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 31 毫秒
1.
ABSTRACT

5-Bromo-, 3,5-dibromo- and 5-nitrosalicylaldehyde acetohydrazones (BSAH, DBSAH and NSAH, respectively, or H2L) were synthesized. The extraction of lanthanoid ions (Ln3+) including praseodymium, europium and ytterbium ions into 1,2-dichloroethane with the synthesized hydrazones in the presence of both tri-n-butylphos-phate (TBP) and perchlorate has been investigated. The extracted species were (Ln3+)(HLP?)2(TBP)3(CI4 ?) for all the hydrazones. The acid dissociation constants, Ka, and the partition coefficients, Ko, of the hydrazones and the extraction constants, Kex, of the lanthanoid complexes were determined. The introduction of electron-withdrawing bromo or nitro groups to the 3- and/or 5-position of the benzene ring in salicylaldehyde acetohydrazone (SAH), the mother compound of the synthesized hydrazones, was very effective for increasing not only the Kex value of SAH but also its KD value and the Kex values of its lanthanoid complexes. Taking into consideration the above-mentioned three constants, of the synthesized hydrazones DBSAH was the most balanced and recommend-able extractant for the lanthanoids.  相似文献   

2.
Sal icylaldehyde n-alkanohydrazones (SAAH or H2L) having 1 to 13 carbon atoms in their alkyl chains were synthesized and used for the extraction of some lanthanoid ions (Ln3+) including praseodymium, europium and ytterbium ions from aqueous solutions containing both tri-n-butyIphosphate (TBP) and perchlorate into I,2-dichloroethane The extracted species were (Ln3+) (HL-)2(TBP)3(C101,-) for all SAAH. The partition coefficients, Kd, of SAAH increased with an increase in the carbon number of the alkyl chain, the increment of log Kb being 0.64 per a carbon atom. The acid dissociation constants of SAAH also increased slightly with an increase in the carbon number, but the extraction constants of their lanthanoid complexes were almost independent of the length of the alkyl chain.  相似文献   

3.
BACKGROUND: 2‐ethylhexylphosphonic acid mono‐(2‐ethylhexyl) ester (HEHEHP, H2A2) has been applied extensively to the extraction of rare earths. However, there are some limitations to its further utilization and the synergistic extraction of rare earths with mixtures of HEHEHP and another extractant has attracted much attention. Organic carboxylic acids are also a type of extractant employed for the extraction of rare earths, e.g. naphthenic acid has been widely used to separate yttrium from rare earths. Compared with naphthenic acid, sec‐nonylphenoxy acetic acid (CA100, H2B2) has many advantages such as stable composition, low solubility, and strong acidity in the aqueous phase. In the present study, the extraction of rare earths with mixtures of HEHEHP and CA100 has been investigated. The separation of the rare earth elements is also studied. RESULTS: The synergistic enhancement coefficient decreases with increasing atomic number of the lanthanoid. A significant synergistic effect is found for the extraction of La3+ as the complex LaH2ClA2B2 with mixtures of HEHEHP and CA100. The equilibrium constant and thermodynamic functions obtained from the experimental results are 10?0.92 (KAB), 13.23 kJ mol?1H), 5.25 kJ mol?1G), and 26.75 J mol?1 K?1S), respectively. CONCLUSION: Graphical and numerical methods have been successfully employed to determine the stoichiometries for the extraction of La3+ with mixtures of HEHEHP and CA100. The mixtures have different extraction effects on different rare earths, which provides the possibility for the separation of yttrium from heavy rare earths at an appropriate ratio of HEHEHP and CA100. Copyright © 2009 Society of Chemical Industry  相似文献   

4.
ABSTRACT

The solvent extraction of some tervalent lanthanoid ions (Ln3+) with five 4-acyl-3-phenyl-5-isoxazolones (AcyPI or HL),i.e. 4-acetyl-, 4-benzoyl-, 4- (4-toluoyl)-, 4- (4-fluorobenzoyl) - and 4-(4-nitrobenzoyl)-3-phenyl-5-isoxazolones (API,BPI, TPI, FBPI and NBPI respectively) into benzene was studied in the presence of tri-n-butylphosphate (TBP). The extracted species were (Ln3+) (L-)3 (TBP)2 for AcyPI.  相似文献   

5.
Experiments at various Sb2O3 concentrations were made in a pilot plant and the effect of Sb2O3 on continuous esterification between terephthalic acid (TPA) and ethylene glycol (EG) was obtained. Reaction rate constants of the previously reported reaction scheme were determined to fit with the experimental data obtained. It was found that the effect of Sb2O3 on reaction rate constant (ki) can be expressed as follows.
  • k1 = (3.75 × 10?4Sb + 1.0) × 1.5657 × 109exp(?19,640/RT)
  • k2 = (4.75 × 10?4Sb + 1.0) × 1.5515 × 108exp(?18,140/RT)
  • k3 = (6.25 × 10?4Sb + 1.0) × 3.5165 × 109exp(?22,310/RT)
  • k4 = (4.50 × 10?4Sb + 1.0) × 6.7640 × 107exp(?18,380/RT)
  • k5 = (3.50 × 10?4Sb + 1.0) × 7.7069 × exp(?2810/RT)
  • k6 = (1.75 × 10?4Sb + 1.0) × 6.2595 × 106exp(?14.960/RT)
  • k7 = (3.75 × 10?4Sb + 1.0) × 2.0583 × 1015exp(?42,520/RT)
Simulation of esterification with these reaction rate constants at various Sb2O3 concentrations was made and the following results were obtained.
  • 1 Sb2O3 accelerates the esterification reaction between TPA and EG.
  • 2 Sb2O3 accelerates the main reaction and its effects on side reactions are minor. The higher the addition rate of Sb2O3, the lower the carboxyl end-group concentration (AV) and diethylene glycol content (DEG).
  • 3 The comparison between simulation with potassium titanium oxyoxalate (PTO) in the previous report and with Sb2O3 in the present report shows that the acceleration of polycondensation reaction by Sb2O3 is higher. DEG formation rate is lower with PTO than Sb2O3.
  相似文献   

6.
Amphiphilic diblock copolymers (DCs) of 2,3,4,5,6-pentafluorostyrene (PFS) and 2-hydroxyethyl methacrylate (HEMA) of different composition and molecular weights were prepared by two-step reversible addition–fragmentation chain transfer (RAFT) polymerization and first used for preparation of superhydrophobic coatings for cotton/polyester fabrics. The transition from hydrophobic to superhydrophobic coatings is controlled by the ratio between poly(2,3,4,5,6-pentafluorostyrene) (PPFS) block and poly(2-hydroxyethyl methacrylate) (PHEMA) block lengths (PnPFS/PnHEMA). The increase in PnPFS/PnHEMA is accompanied by a significant increase in water (θН2О) and diiodomethane (θCH2I2) contact angles, which reach the plateau at PnPFS/PnHEMA = 3.5 and remains almost constant up to PnPFS/PnHEMA = 6.2. Surface modification of the cotton/polyester fabric with the DC having PnPFS/PnHEMA = 6.2 produced superhydrophobic surface with θН2О = 158 ± 4° and contact angle hysteresis CAH = 5 ± 2°, and θCH2I2 = 107 ± 3°.  相似文献   

7.
Red grape pomace (RGP), an abundant wine industry solid waste, was used for the recovery of polyphenols and anthocyanin pigments, using ultrasound-assisted extraction and water/glycerol mixtures as the solvent. Glycerol concentration (Cgl) and liquid-to-solid ratio (RL/S) were first optimized by implementing Box?Behnken experimental design and the process was further studied through kinetics. The optimal conditions were found to be Cgl = 90% (w/v) and RL/S = 90 mL g?1, and under these conditions the extraction of total polyphenols (TP) and total pigments (TPm) obeyed first-order kinetics. Maximum diffusivity (De) values were 4.22 × 10?12 and 12.59 × 10?12 m2 s?1, for TP and TPm, respectively, and the corresponding activation energies were (Ea) 13.94 and 8.22 kJ mol?1.  相似文献   

8.
The unperturbed dimensions and thermodynamic parameters of poly(vinylpyrrolidone) (PVP) have been studied in aqueous salt solutions, e.g. phosphates, mono- and dihydrogen phosphates, carbonates, sulphates of sodium and potassium. Values of K0 ( = [η]ΘM-1/2, where [η]Θ is intrinsic viscosity at the theta temperature and M is molecular weight) with Mw = 78 000 g mol-1 were found to range from 4·63×10-4 to 5·56×10-4 dl g-1, and root-mean-square end-to-end distances, 〈r201/2, ranging from 1·61×10-6 to 1·68×10-6cm were evaluated. Values of the characteristic ratio, Cn, the steric parameter, σ, and the enthalpy and the entropy of dilution parameters, χH and χS, have also been calculated, and the interaction parameter was found to be χ-0·5<-0·001 for aqueous salt solutions of PVP. ©1997 SCI  相似文献   

9.
The graft copolymerization of acrylamide (AAm) and ethylmethacrylate (EMA) monomers onto cellulose has been carried out using ceric ammonium nitrate (CAN) as initiator in presence of nitric acid at (25 ± 1)°C and varying feed molarity from 7.5 × 10?2 mol dm?3 to 60.0 × 10?2 mol dm?3 at fixed feed composition (fAAm = 0.6). The graft yield (%GY) has shown a linear increasing trend upto a feed molarity of 37.5 × 10?2 mol dm?3. The composition of grafted copolymer chains was found to be constant (FAAm = 0.56) during feed molarity variation but shown variations with feed composition (fAAm) and reaction temperature. The grafting parameters have shown increasing trends up to 7.5 × 10?3 mol dm?3 concentration of ceric (IV) ions and decreased on further increasing the concentration of ceric (IV) ions beyond 7.5 × 10?3 mol dm?3. The IR and elemental analysis data were used to determine the composition of grafted chains (FAAm) and reactivity ratio of acrylamide (r1) and ethylmethacrylate (r2) comonomers. The reactivity ratio for acrylamide (r1) and ethylmethacrylate (r2) has been found to be 0.7 and 1.0 respectively, which suggested for an alternate arrangement of average sequence length of acrylamide (mM?1) and ethylmethacrylate (mM?2) in grafted chains. The rate of graft copolymerization of comonomers onto cellulose was found to be proportional to square concentration of comonomers and square root to the concentration of ceric (IV) ions. The energy of activation (ΔEa) of graft copolymerization was found to be 9.57 kJ mol?1 within the temperature range of 20–50°C. On the basis of experimental findings, suitable reaction steps have been proposed for graft copolymerization of selected comonomers. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 2546–2558, 2006  相似文献   

10.
By using laser light scattering (LS) and size exclusion chromatography combined LS, we have investigated the molecular weight and chain conformation of amylopectin from rice of India (II‐b), japonica (IJ‐b), and glutinous (IG‐b) in dimethyl sulfoxide (DMSO) solution. The weight‐average molecular weight (Mw) and radius of gyration (〈S2½) of amylopectin were determined to be 4.06 × 107 and 128.5 nm for India rice, 7.41 × 107 and 169.6 nm for japonica rice, 2.72 × 108 and 252.3 nm for glutinous rice, respectively. The 〈S2½ values were much lower than that of normal polymers, indicating a small molecular volume of amylopectin, as a result of highly branched structure. Ignoring the difference of degree of branching, approximated dependences of 〈S2½ and intrinsic viscosity ([η]) on Mw for amylopectin in DMSO at 25°C were estimated to be 〈S2½ = 0.30Mw0.35 (nm) and [η] = 0.331Mw0.41 (mL g?1) in the Mw range studied. Moreover, from the 〈S2½ values of numberless fractions obtained from many experimental points in the SEC chromatogram detected with LS, the dependence of 〈S2½ on Mw for the II‐b sample was estimated also to be 〈S2½ = 0.34 Mw0.347, coinciding with the above results. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

11.
The kinetics of 2,2′-azobisisobutyronitrile (AIBN) initiated polymerization of styrene in N,N-dimethylformamide (DMF) at 60°C were investigated in the presence of dibromo(N,N-dimethylformamide)copper(II) complex. The complex was prepared in situ by mixing tetrakis (N,N-dimethylformamide)copper(II) perchlorate with LiBr in the molar ratio of 1 : 2. The equilibrium constant for [Cu(DMF)4]2+ + 2Br? ? Cu(DMF)2Br2 + 2DMF was calculated by the limiting logarithmic method as 1.80 × 103 L2 mol?2. The velocity constant at 60°C for the interaction of polystyryl radical with Cu(DMF)2Br2 is 7.46 × 104 L mol?1 s?1.  相似文献   

12.
The gas permeabilities of three polyacetylene films, prepared from poly[1-(trimethylsilyl)-1-propyne], poly(tert-butylacetylene), and poly(1-chloro-2-phenylacetylene), were studied. Although depending on conditions of polymerization and membrane preparation, typical permeability coefficients P of the polymers to oxygen and nitrogen at 25°C were as follows: poly[1-(trimethylsilyl)-1-propyne], PO2 = 40 × 10?8, PNr2 = 20 × 10?8; poly(tert-butylacetylene), PO2 = 3.0 × 10?8, PN2 = 1.0 × 10?8; poly(1-chloro-2-phenylacetylene), PO2 = 9.4 × 10?10, PN2 = 2.0 × 10?10 cm3(STP) · cm/(cm2 · s · cm Hg). Thus PO2 of a poly[1-(trimethylsilyl)-1-propyne] film is the largest among those ever known, and the values of poly(tert-butylacetylene) and poly(1-chloro-2-phenylacetylene) films are also fairly large. Influences of polymer structure, measuring temperature, and so forth on the PO2 and PN2 of these polyacetylene films were studied. The possibility of applying these films to oxygen enrichment of air are being discussed.  相似文献   

13.
M. Zielinska 《Drying Technology》2016,34(10):1147-1161
The objective of this study was to dry–wet distillers grains and centrifuged solubles and to examine the effect of two different drying media, superheated steam and hot air, at different drying temperatures (110, 130, and 160°C), moisture contents (5–30% wb), and percentages of solubles’ presence (0 or 100%) on some thermophysical properties of laboratory-prepared corn/wheat dried distillers co-products, including geometric mean diameter (dg), particle density (ρp), bulk density (ρb), bulk porosity (?b), specific heat (C), effective thermal diffusivity (αeff), and bulk thermal conductivity (λb). The values of dg of corn/wheat dried distillers co-products ranged from 0.358 ± 0.001 to 0.449 ± 0.001 mm. Experimental values of ρp, ρb, and ?b varied from 1171 ± 6 to 1269 ± 3 kg m?3, from 359 ± 7 to 605 ± 5 kg m?3, and from 0.54 ± 0.01 to 0.71 ± 0.01 kg m?3, respectively. The values of αeff were between 0.58 × 10?7 and 0.93 × 10?7 m2 s?1. The calculated values of C ranged from 1887 ± 11 to 2599 ± 19 J kg?1 K?1, and the values of λb of corn/wheat dried distillers co-products ranged from 0.06 ± 0.01 to 0.09 ± 0.01 W m?1 K?1. Multiple linear regression prediction models were developed to predict the changes in dg, ρp, ρb, ?b, C, αeff, and λb of laboratory-prepared corn/wheat dried distillers co-products with different operational factors.  相似文献   

14.
The chemical stability of perfluorinated and non‐perfluorinated low temperature fuel cell model compounds (MCs) against attack by hydroxyl radicals, HO, is compared using a competition kinetics approach in aqueous solutions at ambient temperature. HO radicals were generated in situ by UV photolysis of hydrogen peroxide in the electron spin resonance (ESR) resonator. Acetic acid (AA), trifluoroacetic acid (TFAA), methanesulfonic acid (MSA), trifluorosulfonic acid (TFSA), and perfluoro(2‐ethoxyethane)sulfonic acid (PFEESA) were chosen as MCs, while the rate constants of 5,5‐dimethyl‐1‐pyrroline‐N‐oxide (DMPO) and methanol (CH3OH) served as reference for the determination of relative rate constants by means of steady state ESR signal amplitudes. In decreasing order the rate constants are: kMSA = (4.8 ± 0.2) × 107 M–1 s–1, kAA = (4.2 ± 0.3) × 107 M–1 s–1, kPFEESA = (3.7 ± 0.1) × 106 M–1 s–1, kTFAA = (7.9 ± 0.2) × 105 M–1 s–1, and kTFSA < 1.0 × 105 M–1 s–1. Applying these results to perfluorinated fuel cell membranes like Nafion®, the main points of attack by HO are concluded to be the ether groups of the side chains, followed by the remaining carboxyl groups from the manufacturing process of the polymers.  相似文献   

15.
Benzoyl peroxide (BPO)‐initiated free radical copolymerization of citronellol with butylmethacrylate (BMA) in xylene at 80°C ± 0.1°C under the inert atmosphere of nitrogen has been studied. The kinetics expression is Rp α [I]0.5±0.27 [citronellol]1.0±0.13 [BMA]1.0±0.18. The overall activation energy has been calculated as 65 kJ/mol. Bands at 3436 and 1732 cm?1 in the FTIR spectrum of the copolymer(s) have indicated the presence of hydroxy, ester group of citronellol and butylmethacrylate, respectively. The 1H‐NMR spectrum shows peaks at 7.0–7.7 δ due to ? OH proton of citronellol and at 3.2–4.0 δ due to ? OCH2 proton of butylmethacrylate. The molecular weight Mv and ηint of the copolymers have been measured with the help of gel permeation chromatography in tetrahydrofuran at 25°C to calculate Mark‐Houwink constants as K = 2.68 × 10?4 and α = 0.34 ± 0.40. The alternating nature of the copolymer is confirmed by reactivity ratios r1 (BMA) = 0.023 ± 0.004 and r2 (Citronellol) = 0.0025 ± 0.22. The Alfrey‐Price Qe parameters for citronellol have been calculated as Q2 = 0.13 and e2 = –1.28. Thermal decompositions of copolymer are evaluated with the help of thermal gravimetric analysis technique. The mechanism of copolymerization has been elucidated. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

16.
Spherical nickel oxide (NiO) nanoparticles were prepared by using nickel chloride as precursor in the ethylene glycol as solvent and urea as precipitant. The X‐ray diffraction study showed that NiO has single‐phase cubic structure with average crystallite size of 35 nm. The prepared NiO nanoparticles were incorporated into polyaniline (PANI) matrix during in situ chemical oxidative polymerization of aniline with different molar ratios of aniline: NiO (12 : 1, 6 : 1, and 3 : 1) at 5°C using (NH4)2S2O8 as oxidant in aqueous solution of sodium dodecylbenzene sulfonic acid, as surfactant and dopant under N2 atmosphere. The synthesized composites have been characterized by means of X‐ray diffraction (XRD), thermogravimetric analysis, Fourier transform infrared (FTIR), scanning electron microscopy, TEM, and vibrating sample magnetometer for its structural, thermal, morphological, and magnetic investigation. The XRD and FTIR studies show that the NiO particles are in the composite. The room temperature conductivities of the synthesized PANI, PANI/NiO (12 : 1), (6 : 1), and (3 : 1) composites were found to be 3.26 × 10?4, 1.88 × 10?4, 1.5 × 10?4, and 4.61 × 10?4 S/cm, respectively. The coercivity (Hc) and remnant magnetization (Mr) of NiO, PANI/NiO NCs (12 : 1), (6 : 1), and (3 : 1) at 5 K was found to be 8.22 × 10?2, 6.31 × 10?2, 6.42 × 10?2, 6.27 × 10?2 T, and 6.64 × 10?3, 1.83 × 10?4, 3.07 × 10?4, and 3.98 × 10?4 emu/g, respectively. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

17.
ABSTRACT

During the extraction of lithium from high Mg-containing salt lake brines by tributyl phosphate (TBP) in the presence of Fe(III), H+ is used to stabilize Fe(III). However, the distribution ratio of H+ (DH) is 4–6 times higher than that of Li+ (DLi), which affects the extraction of Li+ significantly. In this study, the competition mechanism between H+ and Li+ was investigated by spectral analysis and thermodynamic equilibrium. The extracted species are determined as HFeCl4 · 2TBP and LiFeCl4 · 2TBP for H+ and Li+, respectively. The apparent equilibrium constants are KH = 799.8 and KLi = 120.6, respectively. Both equilibrium constants and the distribution ratios for H+ and Li+ extraction show that extraction of H+ is stronger than Li+.  相似文献   

18.
Dialkyl vinylphosphonates such as dimethyl vinylphosphonate (DMVP) and diethyl vinylphosphonate were quantitatively polymerized with dicumyl peroxide (DCPO) at 130°C in bulk. The polymerization of DMVP with DCPO was kinetically studied in bulk by fourier transform near‐infrared spectroscopy (FTNIR) and electron spin resonance (ESR) spectroscopy. The initial polymerization rate (Rp) was given by Rp = k[DCPO]0.5[DMVP]1.0 at 110°C, being the same as that of the conventional radical polymerization involving bimolecular termination. The overall activation energy of the polymerization was estimated to be 26.2 kcal/mol. The polymerization system involved ESR‐observable propagating polymer radicals under the practical polymerization conditions. ESR‐determined rate constants of propagation (kp) and termination (kt) were kp = 19 L/mol s and kt = 5.8 × 103 L/mol s at 110°C, respectively. The molecular weight of the resultant poly(DMVP)s was low (Mn = 3.4 ? 3.5 × 103), because of the high chain transfer constant (Cm = 3.9 × 10?2 at 110°C) to the monomer. DMVP (M1) showed a considerably high reactivity in the radical copolymerization with trimethoxyvinylsilane (TMVS) (M2) at 110°C in bulk, giving an inorganic component‐containing functional copolymer with potential flame‐retardant properties; r1 = 1.6 and r2 = 0. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

19.
BACKGROUND: The effect of acrylic acid neutralization on the degradation of alkoxyamine initiators for nitroxide‐mediated polymerization (NMP) was studied using styrene/acrylic acid and styrene/sodium acrylate random copolymers (20 mol% initial acrylate feed concentration) as macro‐initiators. The random copolymers were re‐initiated with fresh styrene in 1,4‐dioxane at 110 °C at SG1 mediator/BlocBuilder® unimolecular initiator ratios of 5 and 10 mol%. RESULTS: The value of kpK (kp = propagation rate constant, K = equilibrium constant) was not significantly different for styrene/acrylic acid and styrene/sodium acrylate compositions at 110 °C (kpK = 2.4 × 10?6–4.6 × 10?6 s?1) and agreed closely with that for styrene homopolymerization at the same conditions (kpK = 2.7 × 10?6–3.0 × 10?6 s?1). All random copolymers had monomodal, narrow molecular weight distributions (polydispersity index M?w/M?n = 1.10–1.22) with similar number‐average molecular weights M?n = 19.3–22.1 kg mol?1. Re‐initiation of styrene/acrylic acid random copolymers with styrene resulted in block copolymers with broader molecular weight distributions (M?w/M?n = 1.37–2.04) compared to chains re‐initiated by styrene/sodium acrylate random copolymers (M?w/M?n = 1.33). CONCLUSIONS: Acrylic acid degradation of the alkoxyamines was prevented by neutralization of acrylic acid and allowed more SG1‐terminated chains to re‐initiate the polymerization of a second styrenic block by NMP. Copyright © 2008 Society of Chemical Industry  相似文献   

20.
Highly transparent (Y0.95?xGdxEu0.05)2O3 (= 0.15–0.55) ceramics have been fabricated by vacuum sintering at the relatively low temperature of 1700°C for 4 h with the in‐line transmittances of 73.6%–79.5% at the Eu3+ emission wavelength of 613 nm (~91.9%–99.3% of the theoretical transmittance of Y1.34Gd0.6Eu0.06O3 single crystal), whereas the = 0.65 ceramic undergoes a phase transformation at 1650°C and has a transparency of 53.4% at the lower sintering temperature of 1625°C. The effects of Gd3+ substitution for Y3+ on the particle characteristics, sintering kinetics, and optical performances of the materials were systematically studied. The results show that (1) calcining the layered rare‐earth hydroxide precursors of the ternary Y–Gd–Eu system yielded rounded oxide particles with greatly reduced hard agglomeration and the particle/crystallite size slightly decreases along with increasing Gd3+ incorporation; (2) in the temperature range 1100°C–1480°C, the sintering kinetics of (Y0.95?xGdxEu0.05)2O3 is mainly controlled by grain‐boundary diffusion with similar activation energies of ~230 kJ/mol; (3) Gd3+ addition promotes grain growth and densification in the temperature range 1100°C–1400°C; (4) the bandgap energies of the (Y0.95?xGdxEu0.05)2O3 ceramics generally decrease with increasing x; however, they are much lower than those of the oxide powders; (5) both the oxide powders and the transparent ceramics exhibit the typical red emission of Eu3+ at ~613 nm (the 5D07F2 transition) under charge transfer (CT) excitation. Gd3+ incorporation enhances the photoluminescence and shortens the fluorescence lifetime of Eu3+.  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