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 375 毫秒
1.
Polycrystalline perovskite La0.67Ca0.33MnO3 was synthesized by a sol–gel method. Its adiabatic temperature change ΔTad induced by a magnetic field change was measured directly. At 268 K, near its Curie temperature TC, ΔTad of La0.67Ca0.33MnO3 induced by a magnetic field change of 2.02 T reaches 2.4 K. The latent heat Q and magnetic entropy change −ΔSM induced by a magnetic field change were calculated from the temperature dependence of ΔTad and zero-field heat capacity Cp. The maximum values of Q and −ΔSM in La0.67Ca0.33MnO3 induced by a magnetic field change of 2.02 T are 1.85 J g−1 and 6.9 J kg−1 K−1, respectively. The former is larger than the phase transition latent heat of heating or cooling, which is about 1.70 J g−1.  相似文献   

2.
The apparent molar volumes, V of diglycine, triglycine and glycyl-l-leucine have been determined in water and in aqueous sodium acetate (0.5, 1.0, 2.0, and 4.0 mB) and magnesium acetate (0.5, 1.0, 1.5, and 2.0 mB) solutions at 298.15 K by the measurement of densities using vibrating-tube digital densimeter. The partial molar volumes, V2,m0 obtained from V have been used to calculate the partial molar volumes of transfer, ΔtrV2,m0 for these peptides from water to aqueous solutions of sodium acetate (SA) and magnesium acetate (MA) solutions. The hydration numbers, nH and volumetric interaction coefficients have also been calculated. The ΔtrV2,m0 data suggest that ion-charged/or peptide group interactions of peptides are stronger with MA in comparison to SA.  相似文献   

3.
Two-loop radiative mechanism, when combined with an U(1)L symmetry generated by LeLμLτ (=L′), is shown to provide an estimate of Δm2m2atm εme/mτ, where ε measures the U(1)L-breaking. Since Δm2atm 3.5×10−3 eV2, we find that Δm2 ε10−6 eV2, which will fall into the allowed region of the LOW solution to the solar neutrino problem for ε 0.1.  相似文献   

4.
A spherical harmonic moment analysis of the reactions Kp → Kπ+n and K+p → K+πΔ++ at 13 GeV/c demonstrates the existence of a broad K* state with mass in the vicinity of 1800 MeV and spin parity 3.  相似文献   

5.
It is known that the Korteweg–de Vries (KdV) equation is a geodesic flow of an L2 metric on the Bott–Virasoro group. This can also be interpreted as a flow on the space of projective connections on S1. The space of differential operators Δ(n)=∂n+u2n−2++un form the space of extended or generalized projective connections. If a projective connection is factorizable Δ(n)=(∂−((n+1)/2−1)p1)(∂+(n−1)/2pn) with respect to quasi primary fields pi’s, then these fields satisfy ∑i=1n((n+1)/2−i)pi=0. In this paper we discuss the factorization of projective connection in terms of affine connections. It is shown that the Burgers equation and derivative non-linear Schrödinger (DNLS) equation or the Kaup–Newell equation is the Euler–Arnold flow on the space of affine connections.  相似文献   

6.
The time-dependence of the decay rate of initially pure K0 into the final state (π+ππ0) has been studied in search for the decay kS0→π+ππ0. No evidence is found in a sample of 384 observed events. The ratio of the CP -violating KS0 amplitude and the KL0 amplitude is η+−0 = (0.13−0.20+0.17) + i(0.17−0.26+0.27); the ratio of the CP-conserving KS0 amplitude and the KL0 amplitude is < 0.4. The energy dependence of the K0→π+ππ0 matrix element is found to be a+−0 = −0.31 ± 0.03.  相似文献   

7.
The iodine complexes of pyridine, -, β-, γ- picolines, 2-aminopyridine and 2,2′-bipyridine in cyclohexane were studied by the constant activity method, their spectral characteristics and thermodynamic parameters were determined. The equilibrium constants, molar extinction coefficients of the charge-transfer band maxima, CT, at 25°C and heats of complexation, ΔH°, are respectively as follows:

1. (a) pyridine-iodine, 107 dm3mol−1 24200 dm3mol−1m−1 and 31 KJ/mol;

2. (b) -picoline-iodine, 172 dm3mol−1m−1, 49700 dm3mol−1m−1 and 33 KJ/mol;

3. (c) β-picoline-iodine, 243 dm3mol−1, 50300 dm3mol−1m−1 and 35 KJ/mol;

4. (d) γ-picoline-iodine, 342 dm3mol−1, 55900 dm3mol−1m−1 and 41 KJ/m;

5. (e) 2-aminopyridine-iodine, 427 dm3mol−1, 60800 dm3mol−1m−1 and 47 KJ/mol;

6. (f) 2,2′-bipyridine-iodine, 6.0 dm3mol−1,−, 21 KJ/mol.

The values of the blue shifted iodine band maxima of the above mentioned donors and their molar extinction coefficients are, 420 nm (1450 dm3mol−1m−1), 419 nm (1470 dm3mol−1m−1), 418 nm (1640 dm3mol−1m−1), 417 nm (1570 dm3mol−1m−1), 410 nm (1574 dm3mol−1m−1) and 452 nm (1370 dm3mol−1m−1) respectively. The oscillator strengths and transition dipole moments of the blue-shifted iodine bands are characterised.  相似文献   


8.
We present a calculation of the Coulomb final-state interaction effects in the decays KL → π±V. Specifically we investigate the charge asymmetry and the transverse polarization of the charged lepton. We consider the effects of possible ΔQ = −ΔS amplitudes and the q2 dependence of the form factors. The variation of the effects over the Dalitz plot is discussed and many numerical results are given.  相似文献   

9.
We comment on CP, T and CPT violation in the light of interesting new data from the CPLEAR and KTeV Collaborations on neutral kaon decay asymmetries. Other recent data from the CPLEAR experiment, constraining possible violations of CPT and the ΔSQ rule, exclude the possibility that the semileptonic-decay asymmetry AT measured by CPLEAR could be solely due to CPT violation, confirming that their data constitute direct evidence for T violation. The CP-violating asymmetry in KLee+ππ+ recently measured by the KTeV Collaboration does not by itself provide direct evidence for T violation, but we use it to place new bounds on CPT violation.  相似文献   

10.
《Physics letters. [Part B]》2002,550(3-4):147-153
Narrow structures in the range of a few MeV have been searched for in ppπ+ and ppπ invariant mass spectra (Mppπ+ and Mppπ) obtained from exclusive measurements of the ppppπ+π reaction at Tp=725,750 and 775 MeV using the PROMICE/WASA detector at CELSIUS. The selected reaction is particularly well suited for the search for dibaryon resonances decoupled from NN and/or NΔ. In the mass range 2020 MeV/c2<mdibaryon<2085 MeV/c2 no narrow structures could be identified on the 3σ level of statistical significance neither in Mppπ nor in Mppπ+ giving an upper limit (95% C.L.) for dibaryon production in this reaction of σ<20 nb.  相似文献   

11.
Conductivities of some tetraalkylammonium halides, viz. tetramethylammonium iodide (Me4NI), tetraethylammonium bromide (Et4NBr), tetraethylammonium iodide (Et4NI), tetra-n-propylammonium bromide (Pr4NBr), tetra-n-butylammonium bromide (Bu4NBr), tetra-n-butylammonium iodide (Bu4NI) and tetra-n-heptylammonium bromide (Hp4NBr) were measured at 298.15 K in 1,3-dioxolane which has a low permittivity (ε = 7.13). A minima in the conductometric curves (molar conductance, Λ vs. square root of concentration, √c) was observed for concentrations which were dependant upon both the salt and the solvent. The observed molar conductivities were explained by the formation of ion-pairs (M+ + X ↔ MX, KP) and triple-ions (2M+ + X ↔ M2X+; M+ + 2X ↔ MX2, KT). A linear relationship between the triple-ion formation constants [log (KT/KP)] and the salt concentrations at the minimum conductivity (log Cmin) was given for all the salts in 1,3-dioxolane. The formation of triple-ions might be attributed to the ion sizes in solutions in which Coulombic interactions and non-Coulombic interactions act as the main forces between the ions (R4N+…..X).  相似文献   

12.
From the temperature dependence of the line—band luminescence intensity ratio of LiBaF3:Eu2+ a 4f−5d activation energy (Δ) of 800 cm−1 is derived, being much higher than the value reported in the literature (100 cm−1). The temperature dependence of the luminescence decay can be well described with Δ = 800 cm−1 and with 4f−4f and 5d−4f radiative probabilities of 4×102s−1 and 6×105s−1, respectively.  相似文献   

13.
The multilayer relaxation of the Rh(311) surface was investigated by means of LEED structure determination both for vertical and surface parallel (registry) relaxations. Excellent agreement between experimental and calculated spectra could be achieved mirrored by a minimum Pendry R-factor R = 0.174. The first three layer spacings are oscillatorily relaxed by Δd12/d0 = −14.5 ± 1.8%, Δd23/d0 = +4.9 ±2.0% and Δd34/d0 = −1.0 ±2.0%. There seems to be a coherent registry shift of the fir Δs = 0.03 ± 0.07 Å which, however, is within the error limits of the structure determination. Moreover, an energy dependent inner potential is detected. The results are discussed in comparison to equivalent surfaces of other materials as well as for the less open surfaces of rhodium.  相似文献   

14.
T. -U. Nahm  R. Gomer 《Surface science》1997,380(2-3):434-443
The kinetics of H2 desorption from H/W(110) and H/Fe1/W(110) were studied by measuring work function changes Δø vs time at a number of temperatures. Combination with previously determined Δø vs coverage data and differentiation at various fixed coverages gave rate vs T data from which activation energies of desorption could be obtained. E vs coverage results agree well with previously determine ΔHdes results. In the case of H/Fe1/W(110) this includes a rise from 20 to 30 kcal mol−1 of H2 at H/Fe = H/W > 0.3. Plots of rate −dθ/dt vs θ (θ being coverage in units of H/W) vary much more steeply than θ2 at most coverages for both systems. The θ dependence can be explained almost quantitatively in terms of the variations of ΔHdes and surface entropy Ss with coverage, by assuming that rates of desorption are equal to the equilibrium rates of adsorption. The latter can be formulated thermodynamically, except for a sticking coefficient, s. Values for s(θ, T) can also be obtained and show relatively little temperature dependence.  相似文献   

15.
The paper reports on a systematic investigation into the effects of process parameters on the growth kinetics and associated changes in the structure, phase composition and mechanical properties of surface layers formed on Ti–6Al–4V alloy by plasma electrolytic oxidation (PEO) treatment in 0.05–0.2 mol l−1 solutions of sodium aluminate. Methods of gravimetric, SEM and XRD analysis, as well as microhardness and scratch testing, are employed to investigate mass transfer and phase-structure transformations in the surface layer. The probable mechanisms of layer formation are discussed, which comprise electrochemical oxidation of the Ti-electrode by OH anions, complimented by chemical precipitation of Al(OH)3 and plasma-induced transformations in the surface discharges. Running with a total yield efficiency of 20–30%, these processes lead to the formation of predominantly the Al2TiO5 phase with heterogeneous precipitation of Al2TiO5·TiO2 and 3Al2TiO5·Al2O3 eutectics. Al- and Ti-enriched constituents of this structure show hardnesses of 1050–1480 and 300–845 HK, 0.02, respectively. The layer growth rate increases with increasing electrolyte concentration, providing a maximum thickness of over 60 μm and a surface roughness (Ra) of 3–4 μm. Increasing the electrolyte pH from 12.0 to 12.8 results in smoothing and thickening of the surface layer but a lower sample weight gain, associated with an enhancement of the Ti electro-oxidation process. Morphological changes during PEO formation of the surface layer include gradual transformation of the original fine grained but porous structure into a dense, fused morphology which is adversely affected by discharge-induced thermal stresses, causing a degradation of the layer adhesion strength.  相似文献   

16.
The results of the impedance spectroscopy measurements on eutectic samples based on zirconium oxide are presented here. Samples of CaZrO3---ZrO2(cubic) and MgO---ZrO2(cubic) have been grown by a directional solidification procedure such that the different phases appear nearly oriented along the growth direction (lamellae in the system of CaZrO3-ZrO2(cubic) and fibers of MgO in a ZrO2 matrix in the other system). The DC electrical conductivity has been measured by impedance spectroscopy along and across the growth axis. For CaZrO3---ZrO2 the coductivity is clearly anisotropic. The following values for σT have been obtained: the conductivity at 600 °C equals 2.0 × 10−6 Ω−1 cm−1 perpendicular to the fiber axis and 1.4 × 10−5 Ω−1 cm−1 parallel to it and with an activation energy of 1.3 eV for σT. For MgO---ZrO2(cubic) the isotropic value of the conductivity at 600 °C is 10−4 Ω−1 cm−1 with activation energy for σT of 1.5 eV. The anisotropic conductivity in the CaZrO3---ZrO2 (cubic) system has been explained by a model of an ordered stacking of oxygen conducting (cubic ZrO2) and non-conducting (CaZrO3 or MgO) phases.  相似文献   

17.
In this investigation, nonstoichiometries and defect structures of tin oxides were studied between 694 and 990 K by coulometric titration using solid state electrolyte (YSZ) cells. The relationship between nonstoichiometry of the oxide (x) and equilibrium oxygen partial pressure (Po2) was expressed by the proportionality: xPO2−1/6. An intermediate oxide phase, Sn3O4 between Sn and SnO2 was observed in the temperature range of 696–732 K. The standard Gibbs energy of formation of Sn3O4 via the reaction; was found to be ΔGoSn3O4 = −1163960+417.36 T (J/mol). The standard Gibbs energy change for the defect formation reaction in SnO2−x was calculated to be ΔGoSnO2−x = 3.05×105−38.97 T (J/mol)).  相似文献   

18.
Fluorescence decay curves for 5,10,15,20-tetrakis(4-sulfonatophenyl)-21H,23H-porphine tetraanion (TPPS4−) have been measured in the absence and presence of the methylviologen dication (MV2+) with various ionic strengths in methanol. In the presence of MV2+ the fluorescence decays can be expressed by a double exponential function, I(t = I1exp(−t1) + I2exp(−t2). The contribution by the faster decay component to the total fluorescence signal increases with increasing MV2+ concentration. The faster decay process is attributed to fluorescence from the excited state of a solvent-separated ion pair (SSIP) formed between TPPS4− and MV2+, and the slower process is attributed to fluorescence from free TPPS4− ions in the solution. Rate constants for the quenching of fluorescence from free TPPS4− by MV2+ (kq) and formation constants for the SSIP (KSSIP) were calculated and both are found to decrease with increasing ionic strength. The decrease in kq and KSSIP values can be interpreted in terms of the shielding of electrostatic attraction between the ions.  相似文献   

19.
In this paper, the high-order perturbation formulas of spin-Hamiltonian (SH) parameters (g factors g, g and zero-field splitting D), including both the crystal-field (CF) and for the first time charge-transfer (CT) mechanisms, are established for 3d8 ions in trigonal octahedral clusters. By using these formulas, the SH parameters of Ni2+ ions in CsMgX3 (X=Cl, Br, I) crystals are calculated. The results are consistent with the experimental values. The calculations suggest that the sign of QCT (Qg, Δg or D, where the g-shift Δgi=gige, ge≈2.0023 is the value of free-electron) due to CT mechanism is the same as that of the corresponding QCF due to CF mechanism, and the relative importance of CT mechanism (characterized by QCT/QCF) increases with the increasing atomic number of ligand X. So, for the 3dn MLm clusters with ligand having large atomic number, the reasonable theoretical explanations of all SH parameters should take both CF and CT mechanisms into account. The defect structure of (NiX6)4− impurity centers in CsMgX3:Ni2+ crystals is also considered in our model.  相似文献   

20.
In our work single crystals of Mg4.5Na7(P2O7)4 were prepared, pulverized, pressed into pellets and sintered in order to measure the electrical conductivity of polycrystalline specimens. The conductivity was also measured on glassy specimens obtained by the melting of previously prepared crystals. The electrical conductivities at 25°C with values of the order of 10−16 Ω−1 cm−1 for polycrystalline samples and a value of the order of 10−14 Ω−1 cm−1 for glass, show that the glassy phase of Mg4.5Na7(P2 because of its greater molar volume and loosely packed structure, is a better matrix for ionic motion.  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