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 140 毫秒
1.
We present quantitative measurements of the photoassociation of cesium molecules inside a far-detuned optical dipole trap. A model of the trap-depletion dynamics is derived which allows us to extract absolute photoassociation rate coefficients for the initial single-photon photoassociation step from measured trap-loss spectra. The sensitivity of this approach is demonstrated by measuring the Franck–Condon modulation of the weak photoassociation transitions into the low-vibrational levels of the outer well of the Og- state that correlates to the 6s+6p3/2 asymptote. The measurements are compared to theoretical predictions. In a magneto-optical trap, these transitions have previously only been observed indirectly through ionization of ground-state molecules. PACS 34.50.Rk; 33.70.Ca; 32.80.Pj  相似文献   

2.
The vibrational structure of the absorption spectra of the first n*–electron transitions of the HNO and DNO molecules is calculated in the Franck–Condon approximation. A structural model of the molecules in the excited electronic state is constructed on the basis of correlations and with the aid of a method of hybrid atomic orbitals. Evaluation of the influence of deuterium substitution on the intensities of the vibrational components upon electronic excitation is made. A comparison of the experimental and theoretical absorption spectra calculated for different models of the molecules is carried out.  相似文献   

3.
4.
We have traced the change in the complex structure of the intense 0–0 band of the electronic S 2 S 0 transition in the quasi-line absorption spectra (n-octane, 77 K) for individual types of impurity centers in the series of compounds tetraazaporphin–tetraazachlorin–di(tert-butylbenzo)barrelene-substituted tetraazachlorin. It has been concluded that the reason for the appearance of this structure is the interaction of the pure electronic S 2 state with the vibronic S 1 states rather than the Franck–Condon manifestation of low-frequency vibrations. Attempts to detect the short-wave S 2 S 0 fluorescence for all three compounds were unsuccessful. The reasons for the difference from the case of bacteriochlorin, in which such a fluorescence was observed earlier, are discussed.  相似文献   

5.
The luminescence spectra of solid xenon excited by a volume discharge are studied for the first time. The majority of the energy is emitted in the VUV spectral region due to the radiative decay of excitons. A specific feature of the spectra in the UV and visible regions is the complete absence of lines belonging to neutral excited particles (one-center Xe* and two-center Xe 2 * excitons) and to diatomic Xe 2 + ions. Numerous lines of the Xe+ ion are shifted by 0.1–0.3 nm with respect to their positions in the gas; the larger the shift, the stronger the distortion of the line shape. Based on the data obtained, a mechanism by which solid xenon is excited and ionized by fast drifting electrons is proposed.  相似文献   

6.
Temperature quenchings of narrow-line and broad-band emissions are pictured through the quantum-mechanical single-configurational-coordinate model. The model is taken in the thermal-Condon approximation with the overlap integrals evaluated through the Manneback recursion formulas. The multiple activation energies m?ωυ to the initial vibrational states produce an increase in the non-radiative rate poorly described, in general, by a single activation energy. The model applies to energy parabolas with large Franck- Condon offset and a consequent crossover and to energy parabolas with small Franck- Condon offset and no relevant crossover. For large offset, the model gives approximately Mott's single-activation-energy rate AM exp(-EX/kT) for upward transitions but faster rates poorly described by a single activation energy for downward transitions. For small offset, the model gives approximately Kiel's multiphonon-emission rate AK?p[1+〈mυ]p for downward transitions. A numerical matrix method is described which can handle all cases and which explicitly exhibits the multiple activation energies m?ωυ in every case. This method is used to work out examples of the various types of quenchings which can occur: a fast bottom crossover, an outside crossover, small-offset multiphonon emission, a tunneling crossover, and two-step quenchings through a higher offset state.  相似文献   

7.
8.
Intermolecular 129Xe–1H nuclear Overhauser effects and 129Xe longitudinal relaxation time measurements were used to demonstrate that the dipole–dipole coupling is the dominant relaxation mechanism for 129Xe in water, at room temperature. 129Xe–1H cross-relaxation rates were derived to be ςXeH 3.2 ± 0.3 × 10−3 s−1, independent of xenon pressure (in the range of 1–10 bar) and of the presence of oxygen. Corresponding xenon–proton internuclear distances were calculated to be 2.69 ± 0.12 Å. Using the magnitude of the dipole–dipole coupling and the spin density ratio between dissolved xenon and bulk water, it is estimated that 129Xe–1H spin polarization-induced nuclear Overhauser effects would yield little net proton signal enhancement in water.  相似文献   

9.
The 40Ca(p,p′ α) reaction has been studied at an incident proton energy Ep = 99.5 MeV for proton laboratory scattering angles Θplab = 17°, 23° and 27°. Emission of α particles coincident with the scattered proton has been measured for an angular range Θα 0° − 180° relative to the recoil axis. A multipole decomposition for the α0-decay channel to the 36Ar ground state has been performed from the angular-correlation functions. The energy distribution of the dominating E2 strength deduced in the excitation energy range Ex = 11–21 MeV agrees reasonably well with the results from electron and α-induced α0-decay investigations. The exhaustion of the E2 energy-weighted sum rule in this channel up to an energy of 17 MeV is 16.1(4.0)%, in accord with the study of the (α, α′ α0) reaction. However, this value is twice what is found in the (e,e′ α0) experiment in the same energy region. Thus, the puzzling discrepancy in the E2 strengths derived from electromagnetic and hadronic probes remains unsolved.  相似文献   

10.
Radiative parameters for vibrational bands (0 ≤ v′ ≤ 30, 0 ≤ v′′ ≤ 59) of the A1?u+ X1 ?g+ {A^1}\sum\nolimits_u^{+} {{X^1}} \sum\nolimits_g^{+} {} -electronic transition of a cesium dimer are calculated. These include vibronic band wavenumbers, Franck–Condon factors, Einstein coefficients for spontaneous emission, absorption band oscillator strengths, and radiative lifetimes for vibrational levels of the excited electronic state. Vibrational energies and wave functions for the ground and excited electronic states were found by a numerical solution of the radial wave equation based on potential curves plotted during the course of the study.  相似文献   

11.
It is shown that in the negative glow of a glow discharge in mixtures of neon with xenon (xenon contents of 5.5·10–3 to 2% of the total pressure P of the mixture) with P=15 torr and discharge currents i=40 mA (cathode diameter 30 mm), the 7s2P1/2, 3/2 levels of the Xe+ ion are excited by the charge exchange reaction Xe m + Ne+ Xe+* + Ne.The effective cross section Q3/2 for excitation of the 7s2P3/2 level is estimated to lie between 10–14 and 10–15 cm2 and Q3/26.5 × Q1/2, where Q1/2 is the effective cross section for excitation of the 7s2P1/2 level.Translated from Izvestiya Vysshikh Uchebnykh Zavedenii Fizika, No. 3, pp. 96–99, March, 1979.  相似文献   

12.
The interaction of xenon with different proteins in aqueous solution is investigated by 129Xe NMR spectroscopy. Chemical shifts are measured in horse metmyoglobin, hen egg white lysozyme, and horse cytochrome c solutions as a function of xenon concentration. In these systems, xenon is in fast exchange between all possible environments. The results suggest that nonspecific interactions exist between xenon and the protein exteriors and the data are analyzed in term of parameters which characterize the protein surfaces. The experimental data for horse metmyoglobin are interpreted using a model in which xenon forms a 1:1 complex with the protein and the chemical shift of the complexed xenon is reported (Locci et al., Keystone Symposia “Frontiers of NMR in Molecular Biology VI”, Jan. 9–15, 1999, Breckenridge, CO, Abstract E216, p. 53; Locci et al., XeMAT 2000 “Optical Polarization and Xenon NMR of Materials”, June 28–30, 2000, Sestri Levante, Italy, p. 46).  相似文献   

13.
In the present paper, an abrupt heterojunction photodetector based on Hg1 − xCdxTe (MCT) has been simulated theoretically for mid-infrared applications. A semi-analytical simulation of the device has been carried out in order to study the performance ratings of the photodetector for operation at room temperature. The energy band diagram, carrier concentration, electric field profile, dark current, resistance–area product, quantum efficiency and detectivity have been calculated and optimized as a function of different parameters such as device thickness, applied reverse voltage and operating wavelength. The effect of energy band offsets in conduction and valance band on the transportation of minority carriers has been studied. The influences of doping concentration, electron affinity gradient and the pn junction position within heterostructure on potential barrier have been analyzed. The optical characterization has been carried out in respect of quantum efficiency, and detectivity of the heterojunction photodetector. In present model the Johnson–Nyquist and shot noise has been considered in calculation of detectivity. The simulated results has been compared and contrasted with the available experimental results. Results of our analytical-cum-simulation study reveal that under suitable biasing condition, the photodetector offers a dark current, ID ≈ 6.5 × 10−12 A, a zero-bias resistance–area product, R0A ≈ 11.3 Ω m2, quantum efficiency, η ≈ 78%, NEP = 2 × 10−12 W Hz1/2 and detectivity D* ≈ 4.7 × 1010 mHz1/2/W.  相似文献   

14.
The charge state dependence of positron lifetime and trapping at divacancy (V2) in Si doped with phosphorus or boron has been studied after 15 McV electron irradiation up to a fluence of 8.0×1017 e/cm2. The positron trapping cross sections for V 2 2– , V 2 and V 2 0 at 300 K were about 6×10–14, 3×10–14 and 0.1–3×10–14 cm2, respectively. For V 2 + , however, no positron trapping was observed. The marked difference in the cross sections comes from Coulomb interaction between the positron and the charged divacancy. The trapping rates for V 2 0 and V 2 2– have been found to increase with decreasing temperature in the temperature range of 10–300 K. These results are well interpreted by a two-stage trapping model having shallow levels with energy of 9 meV (V 2 0 ) and 21 meV (V 2 2– ). The appearance of a shallow level for V 2 0 can not be explained by a conventional Rydberg state model. The lifetime (290–300 ps) in V 2 0 is nearly constant in the temperature range from 10 to 300 K, while that in V 2 2– increases from 260 ps at 10 K to 320 ps at 300 K. The lifetime (260 ps) in V 2 2– is shorter than that in V 2 0 at low temperature, which is due to the excess electron density in V 2 2– . At high temperature, however, the longer lifetime of V 2 2– than that of V 2 0 is attributed to lattice relaxation around V 2 2– .  相似文献   

15.
By employing the dissociation energy and the equilibrium bond length for a diatomic molecule as explicit parameters, we generate an improved form of the Williams-Poulios potential energy model. It is found that the negative Williams-Poulios potential model is equivalent to the Manning-Rosen potential model for diatomic molecules. We observe that the Manning-Rosen potential is superior to the Morse potential in reproducing the interaction potential energy curves for the \({{a}^{3} \Sigma_{{\rm u}}^{+}}\) state of the 6Li2 molecule and the \({{X}^{1} \sum^{+}}\) state of the SiF+ molecule.  相似文献   

16.
We present measurements of the transverse (T 2 –1 ) and longitudinal (T 1 –1 ) spin relaxation rates of muonium (Mu) atoms in solid natural xenon (n-Xe) as well as pure136Xe (which has no nuclear moments). The temperature dependences ofT 2 –1 andT 1 –1 in natural Xe belowT 115 K demonstrate the quantum character of Mu diffusion governed by one-phonon interactions. Taking into account both the polaron effect (PE) and the effect of fluctuational preparation of the barrier (FPB) makes it possible to consistently describe Mu diffusion in Xe. Mu spin relaxation in136Xe at high temperatures is not due to nuclear hyperfine (NHF) interactions.  相似文献   

17.
A wide adiabatic study has been performed for numerous electronic states of CaLi+ molecular ion. The adiabatic potential energy curves and their spectroscopic constants (Re, De, ωe and Te) have been calculated using an ab initio approach including a nonempirical pseudo-potential for the Ca and Li cores with the core polarisation potentials operator through full configuration interaction (FCI). Thereafter, the energies of vibrational levels and their spacing for all these states have been reported. In addition, the electric dipole moment curves have been investigated for the (1-19) Σ, (1-12) Π and (1-8) Δ electric states. Moreover it lets us check the extreme transition dipole moments (TDM). These behaviours of TDM are more accustomed to estimate the radiative lifetimes for all vibrational levels in 21Σ+ and 31Σ+ states. Also, the bound-bound and the bound-free contribution have been calculated precisely and by employing a Franck–Condon (FC) approximation.  相似文献   

18.
The adsorption of xenon has been studied with UV photoemission (UPS), flash desorption (TDS) and work function measurements on differently conditioned Ru(0001) surfaces at 100 K and at pressures up to 3 × 10?5 Torr. Low energy electron diffraction (LEED) and Auger electron spectroscopy (AES) served to ascertain the surface perfectness. On a perfect Ru(0001) surface only one Xe adsorption state is observed, which is characterized byXe5p32,12 electron binding energies of 5.40 and 6.65 eV, an adsorption energy of Ead≈ 5 kcal/mole and dipole moment of μ'T ≈ 0.25 D. On a stepped-kinked Ru(0001) surface, the terrace-width, the step-height and step-orientation of which are well characterized with LEED, however, two coexisting xenon adsorption states are distinguishable by an unprecedented separation inXe 5p32,12 electron binding energies of 800 meV, by their different UPS intensities and line shapes, by their difference in adsorption energy ofΔEad ≈ 3 kcal/mole and finally by their strongly deviating dipole moments of μS = 1.0 D and μT = 0.34 D. The two xenon states (which are also observed on a slightly sputtered surface) are identified as corresponding to xenon atoms being adsorbed at step and terrace sites, respectively. Their relative concentrations as deduced from the UPS intensities quantitatively correlate with the abundance of step and terrace sites of the ideal TLK surface structure model as derived from LEED. Furthermore, ledge-sites and kink-sites are distinguishable via Ead. The Ead heterogeneity on the stepped-kinked Ru(0001) surface is interpreted in terms of different coordination and/or different charge-transfer-bonding at the various surface sites. The enormous increase in Xe 5p electron binding energy of 0.8 eV for Xe atoms at step sites is interpreted as a pure surface dipole potential shift. —The observed effects suggest selective xenon adsorption as a tool for local surface structure determination.  相似文献   

19.
Absorption spectra of the gases SiH4, NH3, C2H2 and of SiH4/Ar and SiH4/B2H6 mixtures have been measured in the spectral range of the CO2 laser from 9.2 to 10.8 µm. In agreement with literature, silane shows the highest absorption (absorption coefficient = 3.3 × 10–2 Pa–1 m–1). The deviation of the measured absorption behaviour of silane from literature, as far as the pressure dependence is concerned, can be explained by the enhanced spectral energy density in our experiment. This is confirmed by a rate-equation model involving the basic mechanisms of V-V and V-T energy transfer between vibrationally excited silane molecules. In contrast to silane, the absorption coefficient of NH3 at the 10P(20) laser line is 4.5 × 10–4 Pa–1 m–1 atp = 20 kPa and has its maximum of 4.5 × 10–3 Pa–1 m–1 at the 10R(6) laser line. For C2H2 and B2H6, is even less ( 2.1 Ò 10–5 Pa–1 m–1 for C2H2).  相似文献   

20.
Lomaev  M. I.  Rybka  D. V.  Tarasenko  V. F.  Lipatov  E. I.  Krishnan  M.  Thompson  J.  Parks  D. 《Russian Physics Journal》2004,47(1):89-93
This paper presents the results of an experimental study of the sensitivity of a detector made of natural diamond of the IIa type as a function of the power density of the radiation of a pulsed xenon lamp. It has been shown that the sensitivity of the detector depends on the power density of the xenon lamp radiation and equals about (1–6)·10–4 A/W. A monotonic decrease in sensitivity of the detector during a radiation pulse has been detected. The minimum resistance of the diamond crystal with an irradiated face of area 3 mm2, placed 7 mm away from the axis of the xenon lamp, was 300 , which corresponds to a specific conductivity of 2.2 –1·m–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