anticlockwise. The sense of rotation in each droplet changes when E is reversed. We propose that a twisted bipolar director structure is stabilized in the droplets by a relatively large splay elastic constant and tangential boundary conditions. The molecules twist along the diameter of the droplets, perpendicular to the applied field, which results in a linear rotation of the director by the inverse flexoelectric effect. Since the molecules are achiral, the handedness of the twist, and hence the sense of the switching, in any droplet is arbitrary.  相似文献   

6.
Coupled cluster and density functional models of specific rotation and vacuum UV (VUV) absorption and circular dichroism spectra are reported for the conformationally flexible molecules (R)-3-chloro-1-butene and (R)-2-chlorobutane. Coupled cluster length- and modified-velocity-gauge representations of the Rosenfeld optical activity tensor yield significantly different specific rotations for (R)-3-chloro-1-butene, with the latter providing much closer comparison (within 3%) to the available gas-phase experimental data at 355 and 633 nm. Density functional theory overestimates the experimental rotations for (R)-3-chloro-1-butene by approximately 80%. For (R)-2-chlorobutane, on the other hand, all three models give reasonable comparison to experiment. The theoretical specific rotations of the individual conformers of (R)-3-chloro-1-butene are much larger than those of (R)-2-chlorobutane, in disagreement with previous studies of the temperature dependence of the experimental rotations in solution. Simulations of VUV absorption and circular dichroism spectra reveal large differences between the coupled cluster and density functional excitation energies and the rotational strengths. However, while these differences lead to very different specific rotations for (R)-3-chloro-1-butene, they have much less impact on the computed specific rotations for (R)-2-chlorobutane. In addition, the coupled cluster VUV absorption spectrum of (R)-2-chlorobutane compares well to experiment.  相似文献   

7.
We observed the orientation of 4-trans-2-(pyrid-4-yl-vinyl)benzoic acid (PVBA) trimers on Pd(111) using scanning tunneling microscopy (STM) under ultrahigh vacuum (UHV). The image showed three different types of trimers, one of which does not follow predicted dimer orientations. This type of trimer displays 10° rotations of each molecule in clockwise or counterclockwise directions. Calculations of adsorbate-substrate energy and hydrogen bonding energy revealed that the rotations are a result of competition between adsorbate-adsorbate and adsorbate-substrate interactions.  相似文献   

8.
The conformational study on N-acetyl- N'-methylamide of l-lactic acid (Ac-Lac-NHMe, the Lac dipeptide) is carried out using ab initio HF and density functional methods with the self-consistent reaction field method to explore its backbone conformational preferences and cis-trans isomerization for the depsipeptide with an ester bond in the gas phase and in solution. In the gas phase and in chloroform, the conformation tB with a trans depsipeptide bond is most preferred for the Lac dipeptide, whose backbone torsion angles are phi approximately -150 degrees and psi approximately -5 degrees , juxtaposed to those of the 3 10-helical structure. The larger shift in phi is brought to reduce the repulsion between the two carbonyl carbons of the acetyl and NHMe groups. However, the polyproline II-like tF conformation becomes more populated and the relative stability of conformation tB decreases significantly as the solvent polarity increases. This may be ascribed to weakening a C(5) hydrogen bond between the depsipeptidyl oxygen and the carboxyl amide hydrogen that plays a role in stabilizing the conformation tB in the gas phase and in chloroform. The cis populations about the depsipeptide bond are nearly negligible in the gas phase and in solution. The rotational barriers to the cis-trans isomerization of the depsipeptide bond for the Lac dipeptide are calculated to be about 11 kcal/mol, which is about half of those for the Ala dipeptide, although they increase somewhat with the increase of solvent polarity. The cis-trans isomerization of the depsipeptide bond proceeds through either clockwise or anticlockwise rotations with torsion angles of about +90 degrees or -90 degrees , respectively, in the gas phase and in solution, whereas it has been known that the isomerization proceeds through only the clockwise rotation for alanyl and prolyl peptide bonds. The pertinent distances between the depsipeptidyl oxygen and the carboxyl amide hydrogen can describe the role of this hydrogen bond in stabilizing the transition state structures in the gas phase and in solution.  相似文献   

9.
A dramatic multilayer substrate relaxation is observed for the (square root 19 x square root 19)-13CO adlayer phase on a Pt(111) electrode by surface X-ray scattering. Within the (square root 19 x square root 19) unit cell, a vertical expansion of 0.28 A was determined for the Pt atoms under near-top-site CO molecules, whereas only 0.04 A was found under near-bridge-site CO molecules. The lateral displacements involve small rotations toward more symmetric bonding. Both the expansions and rotations extend into the bulk with a decay length of 1.8 Pt layers. This nonuniform layer expansion, hitherto unseen, appears to be a manifestation of the differential stress induced by CO adsorption at different sites.  相似文献   

10.
《中国化学》2018,36(7):605-611
Material surfaces can induce cell responses such as contact guidance, yet little attention has been paid to further cell orientation. Herein, we report an interesting phenomenon of cell orientation beyond the classic contact guidance on a stripe‐like micropattern with cell‐adhesive arginine‐ glycine‐aspartate (RGD) peptides on a nonfouling background decorated by poly(ethylene glycol) (PEG). Such a micropattern with cell adhesion contrast led to significant contact guidance after cell seeding. What is more, the localized and elongated cells were found to be further orientated out of the adhesive stripes, and even an anticlockwise rotation was observed for rat mesenchymal stem cells (rMSCs). The left‐right asymmetry of rMSCs stood only in statistics, for we observed all cases including clockwise orientation, anticlockwise orientation or just keeping the orientation of previous contact guidance. We further found that human foreskin fibroblasts (HFFs) preferred a clockwise rotation, while human mesenchymal stem cells (hMSCs) and human umbilical vascular endothelial cells (HUVECs) exhibited no significant preference to either direction, which indicated that the left‐right symmetry or asymmetry was cell‐type dependent. The present report has partially confirmed the cell chirality and revealed its complexity, calling for further careful and comprehensive investigation of the challenging topic of cell chirality on material surfaces.  相似文献   

11.
系统地研究了在由1-(2,4,-三硝基)芴基-2,6-二甲基苯胺(DMTNF),4-(二乙氨基)-苯甲醛-1,1-二苯基腙(DEH)和Y晶型氧钛酞菁(Y-TiOPc)或非金属酞菁(H2Pc)构成的单层结构有机光导体的性能,考查了电荷产生材料(CGM)浓度、电场强度和CGM的种类对光导体静电照相性能的影响.研究结果表明,光导体的量子收率和感光度与CGM浓度有很大关系,正充电时随CGM的浓度的增加而增加,负充电时随CGM的浓度增加而降低.两种光导体在近红外光谱区表现出良好的光敏性,适合LD扫描成像.Y-TiOPc光导体的最高峰在80 nm处,半衰曝光量为0.588 μJ/cm2(正充电),0.828 μJ/cm2(负充电);H2Pc光导体正充电最高峰在800 nm处,半衰曝光量为1.50 μJ/cm2,负充电最高峰在820 nm处,半衰曝光量为1.9 μJ/cm2.  相似文献   

12.
The optical rotation of (S)-(-)-alpha-methylbenzylamine at 589 nm has been measured in 39 different solvents at five different concentrations: 0.25, 0.50, 1.00, 2.00, and 3.00 M. A correlation of the intrinsic rotations (i.e., extrapolation of specific rotations to zero concentration) with Kamlet's and Taft's solvent parameters (alpha, beta, and pi) is established. The polarity/polarizability, pi, and solvent acidity, alpha, terms are found to have a greater effect upon the optical rotation than the basicity of the solvent, beta. The specific rotation for (S)-(-)-alpha-methylbenzylamine has been calculated with Gaussian03 using a PCM model (B3LYP aug-cc-pVDZ) for all 39 solvated systems. Comparisons between the experimental and calculated values show the importance of hydrogen bonding on specific rotation.  相似文献   

13.
Although 99mTc and 51Cr have been used for lymphocyte labeling, these radionuclides have several disadvantages for study on immunological behaviour of lymphocyte; very high rate elution and low labeling efficiency for both radionuclides, and short half life for 99mTc. Indium-111 has quite suitable physical properties for clinical nuclear medicine, i.e. desirable photon energy (247,173 keV) and 2.8 day half life. 111In-oxine is lipid soluble and is known to pass through the cell membrane and attaches firmly to cytoplasmic component of the cell. On the other hand, 3H-thymidine is well known substance which incorporated to nucleic acid in the cell. In this study, distribution patterns of 111In-oxine and/or 3H-thymidine labeled lymphocyte in C3H/He mice were examined and the suitability of 111In-oxine labeled lymphocyte for radionuclide imaging in vivo was discussed. Thirty minutes after intravenous injection of 3H and/or 111In labeled lymphocyte, about 12% of lymphocyte were found in the lungs and rest of them were distributed mainly in the blood, kidneys and liver. After 24 hours the activity in the lung decreased markedly and the activity in the liver and kidneys increased with time. Between lymphocyte labeled with 111In-oxine and 3H-thymidine, there is not so much differences in terms of distribution patterns. From this study, it is concluded that the 111In-oxine labeled lymphocyte distributes in the same way as 3H labeled one, in spite of different labeling sites. This 111In-oxine labeling method can be used as a useful tool of radionuclide imaging in kinetic studies of lymphocyte in vivo.  相似文献   

14.
The reaction of calix[4]arene monoalkyl ethers with 1(S)-camphor-10-sulfonyl chloride yields 1,2- or 1,3-alkoxy-1(S)-camphorsulfonyloxycalixarenes depending on the nature of the base used. In the presence of triethylamine only 1,3-substituted derivatives are formed. Two diastereomers of the 1,2-substituted calixarenes with clockwise and anticlockwise arrangement of alkyl and camphorsulfonyl groups at the narrow rim are formed in the presence of sodium hydride or potassium carbonate. Due to chiral induction of the camphorsulfonyl group the 1,2-substitution is diastereoselective. The ratio of diastereomers formed is dependent on the alkyl groups at the calixarene narrow rim.  相似文献   

15.
The concerted use of ab initio time-dependent density functional theory (TDDFT) calculations of transparent spectral region optical rotation and of circular dichroism has recently become practicable, permitting the concerted use of transparent spectral region optical rotation and circular dichroism in determining the absolute configurations of chiral molecules. Here, we report concerted TDDFT calculations of the transparent spectral region specific rotations and of the circular dichroism spectra originating in n --> pi C=O group excitations of four bicyclo[3.3.1]nonane diones, 1-4. Comparison to experiment yields absolute configurations for 1-4. For each dione, specific rotations and circular dichroism spectra give identical absolute configurations. Our results are consistent with previous work, with the exception of the Octant Rule-derived absolute configuration of the 2,9-dione.  相似文献   

16.
申嗪霉素在辣椒及土壤中残留动态的研究   总被引:1,自引:0,他引:1  
采用高效液相色谱法测定1%申嗪霉素悬浮剂在辣椒及土壤中的残留动态和最终残留量。辣椒和土壤中申嗪霉素的半衰期分别为2.80~3.63d和2.83~4.62d,检出限分别为0.02mg/kg和0.01mg/kg,回收率分别为80.1%~102.5%和82.0%~97.1%,相对标准偏差分别为7.6%~9.4%和6.3%~8.4%。  相似文献   

17.
Homochirality in life has always been a driving force in scientific research and natural exploration. It has not been satisfactorily explained, and systematic investigations are necessary. This paper reported a homochiral expression of proteins dependent on the stirring direction of growing media. By controlling the stirring direction clockwise (CW) and anticlockwise (ACW) of the culture medium, proteins with distinct secondary structures were obtained, and D-amino acid may be included in the protein cultur...  相似文献   

18.
Coupled-cluster and density-functional methods have been used to determine specific rotations and electronic circular dichroism (ECD) rotational strengths for (S)-2-chloropropionitrile. Coupled-cluster specific rotations using both the length- and velocity-gauge representations of the electric-dipole operator, computed with basis sets of triple-zeta quality containing up to 326 functions, compare very well with recently reported gas-phase cavity-ring-down polarimetry data. ECD rotational strengths for the six lowest-lying excited states are found to vary in sign, and the second excited state, which has a larger rotational strength than the first by a factor of 4, was found to yield a much larger contribution (by a factor of 10) to the overall negative specific rotation observed both experimentally and theoretically. However, both valence and Rydberg states appear to make substantial contributions to the total rotation, often of opposite sign from the converged/linear-response result. Furthermore, the sum-over-states approach was found to be inadequate for reproducing the specific rotations derived from the linear-response approach, even when 100 excited states (well beyond the estimated ionization limit) were included in the summation. Density-functional specific rotations using the B3LYP functional with basis sets of quadruple-zeta quality containing up to 588 functions are found to be too large compared to experiment by approximately a factor of 2. This error appears to be related to both the underestimation of the electronic excitation energies, as well as concomitant overestimation of the corresponding ECD rotational strengths. Although earlier studies reported good agreement between density-functional specific rotations and experiment when electric-field-dependent functions were used in conjunction with a double-zeta-quality basis set, the results reported here, which are near the basis-set limit, suggest that this agreement may be fortuitous.  相似文献   

19.
The interaction of low-energy electrons with multilayers of SiCl(4) adsorbed on Si(111) leads to production and desorption of Cl((2)P(32)), Cl((2)P(12)), Si, and SiCl. Resonant structure in the yield versus incident electron energy (E(i)) between 6 and 12 eV was seen in all neutral channels and assigned to dissociative electron attachment (DEA), unimolecular decay of excited products produced via autodetachment and direct dissociation. These processes yield Cl((2)P(32)) and Cl((2)P(12)) with nonthermal kinetic energies of 425 and 608 meV, respectively. The Cl((2)P(12)) is produced solely at the vacuum surface interface, whereas the formation of Cl((2)P(32)) likely involves subsurface dissociation, off-normal trajectories, and collisions with neighbors. Structure in the Cl((2)P(32)) yield near 14 and 25 eV can originate from excitation of electrons in the 2e, 7t(2) and 6t(2), 6a(1) levels, respectively. Although the 14 eV feature was not present in the Cl((2)P(12)) yield, the broad 25 eV feature, which involves complex Auger filling of holes in the 6t(2) and 6a(1) levels of SiCl(4), is observed. Direct ionization, exciton decay, and DEA from secondary electron scattering all occur at E(i)>14 eV. Si and SiCl were detected via nonresonant ionization of SiCl(x) precursors that are produced via the same states and mechanisms that yield Cl. The Si retains the kinetic energy profile of the desorbed precursors.  相似文献   

20.
The electrocatalytic oxidation of ammonia on Pt(111) and Pt(100) has been studied using voltammetry, chronoamperometry, and in situ infrared spectroscopy. The oxidative adsorption of ammonia results in the formation of NH(x) (x = 0-2) adsorbates. On Pt(111), ammonia oxidation occurs in the double-layer region and results in the formation of NH and, possibly, N adsorbates. The experimental current transients show a hyperbolic decay (t(-1)), which indicates strong lateral (repulsive) interactions between the (reacting) species. On Pt(100), the NH(2) adsorbed species is the stable intermediate of ammonia oxidation. Stabilization of the NH and NH(2) fragments on Pt(111) and Pt(100), respectively, is in an interesting agreement with recent theoretical predictions. The Pt(111) surface shows extremely low activity in ammonia oxidation to dinitrogen, thus indicating that neither NH nor N (strongly) adsorbed species are active in dinitrogen production. Neither nitrous oxide nor nitric oxide is the product of ammonia oxidation on Pt(111) at potentials up to 0.9 V, as deduced from the in situ infrared spectroscopy measurements. The Pt(100) surface is highly active in dinitrogen production. This process is characterized by a Tafel slope of 30 mV decade(-1), which is explained by a rate-determining dimerization of NH(2) fragments followed by a fast decay of the resulting surface-bound hydrazine to dinitrogen. Therefore, the high activity of the Pt(100) surface for ammonia oxidation to dinitrogen is likely to be related to its ability to stabilize the NH(2) adsorbate.  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