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 46 毫秒
1.
The exact solution of a random walk model for electron spin flips in a magnetic field is presented and applied to the Stern-Gerlach (SG) deflection of a spin-1/2 molecule. It is found that when the intramolecular spin-relaxation time τ is much longer than t, the residence time of the molecule in the magnetic gradient, the SG deflection spectrum is unaffected, whereas when τ is much shorter than t, the two-line splitting pattern collapses to that of a single line, neither shifted nor broadened with respect to the line profile at zero field gradient. SG band profiles at intermediate values of λ = τ/t are especially interesting. In the range 2.0> λ > 0.5, the Ms = ± 1/2 SG bands are not shifted from their peak position at large λ, however they are strongly broadened asymmetrically toward zero deflection. This broadening assumes the nature of a broad peak at λ ≈ 0.5, which then narrows substantially and rapidly shifts to the position of zero deflection for λ between 0.3 and 0.01. Examples are given of the SG spectra of molecules exhibiting extreme values of λ.  相似文献   

2.
For a small volume (of about 10−6 cm3) of NaCl and other electrolyte solutions (C = 0.1 and 1 M) in thin (r = 5/10 μm) single quartz capillaries, dependencies of the column length l of frozen solutions on the temperature t were measured using comparator IZA-2 in a thermostated chamber. At temperatures range t > −4 °C (for C = 0.1 M) and t > −8 °C (for C = 1 M) the l(t) dependencies are reversible and therefore correspond to establishment of an equilibrium between ice-1 and the solution.

From the constants mass condition of the dissolved salt in a frozen column, the l(t) expression was derived, which includes thermodynamic relation between solution concentration in an equilibrium with ice, Cs, and the temperature t for bulk systems. Deviations from the data known for bulk solutions were observed in thin capillaries when temperature t decreased to −3 °C (for 0.1 M NaCl) and to −6 °C for 1 M NaCl solution.

This effect may be a result of strong adhesion of the ice column to capillary walls. In this case, some internal stresses arise in frozen solution resulting in a deviation from thermodynamic equilibrium conditions for bulk systems. When approaching the temperature of ice melting, adhesion forces decrease due to formation of a thin non-freezing water interlayer on the capillary wall. In this temperature range the experimental data are in agreement with the predictions for bulk systems. It was supposed that the observed deviation in thin capillaries may be caused by formation of an amorphous ice phase with higher density as compared with the ice-1 during rapid freezing, or by an effect of ice microlenses formation. Both effects will result in a deviation from the phase diagram corresponding to a bulk solution.  相似文献   


3.
A flow injection with pulsed amperometric detection for determination of doxycycline or chlortetracycline in pharmaceutical formulations is described. Doxycycline or chlortetracycline were studied at a gold rotating disk electrode with cyclic voltammetry as a function of pH of supporting electrolyte solution. The optimized PAD waveform parameters were obtained with a flow injection system. The optimized pulsed conditions of doxycycline were 1150 mV (versus Ag/AgCl reference electrode) detection potential (Edet) for 220 ms (150 ms delay time and 70 ms integration time), 1500 mV (versus Ag/AgCl reference electrode) oxidation potential (Eoxd) for 70 ms oxidation time (toxd) and 250 mV (versus Ag/AgCl reference electrode) reduction potentail (Ered) for 400 ms reactivation time (tred). The optimized pulsed conditions of chlortetracycline were 1050 mV (versus Ag/AgCl reference electrode) detection potential (Edet) for 300 ms (200 ms delay time and 100 ms integration time), 1300 mV (versus Ag/AgCl reference electrode) oxidation potential (Eoxd) for 70 ms oxidation time (toxd) and 250 mV (versus Ag/AgCl reference electrode) reduction potentail (Ered) for 400 ms reactivation time (tred). The optimized PAD waveform was applied to the determination of doxycycline hydrochloride and chlortetracycline hydrochloride standard solution and in pharmaceutical formulations. The linear dynamic ranges of doxycycline hydrochloride and chlortetracycline hydrochloride were 1 μM–0.1 mM. The sensitivity of this method was found to be 23 μA/mM for doxycycline hydrochloride and 33.76 μA/mM for chlortetracycline hydrochloride. The detection limit for both compounds is 1 μM. The doxycycline hydrochloride and chlortetracycline hydrochloride content in commercially available tablet dosage forms by the proposed method was comparable to those specified by the manufacturer.  相似文献   

4.
We measured FT-IR spectra of intact Acholeplasma laidlawii cells grown at 37 °C on palmitic acid (C16:0) or on binary palmitic acid-d31/oleic acid (C16:0-d31/C18:1(9)) at an initial mole ratio of 2:3, which have been previously reported to produce significant fluctuations in CH2 symmetric stretching (νsCH2) and CD2 asymmetric stretching (νaCD2) frequencies (Biochim. Biophys. Acta 1279 (1996) 49). Time courses for acyl chain νsCH2 and νaCD2 frequencies determined from fourth derivative spectra are presented. Fluctuations were detected with the C16:0 enriched cells at temperatures above 40 °C as well as with the cells enriched in 2:3 C16:0-d31/C18:1(9). These observations at temperatures above 40 °C for the C16:0 enriched cells were not in agreement with the conclusion in the previous work by Moore et al. Our results have suggested that the 2850 cm−1 νsCH2 band comprises two components arising from trans and gauche conformations, and that the fluctuations in νsCH2 frequency are caused by random temporal changes in the relative intensities of these two components.  相似文献   

5.
Xia YX  Chen JF  Choppin GR 《Talanta》1996,43(12):2073-2081
Equilibria in the system of Nd(III) and Th(IV) with 8-hydroxyquinoline (oxine), thenoyltrifluoroacetone (HTTA) and 1,10-phenanthroline (phen) in 5.0 m NaCl solution have been investigated by spectroscopy and potentiometry. The solubility and deprotonation constants of the three organics were measured to be: pKs = 3.09 ± 0.01, pKa1 = 5.82 ±0.02, pKa2= 10.00 ±0.01 for oxine; pKs = 2.49 ± 0.01, pKa1 = 6.47 ±0.03 for HTTA; pKs = 2.86 ± 0.02, pKa2 = 5.82 ± 0.05 for phen. The stabilities of the corresponding metal complexes are in the order M(oxine) > M(TTA) > M(phen), where M = Nd(III), Th(IV). For all three organic ligands, the Th(IV) complexation is stronger than that of Nd(III).  相似文献   

6.
Derivatives of optically active cyanocyclohexylcyclohexanone have been synthesized and used as dipolar chiral dopants to induce ferroelectric SC* phases in an achiral host phase. The dopant molecules are the first examples in which the chiral centres are incorporated into a rigid core with transverse dipoles directly attached. The spontaneous polarization Ps and the tilt angle θ of the induced SC* phases have been measured. Ps is strongly influenced by the relatively small changes of the molecular structure of the cyclohexanones, for example a change of the sign of Ps or a vanishing value of Ps. These effects are discussed in terms of a sterically hindered rotation of the dopant molecules around their long axes and explained by the assumption that the transverse dipole must not be necessarily parallel to Ps in the equilibrium state of rotation.  相似文献   

7.
Equilibrium distribution constants, Ks, of phenol between surfactant micelles and water have been determined by micellar enhanced ultrafiltration (MEUF) using commercial ultrafiltering centrifuge tubes. Three surfactants: sodium dodecyl sulphate (SDS), polyoxyethylene 20 cetyl ether (C16E20) and cetylpiridinium chloride (CPC) were tested with a 10 000 molecular weight cut off (MWCO) membrane. Additionally, membranes of 5000 and 30 000 MWCO were used for CPC. A phenomenological mathematical model has been proposed for the batch MEUF process and checked with the experimental permeate or retentate composition. The model is based on two assumptions: monomeric molecules are not rejected by the membrane and the rejection of micelles is independent of the retentate concentration. The measured micelles rejections for different surfactants and the equivalent molecular weight of the micelles are correlated and they are not significantly affected by the addition of phenol. The estimates of Ks for SDS and CPC agree with previously reported values determined by other methods. Ks values for CPC, calculated using 5000, 10 000 and 30 000 MWCO membranes, have not been significantly different. Ks estimate has allowed to predict the phenol permeate concentration measured in continuous tangential MEUF experiments.  相似文献   

8.
Kanji Miyabe   《Talanta》2007,71(5):1915-1925
Surface diffusion in reversed-phase liquid chromatography (RPLC) using silica gels bonded with C1 and C18 alkyl ligands of different densities was studied from the viewpoints of two extrathermodynamic relationships, i.e., enthalpy-entropy compensation (EEC) and linear free energy relationship (LFER). First, according to the four methods proposed by Krug et al., the values of surface diffusion coefficient (Ds) were analyzed to confirm that an actual EEC resulting from substantial physico-chemical effects takes place for surface diffusion. Then, it was also demonstrated that a LFER is observed between surface diffusion and the retention equilibrium. The establishment of EEC and LFER suggests a mechanistic similarity of molecular migration by surface diffusion, irrespective of the alkyl chain length and the densities of C1 and C18 ligands. Finally, a thermodynamic model for the LFER based on the real EEC was used to estimate Ds values under various RPLC conditions. The Ds values can be estimated with a mean square deviation of about 25–30%. The agreement between the Ds values estimated and those experimentally measured suggests that the total mass flux by surface diffusion consists of the two contributions due to C1 and C18 ligands and that the contribution of each ligand is proportional to the ligand density.  相似文献   

9.
Measurements are reported of the ferroelectric polarization Ps induced in a non-chiral smectic C phase by a variety of chiral dopants having different molecular structural features. Using molecular calculations of contributing dipole moments, ferroelectric order parameters are determined from the experimental results. The relationships between the Ps and various other molecular properties are discussed, and it is shown that restricted rotation of the molecule due to its shape and internal energy barriers to rotation can result in relatively high values of Ps. In contrast dipolar groups flexibly attached to a chiral centre may have their contribution to Ps greatly reduced through internal rotation.  相似文献   

10.
《Chemical physics》1995,200(3):309-318
Dynamics of electronic polarization in the vicinity of charge carriers in molecular crystals is for the first time investigated here in connection with the carrier transport and intramolecular vibronic polarization. According to standard picture it has been assumed that the electronic polarization relaxation time is extremely short, as estimated from the relation τc = τd1h/Eexc, where Eexc is the energy of the first single exciton state. In the case of anthracene (Ac) crystals, the value of τe is about 2 × 10−16 s, i.e. by several orders of magnitude shorter than a typical hopping (residence) time of charge carriers τh = 10−14 -10−13 s. It is argued that typical time of full reconstruction of the electronic polarization after individual carrier hops equals, in the slow carrier regime, approximately to td2hEexc is the width of the lowest singlet-exciton band. In Ac, this means td2 ≈ 0.73 × 10−14 s. Physical implications of this relatively high value of td2 in connection with carrier transport and molecular (vibronic) polarization are discussed.  相似文献   

11.
Experimental data on the spatial distribution of the energy deposited around an energetic heavy ion, from 1 MeV proton to 5.9 MeV/n uranium ion, which have been reported in the literature were documented to obtain a scaled radial dose distribution; (β/Z*)2 D(Z*, β,t)=200 (for t=0–1), 200/t2 (for t=1tc), and 200 tc/t3 (for t>tc) where Z* and β are the effective charge and velocity relative to c, the velocity of light, of the incident ion, respectively, D the dose in unit of Gy, t the radial distance in unit of nm, tc the critical distance empirically determined.

Then, if we know the yield of any chemical reaction as a function of dose from the results of experiments using γ-radiations or fast electrons or theoretical calculations, we can calculate the probability for the yield of the chemical reaction in the system bombarded with a heavy ion of the effective charge Z* and velocity β. The results of the present calculation of the LET-values and of G(Fe3+) in the ferrous sulfate acidic solution are presented and compared with reported experimental results.  相似文献   


12.
Reactions of the hydrated electron, H atoms, 2-propanol, and methanol radicals with the TiO2 nano-particles have been studied either directly or by competition kinetics. The radicals were produced by radiolysis of 2-propanol, t-butanol, or methanol aqueous solutions in acid pH's. The reactions involve electron injection to the conduction band. As expected, the t-butanol radical is inert towards TiO2 under our conditions, while the other reducing radicals react with TiO2. The reactivity decreases in the order: eaq>H>CH3COHCH3>CH2OH. Two TiO2 nanocrystallite sizes, with average diameters of 1.0 and 4.7 nm were compared. For equal concentrations (in terms of TiO2 molecules), the rate of electron injection shows relatively little dependency on particle size. The rates of interfacial electron transfer and transfer coefficient are also reported.  相似文献   

13.
A novel method for determining the viscosity of polymer solution   总被引:1,自引:0,他引:1  
The relative viscosity ηr and, thus, the reduced viscosity ηsp/C of polymer solution could be obtained by recording the flow times of the polymer solution and the pure solvent in a capillary viscometer. Our experimental results indicated that the measurement of the flow time of the pure solvent was unnecessary. In particular, if the recorded flow time of the pure solvent was used to determine the viscosity of polymer solution, the reduced viscosity ηsp/C exhibited either a drastic increase or a significant decrease in an extremely dilute solution, depending upon the properties of the polymer solution investigated. In this research work, a new method for determining the viscosity of polymer solutions is reported. In the proposed method, the flow time of polymer solution at zero concentration, t0*, instead of the measured flow time of the pure solvent, was used to determine the viscosity of polymer solution. The reduced viscosity ηsp/C determined by the new method is proportional to concentration C even in an extremely dilute solution. The relative viscosity ηr vs. C plot also indicated clearly that t0*, instead of the measured flow time of the pure solvent, should be used for determining the viscosity of polymer solution. At low concentrations, the flow time of the polymer solution was proportional to C. As a result, t0* could be determined by extrapolating the flow time of the polymer solution to C=0.  相似文献   

14.
Fan J  Wang J  Ye C 《Talanta》1998,46(6):1285-1292
The acid dissociation constants (Ka), base dissociation constants (Kb) and the autoprotolysis constants (Ks) for 2,2′-bipyridyl in water and in water+alcohol(methanol, ethanol, iso-propanol) mixed solvents have been determined at 25°C and an ionic strength of 0.1 mol l−1, from a direct potentiometric method based on the treatment of the data of a single pH titration. It has been shown that Ka increases, whereas Kb and Ks decrease, with increasing proportion of the alcohol in the mixed solvents. Linear relations between pKa, pKb, pKs and the mole fraction of the alcohol were observed in the composition range investigated. These results are discussed in terms of the properties of solvent and the interactions of the different species existing in dissociation equilibrium with solvents. It is concluded that the higher stabilization of both 2,2′-bipyridyl and its protonated form by dispersion forces and of the proton by its interaction with solvent molecules in the mixed solvents compared with that in water are largely responsible for the observed changes of pKa with composition. On the other hand, the low stabilization of OH in the mixed solvents relative to that in water and the electrostatic effect are the main factors in determining the solvent effect on pKb.  相似文献   

15.
The complex permittivity of a room-temperature ferroelectric liquid crystal 4-n-octyloxy benzoic acid 4'-[(2-methylbutyloxy)carbonyl]phenyl ester, has been measured in the vicinity of the phase transition in the frequency range 40 Hz-300 kHz. In the para-electric phase the contribution εs of the soft mode to the permittivity and the soft-mode relaxation frequency fs satisfy the Curie-Weiss law. Under a biasing field E, the helicoidal structure is unwound, and εs and fs can then be measured even in the ferroelectric phase. On the other hand, the phase transition is smeared under the influence of E. This smearing results in deviations from the Curie-Weiss law for both εs and fs in the vicinity of the transition. On increasing E, the maximum of the permittivity, εmax, is lowered and shifted to higher temperatures. Both the shift and ε-1max are proportional to E2/3. From experimentally found dependences, some constants in the free energy are determined.  相似文献   

16.
Impact tensile fracture testing of a brittle polymer   总被引:1,自引:0,他引:1  
The fracture behavior of a brittle polymer, methylmethacrylate–butadiene–styrene resin, under impact tensile loading was studied using single-edge-cracked specimens. The dynamic load and displacement were measured with a Piezo sensor and a high-speed extensometer, respectively. The load and displacement diagram, i.e., the external work, Uex, applied to the specimen was used to determine the elastic energy, Ee, and non-elastic energy, En, due to viscoelastic and plastic deformation, and the fracture energy, Ef, for creating new fracture surface, As. The energy-release rates were then estimated using Gt=Uex/As and Gf=Ef/As. The values of Gt and Gf were correlated with the fracture loads and the mean crack velocities determined from the load and time relationships.  相似文献   

17.
Anion exchange membrane has been investigated in different electrolyte solutions by chronopotentiometry to explore the influence of co-ion and counterion of the exchange group of the membrane, on the transport phenomena. Chloride, nitrate, sulfate and acetate in sodium salts were used as counterions and sodium, potassium, calcium and ammonium in chloride salts were used as co-ions. The membrane showed a potential drop (E0) in all these electrolytes when a constant current was applied across it, which remained constant for a period less than τ, called the transition time and rose gradually to a maximum (Emax) value. The parameters such as τ, E0 and Emax and the potential jump (ΔE) and τ and the inflection zone (Δt) along the time axis have been measured and compared at an applied current density (I) of 10 mA cm−2 in 10 mM solutions. The values of τ1/2/zA[A0] or τ1/2/zC[C0], with or , E0 and ΔE with or (where rA and rC are the ionic radii of counter and co-ions, respectively) have been correlated. Permselectivity (P) and transference number of the membrane with respect to each one of the above electrolytes have been evaluated and discussed.  相似文献   

18.
Induced S*C phases can be obtained by dissolving chiral dopants in achiral SC host phases. If the chiral guest molecules bear a transverse dipole, ferroelectricity will occur. The novel dopants under discussion are characterized by chiral centres and the transverse dipole situated not in the alkyl end groups of the mesogenic molecules, but directly in their rigid cores. As a rigid core, analogues of decalin were used. In those dopants, rotation around the molecular long axis is sterically restricted. According to the microscopic model of Zeks, this leads to enhanced values of the spontaneous polarization Ps. The magnitude as well as the sign of the spontaneous polarization Ps of the S*C phases induced by the novel dopants in different host phases has been investigated. It has been found for the first time that for a given dopant, the polarization as well as the sign of Ps depends on the structure of the host phase. The results are discussed in terms of two microscopic models. They can be understood taking into account the situation that the potential of the restricted long axial rotation is determined by the hard core interactions of the molecules involved or that an orientation of the host dipoles by a guest/host interaction takes place.  相似文献   

19.
We report new photomechanical effects in the ferroelectric liquid crystal SCE13 doped with a photoisomerizing guest azo dye. Low concentrations of dye (∼5 per cent wt:wt) are shown to cause an isothermal, reversible disruption of smectic phases when the system is illuminated with low power density (∼ 1 mW cm-2) UV light. In the case of a sample initially in the S*c phase, this results in a fall in the magnitude of spontaneous electrical polarization (Ps) and changes in electro-optic switching characteristics. If the sample is illuminated in the SA phase, the electroclinic switching decreases. In contrast to this, when systems containing higher concentrations of dye (≥ 10 per cent wt: wt) are UV illuminated in the SA phase, a reversible, isothermal transition to a biphasic S*c/isotropic state occurs. In this case, the Ps is seen to rise from zero in the SA phase to a finite value(∼2 nC cm-2) in the biphasic mixture and hysteresis occurs in the electro-optic switching. When these higher dye concentration mixtures are held initially in the S*c phase and UV illuminated, a more complicated variation of Ps occurs with the sample again undergoing a transition to a biphasic S*c/isotropic state. Possible mechanisms for the transition are discussed.  相似文献   

20.
Reduction of cyclo-(t-Bu4Sb4) (1) with sodium or potassium in boiling tetrahydrofuran leads to the anions [t-Bu4Sb3] and [t-Bu3Sb2]. Crystallization with pentamethyldiethylenetriamine (L) gives [M(L)n(t-Bu4Sb3)] (n=1, M=Na (2), K (3); n=2, M=K (4)) and [K(L)(t-Bu3Sb2)] (5). Crystal structure analyses reveal coordination of the anionic antimony ligands on the alkali metal ions for 2, 3, and 5. In contrast, no Sb---K interactions were observed in the structure of 4.  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