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 509 毫秒
1.
TheY2Σ+–X2Πinear-infrared electronic transition of CuO was observed at high resolution for the first time. The spectrum was recorded with the Fourier transform spectrometer associated with the McMath–Pierce Solar Telescope at Kitt Peak. The excited CuO molecules were produced in a low pressure copper hollow cathode sputter with a slow flow of oxygen. Constants for theY2Σ+states of CuO are:T0= 7715.47765(54) cm−1,B= 0.4735780(28) cm−1,D= 0.822(12) × 10−6cm−1,H= 0.46(10) × 10−10cm−1, γ = −0.089587(42) cm−1, γD= 0.1272(79) × 10−6cm−1,bF= 0.12347(22) cm−1, andc= 0.0550(74) cm−1. ImprovedX2Πiconstants are also presented.  相似文献   

2.
The pure rotational spectrum of CH2F2 was recorded in the 20–100 cm−1 spectral range and analyzed to obtain rotation and centrifugal distortion constants. Analysis of the data yielded rotation constants: A = 1.6392173 ± 0.0000015, B = 0.3537342 ± 0.00000033, C = 0.3085387 ± 0.00000027, τaaaa = −(7.64 ± 0.46) × 10−5, τbbbb = −(2.076 ± 0.016) × 10−6, τcccc = −(9.29 ± 0.12) × 10−7, T1 = (4.89 ± 0.20) × 10−6, and T2 = −(1.281 ± 0.016) × 10−6cm−1.  相似文献   

3.
The ν3±1 perpendicular band of 14NF3 ( cm−1) has been studied with a resolution of 2.5 × 10−3 cm−1, and 3682 infrared (IR) transitions (Jmax=55, Kmax=45) have been assigned. These transitions were complemented by 183 millimeterwave (MMW) rotational lines (Jmax=25, Kmax=19) in the 150–550 GHz region (precision 50–100 kHz). The kl=+1 level reveals a strong A1/A2 splitting due to the l(2,2) rotational interaction (q=−4.05 × 10−3 cm−1) while the kl=−2 and +4 levels exhibit small A1/A2 splittings due to l(2,−4) and l(0,6) rotational interactions. All these splittings were observed by both experimental methods. Assuming the v3=1 vibrational state as isolated, a Hamiltonian model of interactions in the D reduction, with l(2,−1) rotational interaction (r=−1.96 × 10−4 cm−1) added, accounted for the observations. A set of 26 molecular constants reproduced the IR observations with σIR=0.175 × 10−3 cm−1 and the MMW data with σMMW=134 kHz. The Q reduction was also performed and found of comparable quality while the QD reduction behaved poorly. This may be explained by a predicted Coriolis interaction between v3=1 and v1=1 (A1, 1032.001 cm−1) which induces a slow convergence of the Hamiltonian in the QD reduction but has no major influence on the other reductions. The experimental equilibrium structure could be calculated as: re(N–F)=1.3676 Å and (FNF)=101.84°.  相似文献   

4.
The 2ν3(A1) band of 12CD3F near 5.06 μm has been recorded with a resolution of 20–24 × 10−3 cm−1. The value of the parameter (αB − αA) for this band was found to be very small and, therefore, the K structure of the R(J) and P(J) manifolds was unresolved for J < 15 and only partially resolved for larger J values. The band was analyzed using standard techniques and values for the following constants determined: ν0 = 1977.178(3) cm−1, B″ = 0.68216(9) cm−1, DJ = 1.10(30) × 10−6 cm−1, αB = (B″ − B′) = 3.086(7) × 10−3 cm−1, and βJ = (DJDJ) = −3.24(11) × 10−7 cm−1. A value of αA = (A″ − A′) = 2.90(5) × 10−3 cm−1 has been obtained through band contour simulations of the R(J) and P(J) multiplets.  相似文献   

5.
Molecular constants for the E0+(3P2) and 1(3P2) ion-pair states of ICl vapor have been determined using sequential two-photon polarization-labeling spectroscopy. The two states are coupled by a heterogeneous perturbation which is analyzed in some detail for low-lying vibrational levels of 1(3P2). The I35Cl potential constants for the 1(3P2) state and the rotation-vibration constants for the set of f sublevels—i.e., the constants unaffected by coupling with the E state—are (in cm−1) 1(3P2): Y0,0= 39103.814(32), Y1,0= 170.213(15), Y2,0= −0.4528(22), Y3,0= −7.0(12) × 10−4, Y4,0= −1.48(24) × 10−5 and Y5,0= −6.6(19) × 10−8, Y(f)0,1= 5.6878(17) × 10−2 Y(f)1,1= −2.110(24) × 10−4, Y(f)2,1= −1.23(62) × 10−7, and Y(f)0,1= −3.08(22) × 10−8Likewise, the I35Cl constants determined for the E 0+(3P2) state are E 0+(3P2: Y0,0= 39054.38(61), Y1,0= 166.96(10), Y2,0 = −0.3995(42), Y0,1= 5.738(31) × 10−2, and Y1,1= −1.67(26) × 10−4Practical constraints in pumping the sequence E 0+B 0+ ← × 0+ restrict the analysis of the E state to levels v = 9–15. Given the long extrapolation to the equilibrium state the 3σ statistical uncertainties quoted for these constants should be treated with caution.  相似文献   

6.
Single-crystal ZnO has been hydrothermally grown with additional In2O3 in the solution. Schottky barrier contacts have been deposited by electron beam evaporation of Pd onto the face. Capacitance–voltage measurements have been performed to reveal the carrier concentration as a function of the In2O3 content in the solution, and secondary-ion mass spectrometry was used to measure the resulting In concentration in the samples. For an In2O3 content of 2×1019 cm−3, the average free electron concentration increased to 5×1018 cm−3 compared to 4×1017 cm−3 for the non-doped material. An increase of the In2O3 content to 4×1019 cm−3 leads to a measured carrier concentration of approximately 1×1019 cm−3; however, only up to a quarter of the incorporated In became electrically active. From thermal admittance spectroscopy measurements two prominent electronic levels are found, and compared with to the non-doped material case, the freeze-out of the shallow doping in the In-doped samples takes place at lower temperatures (below 80 K).  相似文献   

7.
The Ag2O–TiO2–SiO2 glasses were prepared by Ag+/Na+ ion-exchange method from Na2O–TiO2–SiO2 glasses at 380–450 °C below their glass transition temperatures (Tg), and their electrical conductivities were investigated as functions of TiO2 content and the ion-exchange ratio (Ag/(Ag+Na)). In a series of glasses 20R2xTiO2·(80−x)SiO2 with x=10, 20, 30 and 40 in mol%, the electrical conductivities at 200 °C of the fully ion-exchanged glasses of R=Ag were in the order of 10−5 or 10−4 S cm−1 and were 1 or 2 orders of magnitude higher than those of the initial glasses of R=Na. The glass of x=30 exhibited the highest increase of conductivity from 3.8×10−7 to 1.3×10−4 S cm−1 at 200 °C by Ag+/Na+ ion exchange among them. When the ion-exchange ratio was changed in 20R2O·30TiO2·50SiO2 system, the electrical conductivity at 200 °C exhibited a minimum value of 7.6×10−8 S cm−1 around Ag/(Ag+Na)=0.3 and increased steeply in the region of Ag/(Ag+Na)=0.5–1.0. When the ion-exchange temperature was changed from 450 to 400 °C, the conductivity of the ion-exchanged glass of x=30 decreased. The infrared spectroscopy measurement revealed that the ion-exchange temperature of 450 °C induced a structural change in the glass of x=30. The Tg of the fully ion-exchanged glass of x=30 was 498 °C. It was suggested that the incorporated silver ions changed the average coordination number of titanium ions to form higher ion-conducting pathway and resulted in high conductivity in the titanosilicate glasses.  相似文献   

8.
The electron impact behavior of CO adsorbed on was investigated. The desorption products observed were neutral CO, CO+, and O+. After massive electron impact residual carbon, C/W = 0.15, but not oxygen was also found, suggesting that energetic neutral O, not detected in a mass analyzer must also have been formed. Formation of β-CO, i.e., dissociated CO with C and O on the surface was not seen. The total disappearance cross section varies only slightly with coverage, ranging from 9 × 10 −18 cm2 at low to 5 × 10−18 cm2 at saturation (CO/W = 0.75). The cross section for CO+ formation varies from 4 × 10−22 cm2 at satura to 2 × 10−21 cm2 at low coverage. That for O+ formation is 1.4 × 10−22 cm2 at saturation and 2 × 10−21 cm2 Threshold energies are similar to those found previously [J.C. Lin and R. Gomer, Surf. Sci. 218 (1989) 406] for and CO/Cu1/W(110) which suggests similar mechanisms for product formation, with the exception of β-CO on clean W(110). It is argued that the absence or presence of β-CO in ESD hinges on its formation or absence in thermal desorption, since electron impact is likely to present the surface with vibrationally and rotationally activated CO in all cases; β-CO formation only occurs on surfaces which can dissociate such CO. It was also found that ESD of CO led to a work function increase of the remaining Pd1/W(110) surface of 500 meV, which could be annealed out only at 900 K. This is attributed to surface roughness, caused by recoil momentum of energetic desorbing entities.  相似文献   

9.
The structural properties of a 10 μm thick In-face InN film, grown on Al2O3 (0001) by radio-frequency plasma-assisted molecular beam epitaxy, were investigated by transmission electron microscopy and high resolution x-ray diffraction. Electron microscopy revealed the presence of threading dislocations of edge, screw and mixed type, and the absence of planar defects. The dislocation density near the InN/sapphire interface was 1.55×1010 cm−2, 4.82×108 cm−2 and 1.69×109 cm−2 for the edge, screw and mixed dislocation types, respectively. Towards the free surface of InN, the density of edge and mixed type dislocations decreased to 4.35×109 cm−2 and 1.20×109 cm−2, respectively, while the density of screw dislocations remained constant. Using x-ray diffraction, dislocations with screw component were found to be 1.2×109 cm−2, in good agreement with the electron microscopy results. Comparing electron microscopy results with x-ray diffraction ones, it is suggested that pure edge dislocations are neither completely randomly distributed nor completely piled up in grain boundaries within the InN film.  相似文献   

10.
This paper reports the spectral properties and energy levels of Cr3+:Sc2(MoO4)3 crystal. The crystal field strength Dq, Racah parameter B and C were calculated to be 1408 cm−1, 608 cm−1 and 3054 cm−1, respectively. The absorption cross sections σα of 4A24T1 and 4A24T2 transitions were 3.74×10−19 cm2 at 499 nm and 3.21×10−19 cm2 at 710 nm, respectively. The emission cross section σe was 375×10−20 cm2 at 880 nm. Cr3+:Sc2(MoO4)3 crystal has a broad emission band with a broad FWHM of 176 nm (2179 cm−1). Therefore, Cr3+:Sc2(MoO4)3 crystal may be regarded as a potential tunable laser gain medium.  相似文献   

11.
Nonlinear optical properties of Fe2O3 nanoparticles were investigated by the signal-beam Z-scan technique with Ar+ and Ne–He lasers. The largest reported effective nonlinear coefficient, n2=−8.07×10−7 cm2/W, was obtained. It is demonstrated that the nonlinear optical response originals from quantum confinement effect.  相似文献   

12.
Using a Fourier transform spectrometer, we have recorded the spectra of ozone in the region of 4600 cm−1, with a resolution of 0.008 cm−1. The strongest absorption in this region is due to the ν1+ ν2+ 3ν3band which is in Coriolis interaction with the ν2+ 4ν3band. We have been able to assign more than 1700 transitions for these two bands. To correctly reproduce the calculation of energy levels, it has been necessary to introduce the (320) state which strongly perturbs the (113) and (014) states through Coriolis- and Fermi-type resonances. Seventy transitions of the 3ν1+ 2ν2band have also been observed. The final fit on 926 energy levels withJmax= 50 andKmax= 16 gives RMS = 3.1 × 10−3cm−1and provides a satisfactory agreement of calculated and observed upper levels for most of the transitions. The following values for band centers are derived: ν01+ ν2+ 3ν3) = 4658.950 cm−1, ν0(3ν1+ 2ν2) = 4643.821 cm−1, and ν02+ 4ν3) = 4632.888 cm−1. Line intensities have been measured and fitted, leading to the determination of transition moment parameters for the two bands ν1+ ν2+ 3ν3and ν2+ 4ν3. Using these parameters we have obtained the following estimations for the integrated band intensities,SV1+ ν2+ 3ν3) = 8.84 × 10−22,SV2+ 4ν3) = 1.70 × 10−22, andSV(3ν1+ 2ν2) = 0.49 × 10−22cm−1/molecule cm−2at 296 K, which correspond to a cutoff of 10−26cm−1/molecule cm−2.  相似文献   

13.
Absorption spectra of C2H2 have been recorded between 50 and 1450 cm−1, with a resolution always better than 0.005 cm−1, using two different Fourier transform spectrometers. Analysis of the data provided two sets of results. First, the bending levels with Σt Vt(t = 4, 5) ≤ 2 were characterized by a coherent set of 34 parameters derived from the simultaneous analysis of 15 bands, performed using a matrix Hamiltonian. The following main parameters were obtained (in cm−1): ω40 = 608.985196(14), ω50 = 729.157564(10); B0 = 1.17664632(18), α4 = −1.353535(86) × 10−3, α5 = −2.232075(40) × 10−3; q40 = 5.24858(12) × 10−3, and q50 = 4.66044(12) × 10−3, with the errors (1σ) on the last quoted digit. Second, a more complete set of bending levels with Σt Vt ≤ 4, some of which have never previously been reported, and also including V2 = 1 have been fitted to 80 parameters. This simultaneous fit involved 43 bands and used the same full Hamiltonian matrix. Some perturbations which affect the higher excited levels are discussed.  相似文献   

14.
Molybdenum oxide films (MoO3) were deposited on glass and crystalline silicon substrates by sputtering of molybdenum target under various oxygen partial pressures in the range 8 × 10−5–8 × 10−4 mbar and at a fixed substrate temperature of 473 K employing dc magnetron sputtering technique. The influence of oxygen partial pressure on the composition stoichiometry, chemical binding configuration, crystallographic structure and electrical and optical properties was systematically studied. X-ray photoelectron spectra of the films formed at 8 × 10−5 mbar showed the presence of Mo6+ and Mo5+ oxidation states of MoO3 and MoO3−x. The films deposited at oxygen partial pressure of 2 × 10−4 mbar showed Mo6+ oxidation state indicating the films were nearly stoichiometric. It was also confirmed by the Fourier transform infrared spectroscopic studies. X-ray diffraction studies revealed that the films formed at oxygen partial pressure of 2 × 10−4 mbar showed the presence of (0 k 0) reflections indicated the layered structure of α-phase MoO3. The electrical conductivity of the films decreased from 3.6 × 10−5 to 1.6 × 10−6 Ω−1 cm−1, the optical band gap of the films increased from 2.93 to 3.26 eV and the refractive index increased from 2.02 to 2.13 with the increase of oxygen partial pressure from 8 × 10−5 to 8 × 10−4 mbar, respectively.  相似文献   

15.
Ultrasonic irradiation (22 kHz, Ar atmosphere) of Th(IV) β-diketonates Th(HFAA)4 and Th(DBM)4, where HFAA and DBM are hexafluoroacetylacetone and dibenzoylmethane respectively, causes them to decompose in hexadecane solutions, forming solid thorium compounds. The first-order rate constants for Th(IV) β-diketonate degradation were found to be (9.3±0.8)×10−3 for Th(HFAA)4 and (3.8±0.4)×10−3 min−1 for Th(DBM)4, (T=92°C, I=3 W cm−2). The rate of the sonochemical reaction increased with the rising β-diketonate volatility and decreased with the rising hydrocarbon solvent vapor pressure. Solid sonication products consisted of a mixture of thorium carbide ThC2 and Th(IV) β-diketonate partial degradation products. The average ThC2 particle size was estimated to be about 2 nm. ThC2 formation was attributed to the high-temperature reaction occurring within the cavitating bubble. The thorium β-diketonate partial degradation products formed in the liquid reaction zones surrounding the cavitating bubbles.  相似文献   

16.
To support planetary studies of the Venus atmosphere, we measured line strengths of the 2v3, v1+2v2+v3, and 4v2+v3 bands of the primary isotopologue of carbonyl sulfide (16O12C32S), whose band centers are located at 4101.387, 3937.427, and 4141.212 cm−1, respectively. For this, infrared absorption spectra in normal carbonyl sulfide (OCS) sample gas were recorded at an unapodized resolution of 0.0033 cm−1 at ambient room temperatures using a Bruker Fourier transform spectrometer (FTS) at the Jet Propulsion Laboratory. The FTS instrumental line shape (ILS) function was investigated, which revealed no significant instrumental line broadening or distortions. Various custom-made short cells and a multi-pass White cell were employed to achieve optical densities sufficient to observe the strong 2v3 and the weaker bands in the region. Gas sample impurities and the isotopic abundances were determined from mass spectrum analysis. Line strengths were retrieved spectrum by spectrum using a non-linear curve fitting algorithm adopting a standard Voigt line profile, from which Herman–Wallis factors were derived for the three bands. The band strengths of 2v3, v1+2v2+v3, and 4v2+v3 of 16O12C32S (normalized at 100% of isotopologue) are observed to be 6.315(13)×10−19, 1.570(2)×10−20, and 7.949(20)×10−21 cm−1/molecule cm−2, respectively, at 296 K. These results are compared with earlier measurements and the HITRAN 2004 database.  相似文献   

17.
Rashba polarization in HgCdTe inversion layers at large depletion charges   总被引:1,自引:0,他引:1  
The Rashba effect in metal–insulator–semiconductor (MIS) structures based on zero-gap HgCdTe is investigated experimentally and theoretically over a wide doping range NAND=3×1015–3×1018 cm−3. Increase of doping enlarges the magnitude of the effect at the same 2D concentration and strengthens a gate-voltage dependence of the Rashba splitting. The results demonstrate values of Rashba polarization as high as PR0.5 and a capability to control the Rashba effect strength at constant electron concentration.  相似文献   

18.
We report the observation of sharp plasmon and magnetoplasmon modes in ultra-low-density 2D electron systems. Well defined dispersions for the modes are observed at densities as low as 1.1×109 cm−2 and with excitation wave vectors as large as 1.2×105 cm−1. Interestingly, both modes are found to be more easily measured in low-density systems than in high-density systems. The strength of the light scattering cross-sections at low density suggests potential applications to the study of quantum phase transitions at large rs.  相似文献   

19.
High resolution Fourier transform spectra of deuterated hydrogen sulfide have been recorded in the region 2400-3000 cm−1. Rotational structures of the ν1 + ν2, ν2 + ν3 bands of D232S, of the ν3 and ν1 + ν2 bands of HD32S, and of the ν1 + ν2 band of HD34S were analyzed. Band centers and rotational, centrifugal distortion, and resonance parameters were obtained, which reproduce the initial values of the upper energy levels within a mean accuracy of 1.39 × 10−4 cm−1 for the states (110) and (011) of D232S, 1.61 × 10−4 cm−1 and 1.82 × 10−4 cm−1 for the states (001) and (110) of HD32S, and 2.09 × 10−4 cm−1 for the state (110) of HD34S, respectively.  相似文献   

20.
We examine the reach of a Beta-beam experiment with two detectors at carefully chosen baselines for exploring neutrino mass parameters. Locating the source at CERN, the two detectors and baselines are: (a) a 50 kton iron calorimeter (ICAL) at a baseline of around 7150 km which is roughly the magic baseline, e.g., ICAL@INO, and (b) a 50 kton Totally Active Scintillator Detector at a distance of 730 km, e.g., at Gran Sasso. We choose 8B and 8Li source ions with a boost factor γ of 650 for the magic baseline while for the closer detector we consider 18Ne and 6He ions with a range of Lorentz boosts. We find that the locations of the two detectors complement each other leading to an exceptional high sensitivity. With γ=650 for 8B/8Li and γ=575 for 18Ne/6He and total luminosity corresponding to 5×(1.1×1018) and 5×(2.9×1018) useful ion decays in neutrino and antineutrino modes respectively, we find that the two-detector set-up can probe maximal CP violation and establish the neutrino mass ordering if sin22θ13 is 1.4×10−4 and 2.7×10−4, respectively, or more. The sensitivity reach for sin22θ13 itself is 5.5×10−4. With a factor of 10 higher luminosity, the corresponding sin22θ13 reach of this set-up would be 1.8×10−5, 4.6×10−5 and 5.3×10−5 respectively for the above three performance indicators. CP violation can be discovered for 64% of the possible δCP values for sin22θ1310−3 (8×10−5), for the standard luminosity (10 times enhanced luminosity). Comparable physics performance can be achieved in a set-up where data from CERN to INO@ICAL is combined with that from CERN to the Boulby mine in United Kingdom, a baseline of 1050 km.  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