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 109 毫秒
1.
The field dependences of photocurrent, the two-beam coupling gain coefficient, and the grating formation time constant in polymer composites made from polyvinylcarbazole (PVK) and single-wall carbon nanotubes (SWNTs) were measured under the conditions of one-photon SWNT excitation with continuous laser radiation at a wavelength of 1550 nm. Carbon nanotubes are responsible for optical electronic absorption up to ~2000 nm in this composite. The dependence of the quantum efficiency of generation of mobile charge carriers on the electric field E 0 as determined from the photocurrent coincides with the curves calculated via the Onsager equation expanded to the (E 0)4 term, at a quantum yield of thermalized electron-hole pairs of η0 = 0.07 and a charge separation distance in the pair of r 0 = 9.8 Å. An analysis of the photorefractive characteristics showed that the admixture of fullerene C60 in an amount of 3 wt % to the PVK composite with 0.26 wt % SWNT leads to a twofold increase in the beam-coupling gain coefficient. In the PVK-matrix composite containing 0.26 wt % SWNT and 3 wt % C60, the beam-coupling gain coefficient Γ of a 1550-nm laser beam and the net gain Γ-α are 32 and ~27 cm?1, respectively, at a constant field of E 0 = 140 V/μm.  相似文献   

2.
The photoelectric and photorefractive characteristics of polyvinylcarbazole (PVK) composites with 0.15 wt % graphene at a wavelength of 532 nm have been measured. The dependence of the quantum efficiency of generation of mobile charge carriers as determined from the photocurrent is well approximated by the Onsager equation calculated accurate to E 0 3 for the quantum yield of thermalized electron-hole pairs of φ0 = 1 and their initial separation radius of r 0 = 10.9 Å. The long-wavelength edge of optical absorption for PVK lies at 365 nm. Thus, the photogeneration of mobile carriers during illumination of the composite at 532 nm is due to photoexcitation of graphene. The measurement of the photorefractive properties has revealed that the two-beam coupling gain coefficient is Γ = 50 cm?1 at E 0 = 150 V/μm and equal intensities of incident beams. It has been found that at E 0 = 83.3 V/μm, the two-beam coupling gain coefficient increases from 8 to 14 cm?1 as the ratio of incident beam intensities I 1(0)/I 2(0) increases from 1 to 2.4.  相似文献   

3.
Poly(N-vinylcarbazole) layers containing tetra-5-crown-5-gallium phthalocyaninate (R4Pc)Ga(OH) are shown to possess photoelectric and photorefractive sensitivity at a wavelength of 1064 nm. This effect is associated with the formation of supramolecular ensembles of (R4Pc)Ga(OH) molecules with electronic optical absorption in the near-IR range and nonlinear optical properties. For the composite containing 5 wt % (R4Pc)Ga(OH) supramolecular ensembles, the dependence of the quantum efficiency of mobile-charge photogeneration on electric field E 0 is well fit by the Onsager equation expanded to E 0 3 at a quantum yield of electron-hole pairs of φ0 = 0.9 s with an initial separation radius of r 0 = 9.8 Å susceptibility χ(3) equal to 1.85 × 10?10 esu is measured via the well-known method of electric-field-induced second-harmonic generation. Two-beam-coupling gain coefficient Γ is found to be 80 cm?1 at E 0 = 120 V/μm.  相似文献   

4.
Photorefractive materials based on unplasticized polymers that have a high glass-transition temperature and the frozen random orientation of chromophores were prepared by layer casting. Under these conditions, only the third-order susceptibility has a nonzero value, increasing with an increase in the conjugation-chain length and reaching considerable values in the case of nanosized molecules, such as single-wall carbon nanotubes (SWNTs). In SWNT-containing polyvinylcarbazole, the photoelectric sensitivity and photorefractive characteristics were measured at 1064 nm. Using electric-field induced second-harmonic generation, the third-order susceptibility was estimated. In a composite containing 0.26 wt % SWNT, the diffraction grating is displaced by 5° relative to the interference pattern, the result that is presumably due to the close mobility of unlike charge carriers. Therefore, the beam-coupling gain coefficient Γ and the net gain Γ-α have low values, which are 53 and 42 cm?1, respectively, at 115 V/μm.  相似文献   

5.
The photoelectric and photorefractive characteristics of composites based on poly(vinylcarbazole) containing crown-substituted ruthenium phthalocyanine with axially coordinated pyrazine molecules, (R4Pc)Ru(pyz)2 (where R4Pc?2 is [4,5,4′,5′,4″,5″,4″′,5″′-tetrakis-(1,4,7,10,13-pentaoxatridecamethylene)phthalocyanine ion], pyz is pyrazine), have been studied. Supramolecular ensembles of these complexes are responsible for optical absorption and photoelectric and photorefractive sensitivity in the near IR region. It has been found that under the action of a 1064-nm laser, the quantum efficiency of formation of mobile charge carriers, estimated from the photocurrent, corresponds to the Onsager equation when the quantum yield of formation of thermalized electron-hole pairs φ0 = 0.35 and the separation distance is 9.8 Å, irrespective of the (R4Pc)Ru(pyz)2 content. The kinetic curves of amplification of the information laser beam are bellshaped. This suggests that the photorefractive characteristics are underestimated owing to lowering to zero of the field E 0 inside the layer as a result of buildup of space charge in the near-electrode space upon passing the dark current. The two-beam gain coefficient at an (R4Pc)Ru(pyz)2 content of 7 wt % is Γ = 62 cm?1 as estimated from the maximum of the bell-shaped curve.  相似文献   

6.
The formation of complexes at pH 4.7 of the Hg(II) with five monothiosemicarbazone and two dithiosemicarbazone has been studied. The mercury(II) reacts with monothiosemicarbazones of salicylaldehyde (λmax = 363 nm, E = 1.69 × 104liters · mol?1cm?1), pi-colinadehyde (λmax = 363 nm, E = 2.38 × 104liters · mol?1cm?1), 6-methyl-picolinaldehyde (λmax = 363 nm, E = 2.28 × 104liters · mol?1cm?1), di-2-pyridylketone (λmax = 380 nm, E = 2.08 × 104liters · mol?1cm?1), and o-naphthoquinone (λmax = 540 nm, E = 1.03 × 104liters · mol?1cm?1) and with dithiosemicarbazones of 1,4-dihydroxyphthalimide (λmax = 430 nm, E = 2.56 × 104liters · mol?1cm?1) and dipyridylglyoxal (λmax = 363 nm, E = 2.37 × 104liters · mol?1cm?1). A critical comparison of the stoichiometry and apparent stability constant of complexes with mono- and dithiosemicarbazones is given.  相似文献   

7.
The third-order nonlinear optical properties of the ruthenium (II) complex with tetra-15-crown-5-phthalocyanine and axially coordinated triethylenediamine molecules (R4Pc)Ru(TED)2 were analyzed by means of the z-scanning technique. A solution of (R4Pc)Ru(TED)2 in tetrachloroethane was exposed to nanosecond laser pulses at a wavelength of 1064 nm. It was found that the third-order molecular polarizability of the Ru(II) complex is 4.5 × 10?32 cm4/C (esu). The polarizability per molecule increases by a factor of 3.6 when the single molecule occurs in a supramolecular assembly of (R4Pc)Ru(TED)2 complexes. The photoelectric and photorefractive properties at 1064 nm of polymer composites, determined by the supramolecular assemblies that exhibits optical absorption and photoelectric sensitivity in the near IR region, are reported.  相似文献   

8.
The absorption spectrum of the U3+ ion in anhydrous formic acid was recorded in the range 4100–23,000 cm?1. The electrostatic, spin-orbit and configuration interaction parameters obtained from a least-squares fit to eight observed levels are: E1 = 2882.7 cm?1, E2 = 13.8 cm?1, E3 = 285.4 cm?1, ξ5f = 1654 cm?1, α = 19.8 cm?1, β = ?380.3 cm?1 and γ = 1000.0 cm?1. Intensity calculations gave good agreement with the oscillator strengths of observed bands only if some of the experimental band areas were combined.  相似文献   

9.
Single crystals of PdPSe were shown to be n-type semiconductors. Weak Pauli paramagnetic behavior was observed, which is consistent with the presence of delocalized electrons. Electrical measurements showed a room-temperature resistivity ? = 70 ohm-cm, activation energy of resistivity Ea = 0.32 eV, and Hall mobility μ = 34 cm2 V?1 sec?1. Photoelectronic measurements in aqueous solutions of I?I?3 indicate that PdPSe has high quantum efficiencies below 800 nm. The indirect optical band gap is 1.28(2) eV.  相似文献   

10.
The energies of low-energy optical transitions were calculated, using the linearized attached cylindrical waves (LACW) method, as a function of inverse diameter d ?1 for (n, n) metal nanotubes with n ranging from 3 to 12 and for (n, 0) semiconductor nanotubes with n ranging from 10 to 25. The calculations show that E 11 in the metal nanotubes is higher than in the semiconductor nanotubes. Significant violations of the E 11d ?1 relationship are observed. For metal nanotubes, the situation is more complex because of the close energies of the π-π*-and σ-π* vertical transitions and because of the intersection of these characteristics in the range of 0.7 nm?1 < d ?1 < 1.0 nm?1 (n = 8, 9, 10). For the semiconductor nanotubes, the E 11 versus d ?1 relationship is not linear, rather, it is oscillating; the E 11(d ?1) function alternates between two curves that refer to the (n, 0) nanotubes for which division of n by 3 gives 1 or 2 in the residue. On one hand, these features hamper the use of optical measurements in structure determination for the nanotubes; on the other, new criteria for nanotube classification appear.  相似文献   

11.

Silicate‐based inorganic‐organic hybrid polymer systems have many unique properties including thermal stability and photo‐stability, chemical resistance with the combination of tunable optical properties. Two kinds of new UV‐patternable hybrid materials PSQ‐Ls were synthesized by a sol‐gel process at room temperature, which can be used for low cost fabrication of optical waveguides. Thick films (up to 8.31 µm) can be coated by a single spin‐coating process without any cracking and the average surface roughness (Ra), detected by atomic force microscopy (AFM), is below 0.5 nm. The optical properties (refractive index, birefringence, and optical loss at 1310 nm and 1550 nm, respectively) of the PSQ‐Ls films are investigated by a prism coupler. The refractive index of PSQ‐Ls can be exactly tuned from 1.4483 to 1.5212 by blending PSQ‐LH (nTE=1.5212 @ 1310 nm) and PSQ‐LL (nTE=1.4483 @ 1310 nm). The maximum refractive index contrast is about 4.8%. After post‐baking, birefringences of the films are below 0.0005 and optical losses are about 0.2 dB · cm?1 at 1310 nm, 0.7 dB · cm?1 at 1550 nm, respectively. Furthermore, the PSQ‐Ls films also show outstanding thermal stability in air atmospheres.  相似文献   

12.
New soluble MoS2 nanosheets covalently functionalized with poly(N‐vinylcarbazole) (MoS2–PVK) were in situ synthesized for the first time. In contrast to MoS2 and MoS2/PVK blends, both the solution of MoS2–PVK in DMF and MoS2–PVK/poly(methyl methacrylate) (PMMA) film show superior nonlinear optical and optical limiting responses. The MoS2–PVK/PMMA film shows the largest nonlinear coefficients (βeff) of about 917 cm GW?1 at λ=532 nm (cf. 100.69 cm GW?1 for MoS2/PMMA and 125.12 cm GW?1 for MoS2/PVK/PMMA) and about 461 cm GW?1 at λ=1064 nm (cf. ?48.92 cm GW?1 for MoS2/PMMA and 147.56 cm GW?1 for MoS2/PVK/PMMA). A larger optical limiting effect, with thresholds of about 0.3 GW cm?2 at λ=532 nm and about 0.5 GW cm?2 at λ=1064 nm, was also achieved from the MoS2–PVK/PMMA film. These values are among the highest reported for MoS2‐based nonlinear optical materials. These results show that covalent functionalization of MoS2 with polymers is an effective way to improve nonlinear optical responses for efficient optical limiting devices.  相似文献   

13.
For composites based on poly(vinylcarbazole) and multiwall carbon nanotubes, photorefractive characteristics at 1064 nm have been studied in the presence and absence of fullerene C60. It has been found that the introduction of fullerene C60 provides an increase in the two-beam gain coefficient of a laser beam and a reduction in the time of hologram recording. The preillumination of the fullerene-containing layer with a continuous light from a He-Ne laser (633 nm) leads to an additional increase in the two-beam gain coefficient.  相似文献   

14.
The nonlinear optical properties of solutions of (2,3,9,10,16,17,23,24-tetra-15-crown-5-phthalocyaninato)indium(III) [(15C5)4Pc]In(OH) in tetrachloroethane (TCE) have been studied by the z-scan method. It has been found that a nonlinear optical response is due to supramolecular associates formed in a tetrachloroethane solution by heating to 90°C/slow cooling to room temperature cycling. The formation of the supramolecular associates has been studied by atomic force microscopy (AFM) and electronic absorption spectroscopy (EAS). It has been shown that a single thermal treatment of [(15C5)4Pc]In(OH) solutions in TCE results in the predominant formation of dimers, as evidenced by both a short-wavelength shift of the Q-absorption band of the monomeric complex (λmax = 692 nm) to the band of λmax = 653 nm and height doubling of molecular entities as measured by AFM. The dimers are responsible for the two-photon absorption measured in the femtosecond range, which has a relatively high cross section of σ2 = 1.38 × 10?46 cm4 s/(molecule, photon) or 1.38 × 104 GM. According to the AFM data, three cycles of heat treatment of the solution leads to the formation of supramolecular assemblies of about 200 nm length. The optical spectrum exhibits long-wavelength absorption at λmax = 841 nm and the long-wavelength edge near 1300 nm. In the case of nanosecond 1064-nm laser irradiation, the linear absorption S 0S 1 is primary, having the cross section of σ0 = α0/N = 2.3 × 10?20 cm2. The known high quantum yield (close to unity) of triplet states of indium phthalocyanines suggests that the main nonlinear optical effect is determined by intersystem crossing S 1T 1 and triplet-triplet absorption T 1T 2. The absorption cross section is σ T-T = 1.14 × 10?19 cm2.  相似文献   

15.
In this paper we report the results of an experimental study of the vacuum ultraviolet absorption spectra of molecular impurity states of methyl iodide in Ar (density range ? = 0–1.4 g cm?3) and in Kr (? = 0–2.3 g cm?3), of carbon disulphide in Ar (? = 0–1.4 g cm?3) and of formaldehyde in Ar (? = 0–1.25 g cm?3). The experimental results provide new information regarding medium perturbations of intravalenc transitions, of the lowest extravalence transitions and of transitions to mixed valence—Rydberg configurations, which serve as a diagnostic tool to distinguish between different types of electronic excitations. All the lowest extravalence molecular excitations exhibit appreciable blue spectral shifts at moderate and at high fluid densities, intravalence transitions are practically insensitive to medium effects, while excitations to mixed valence—Rydberg configurations are characterized by a moderate blue spectral shift. New information has been obtained concerning the energetics of molecular ionization processes in a dense fluid. The high n = 2–5 Rydberg states of CH3l exhibit a large red shift at moderate (? = 0–0.5 cm?3) Ar densities. The ionization potential Eg and the effective Rydberg constant G for CH3I in Ar was found to decrease from G = 13.6 eV and Eg = 9.55 eV at ? = 0 and Eg = 9.08 eV and constant G for CH3l in Ar was found to decrease from G = 13.6 eV and Eg = 9.55eV at ? = 0 and Eg = 9.08 eV and G ≈ 7.15 eV at ? = 0.5 g cm?3. Experimental evidence was obtained for the identification of n = 2 molecular Wannier impurity states of CH3I and of CH2O in liquid Ar. These spectroscopic data result in Eg ≈ 8.6 eV for CH3I in liquid Ar and Eg ≈ 10.2 eV for CH2O in liquid Ar.  相似文献   

16.
Oxygen can be determined in a perfluorotributylamine emulsion used as a blood substitute by coulometry and polarography. The oxygen uptake of the emulsion (4.3 × 10?3 mol l?1 or 11 ml-% at 25°C and PO2 = 760 mm Hg) is about three times greater than that of water. The adsorption of surfactant on a dropping mercury electrode changes the electrochemical parameters E12, α and k3. The marked difference between the diffusion coefficients of oxygen and hydrogen peroxide (ratio 2.6) seems to be proceed from the ejection of oxygen molecules from the perfluorotributylamine droplets and from the insolubility of hydrogen peroxide molecules in the droplets. The constant rate of oxygen release by these droplets was estimated to be 104 s?1 by a stopped-flow spectrophotometric method. This constant rate seems to be linked with the particle diameter (0.2 μm) and the diffusion coefficient in the perfluorotributylamine droplets (1.6 × 10?6 cm2 s?1).  相似文献   

17.
The interaction of hydrogen atoms with strong laser fields at intensities up to some 1013 W cm?2 was studied experimentally at the wavelengths λ=355 nm, 532 nm and 1064 nm. The ion yield, the energy spectrum of the photoelectrons and their angular distributions were measured. The angular distributions at λ=355 nm and λ=532 nm provide a sensitive test for theoretical calculations. Comparison with the calculations available shows that perturbation theory with proper inclusion of atomic structure yields results which agree with experiment. Intensity dependent changes of angular distributions at λ=532 nm are observed, which indicate that at 1013 W cm?2 higher order processes become noticable. At λ=1064 nm the situation is more complicated, experimentally as well as theoretically. Intensities of some 1013 W cm?2 are necessary to observe ionization. Strong distortions of the atomic structure can be expected. Presently only qualitative aspects of the angular distributions can be discussed.  相似文献   

18.
The energy transfer rate for the reaction DF (ν=1) + DF (ν=1)kνν→ DF (ν=0) + DF (ν=2) + ΔE=91.6 cm?1 has been studied in a combined shock-tube laser-induced fluorescence experiment at temperatures from 295 to 720°K. The rate coefficient kνν for the exothermic reaction was found to vary as T?1 when expressed in units of cm3/mole sec. At T=295°K, the probability of the reaction is approximately 0.2 per collision.  相似文献   

19.
EPR and optical absorption spectra of Cu2+ ion were investigated in natural elbaites from Brazil and Zambia and in synthetic olenite single crystal. In elbaite from Zambia, the content of Cu2+ ions was found to be about 0.006 pfu, whereas in Brazilian elbaite the amount of this ion can approach up to 0.2 pfu. The rose color of elbaite from Zambia is mainly due to optical absorption at 515 nm related to Mn3+ ions. The blue color of Brazilian elbaite is related to Cu2+ absorption bands at 695 nm and 920 nm. Spin Hamiltonian parameters of Cu2+ calculated from the angular dependence of the EPR spectra are: g x = 2.054, g y = 2.092, g z = 2.374; A x = 27.8·10?4 cm?1, A y = 59.3·10?4 cm? 1, A z = 133.2·10?4 cm?1. We propose that Cu2+ ions enter into Y octahedra with common edges; the symmetry of these Y octahedra is lowered because of local disorder induced by occupancy of the Y site by cations of very different size and charge, such as Li+, Al3+, and Cu2+.  相似文献   

20.
The photorefractive properties of a polyimide containing J-aggregates of a thiacarbocyanine dye were examined. Measurements were made with a holographic apparatus at Ar–Kr laser wavelengths of 647 and 514 nm. The J-aggregate and the monomeric form of the dye, respectively, are responsible for the photoelectric sensitivity at these wavelengths. It was shown that the J-aggregates are also responsible for the nonlinear optical properties of the polymeric system. The combination of these features determines the photorefractive characteristics of the polymer composite: under interference conditions, the intensity of one of the interacting beams (signal beam) increases at the expense of extinction of the other (pumping) beam. The difference between the beam-coupling gain coefficient and the absorption coefficient increases at 647 nm from 217 to 361 cm–1 with the increasing strength of a constant electric field E 0 applied to the layer from 16 to 123 V/m. Only the samples containing J-aggregates display the photorefractive effect at 514 nm.  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