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.
Upconversion luminescence tuning of β‐NaYF4 nanorods under 980 nm excitation has successfully been achieved by tridoping with Ln3+ ions with different electronic structures. The effects of Ce3+ ions on NaYF4:Yb3+/Ho3+ as well as Gd3+ ions on NaYF4:Yb3+/Tm3+(Er3+) have been studied in detail. By tridoping with Ce3+ ions, not only were unusual 5G55I7 and 5F2/3K85I8 transitions from Ho3+ ions and 5d→4f transitions from Ce3+ ions observed in NaYF4:Yb3+/Ho3+ nanorods, but also an increase in the intensity of 5F55I8 relative to 5S2/5F45I8 with increasing Ce3+ concentration, which can be attributed to efficient energy transfers of 5I6 (Ho)+2F5/2 (Ce)→5I7 (Ho)+2F7/2 (Ce) and 5S2/5F4 (Ho)+2F5/2 (Ce)→5F5 (Ho)+2F7/2 (Ce). Interestingly, with increasing pump power density, the luminescence of NaYF4:Yb3+/Ho3+ nanorods is always dominated by the 5S2/5F45I8 transition, whereas the luminescence of Ce3+‐tridoped NaYF4:Yb3+/Ho3+ nanorods is dominated by the 5S2/5F45I8 and 5G55I7 transitions in turn. These observations are discussed on the basis of a rate equation model. Furthermore, Gd3+‐tridoped NaYF4:Yb3+/Tm3+(Er3+) nanorods can emit multicolor upconversion emissions spanning from the UV to the near‐infrared under 980 nm excitation. 6P5/28S7/2 (≈306 nm) and 6P7/28S7/2 (≈311 nm) transitions from Gd3+ ions were observed. In addition to the aforementioned luminescence properties, these Gd3+‐tridoped nanorods also exhibit paramagnetic behavior at room temperature and superparamagnetic behavior at 2 or 5 K.  相似文献   

2.
UV photolysis of [CpFeII(CO)3]+ PF66? (I) or [CpFeII6-toluene)]+ PF6?? (II) in CH3CN in the presence of 1 mole of a ligand L gives the new air sensitive, red complexes [CpFeII(NCCH3)2L]+PF6? (III, L = PPh3; IV; L = CO; VIII, L = cyclohexene; IX, L = dimethylthiophene) and the known air stable complex [CpFeII(PMe3)2(NCMe)]+ PF6? (V). The last product is also obtained by photolysis in the presence of 2 or 3 moles of PMe3. In the presence of dppe, the known complex [CpFeII (dppe)(NCCH3)]+ (XI) is obtained. Complex III reacts with CO under mild conditions to give the known complex [CpFe(NCCH3)(PPh3)CO]+ PF6? (X). UV photolysis of I in CH3CN in the presence of 1-phenyl-3,4-dimethylphosphole (P) gives [CpFeIIP3]+ PF6? (XII); UV photolysis of II in CH2Cl2 in the presence of 3 moles of PMe3 or I mole of tripod (CH3C(CH2Ph2)3) provides an easy synthesis of the known complexes [CpFeII(PMe3)3]+ PF6? (VII) or [CpFeIIη3-tripod]+ PF6t- (XIII). Since I and II are easily accessible from ferrocene, these photolytic syntheses provide access to a wide range of piano-stool cyclopentadienyliron(II) cations in a 2-step process from ferrocene.  相似文献   

3.
The paper presents a radiokinetic study on the appearance and growth of*Fe2S3,*Fe(OH)3,*Fe2(C2O4)3,*Fe(IO3)3 crystals in a colloidal medium of agar and gelatine. The values of the diffusion constants through gels of55+59Fe3+ radioactive cations and of the rate of global growth process of these crystals in agar or gelatine were calculated using the experimental data. A new method for the determination of the starting time for the complex nucleation process was proposed. The formation rate of crystals under study decreases in the order:*Fe(OH)3>*Fe(IO3)3>*Fe2S3>*Fe2(C2O4)3, in agar medium and*Fe(OH)3>*Fe(IO3)3>*FeC2O4)3>*FeS3, in gelatine medium.  相似文献   

4.
Reactions of Li+ [(η5-C5H5)Re(NO)(PPh3)] with 2- and 4-chloroquinoline or 1-chloroisoquinoline give the corresponding σ quinolinyl and isoquinolinyl complexes 3 , 6 , and 8 . With 3 and 8 there is further protonation to yield HCl adducts, but additions of KH give the free bases. Treatment of 3 with HBF4⋅OEt2 or H(OEt2)2+ BArf gives the quinolinium salts [(η5-C5H5)Re(NO)(PPh3)(C(NH)C(CH)4C (CH)(CH))]+ X ( 3-H + X; X=BF4/BArf, 94–98 %). Addition of CF3SO3CH3 to 3 , 6 , or 8 affords the corresponding N-methyl quinolinium salts. In the case of [(η5-C5H5)Re(NO)(PPh3)(C(NCH3)C(CH)4C (CH)(CH))]+ CF3SO3 ( 3-CH3 + CF3SO3), addition of CH3Li gives the dihydroquinolinium complex (SReRC,RReSC)-[(η5-C5H5)Re(NO)(PPh3)(C(NCH3)C(CH)4C (CHCH3)(CH2))]+ CF3SO3 ((SReRC,RReSC)- 5 + CF3SO3, 76 %) in diastereomerically pure form. Crystal structures of 3-H + BArf, 3-CH3 + CF3SO3, (SReRC, RReSC)- 5 + Cl, and 6-CH3 + CF3SO3 show that the quinolinium ligands adopt Re⋅⋅⋅ C conformations that maximize overlap of their acceptor orbitals with the rhenium fragment HOMO, minimize steric interactions with the bulky PPh3 ligand, and promote various π interactions. NMR experiments establish the Brønsted basicity order 3 > 8 > 6 , with Ka(BH+) values >10 orders of magnitude greater than the parent heterocycles, although they remain less active nucleophilic catalysts in the reactions tested. DFT calculations provide additional insights regarding Re⋅⋅⋅ C bonding and conformations, basicities, and the stereochemistry of CH3Li addition.  相似文献   

5.
The ultraviolet absorption spectrum of CF3CFClO2 and the kinetics of the self reactions of CF3CFCl and CF3CFClO2 radicals and the reactions of CF3CFClO2 with NO and NO2 have been studied in the gas phase at 295 K by pulse radiolysis/transient UV absorption spectroscopy. The UV absorption cross section of CF3CFCl radicals was measured to be (1.78 ± 0.22) × 10?18 cm2 molecule?1 at 220 nm. The UV spectrum of CF3CFClO2 radicals was quantified from 220 nm to 290 nm. The absorption cross section at 250 nm was determined to be (1.67 ± 0.21) × 10?18 cm2 molecule?1. The rate constants for the self reactions of CF3CFCl and CF3CFClO2 radicals were (2.6 ± 0.4) × 10?12 cm3 molecule?1 s?1 and (2.6 ± 0.5) × 10?12 cm3 molecule?1 s?1, respectively. The reactivity of CF3CFClO2 radicals towards NO and NO2 was determined to (1.5 ± 0.6) × 10?11 cm3 molecule?1 s?1 and (5.9 ± 0.5) × 10?12 cm3 molecule?1 s?1, respectively. Finally, the rate constant for the reaction of F atoms with CF3CFClH was determined to (8 ± 2) × 10?13 cm3 molecule?1 s?1. Results are discussed in the context of the atmospheric chemistry of HCFC-124, CF3CFClH. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
Ionization energies of substituted trimethylamines (CH3)2NCH2X, together with appearance energies of [M ? X]+ ions, have been measured by the photoionization technique. The heats of formation for the following radicals have been derived (in kJ mol?1): C2H3˙ (+ 314), Ph. (+ 347), p-CH3C6H4˙ (+ 326), HC2˙ (+ 494), (CH3)2N?H2 (+ 134), NC˙ (+ 452), CF3. (? 435), (CH3)3Si˙ (+ 54.5). The heats of formation obtained for [(CH3)2NCH2]+ (+ 649) and (CH3)3Si+ (+ 610) lead to ionization energies of 5.35 eV for (CH3)2N?H2 and 5.75 eV for (CH3)3Si˙.  相似文献   

7.
The reactions of CCl3 with O(3P) and O2 and those of CCl3O2 with NO have been studied at 295 K using discharge flow methods with helium as the bath gas. The rate coefficient for the reaction of CCl3 with O was found to be (4.2 ± 0.6) × 10?11 cm3/s and that for CCl3O2 with NO was (18.6 ± 2.8) × 10?12 cm3/s with both coefficients independent of [He]. For reaction between CCl3 and O2 the rate coefficient was found to increase from 1.51 7times; 10?14 cm3/s to 7.88 × 10?14 cm3/s as the [He] increased from 3.5 × 1016 cm?3 to 2.7 × 1017 cm?3. There was no evidence for a direct two-body reaction, and it is concluded that the only product of this reaction is CCl3O2. Examination of these results for CCl3 + O2 in terms of current simplified falloff treatment suggests that the high-pressure limit for this reaction is ~ 2.5 × 10?12 cm3/s, which may be compared with a direct measurement of the high-pressure limit of 5 × 10?12 cm3/s. A value of (5.8 ± 0.6) × 10?31 cm6/s has been obtained for k0, the coefficient in the low-pressure region. This value is compared with corresponding values found earlier for the (CH3, O2) and (CF3, O2) systems and with estimates based on unimolecular rate theory.  相似文献   

8.
1,1,1,4,5,5,5-Heptafluoro-4-(trifluoromethyl)-2,3-pentanedione reacted with λ3σ3-phosphorus compounds, PR1R2R3 (R1 = CF3, R2 = R3 = Me, iPr, NEt2; R1 = NCO, R2 = R3 = OMe, OEt, R2−R3 = OCH2CH2O, OCMe2CMe2O; R1 = OSiMe3, R2 = R3 = OEt; R1 = NEt2, R2 = R3 = OCH2CF3; R1 = R2 = Et2N, R3 = OCH2CF3, OCH(CF3)2, OCH2Ph, OC6F5) to give new 1,3,2λ5σ5-dioxaphospholenes. The first λ5σ5 phosphoranes with an OCN group bonded to phosphorus were obtained. © 1998 John Wiley & Sons, Inc. Heteroatom Chem 9:109–113, 1998  相似文献   

9.
Different methods for the preparation of fluorinated iminium salts RR1CNR2R3+MF6? (R=R1=F ; R2=R3=CH3, C2H5 M=As, Sb 4a ? c R=H, R1=F; R2=R3=CH3 M=As, Sb 5a, b R=R1=CF3; R2=H, R3=CH3 M=Sb 12 R=R1=CF3; R2=R3=CH3 M=As 14) are reported, the spectroscopic properties (IR, NMR) of the cations of these salts are briefly discussed. By F?-addition to these salts, e.g. to 16, perfluoroalkyl-bis(alkyl)-amines (e.g. (CF3)2CFN(CH3)2 15) can be prepared; from the methylation of CF3NCF2 bis(trifluoromethyl) methylamine (CF3)2NCH3 (11) was obtained.  相似文献   

10.
Two-photon induced fluorescence and resonance-enhanced photoionization have been observed in atomic sulfur originating from both the 3P2,1,0 and the 1D2 states. Sulfur atoms are generated by the sequential multiphoton dissociation of CS2 at probing wavelengths. The two-photon absorption process involves the 3 3P2,1,0 → 4 3P2,0,1 or the 3 1D2 → 4 1F3 transitions with resolution of the individual J″ → J′ transitions in most cases. Intensities of the 33PJ → 4 3PJ transitions have been compared with Hartree-Fock calculated transition probabilities from the analogous transitions in atomic oxygen. Photoionization is observed in a three-photon (two to resonance) ionization originating from the 3P2,1,0 and the 1D2 states. Induced fluorescence is observed at 167 and 180 nm which is dipole-allowed radiation from the intermediate 3S01 and 1D02 states, respectively.  相似文献   

11.
LaInO3: Sm3+, LaInO3: Pr3+ and LaInO3: Tb3+ phosphors were prepared through a Pechini-type sol–gel process. X-ray diffraction, field emission scanning electron microscopy, photoluminescence, and cathodoluminescence (CL) spectra were utilized to characterize the synthesized phosphors. XRD results reveal that the pure LaInO3 phase can also be obtained at 700 °C. FE-SEM images indicate that the LaInO3: Sm3+, LaInO3: Pr3+ and LaInO3: Tb3+ phosphors are composed of aggregated spherical particles with sizes around 80–120 nm. Under the excitation of ultraviolet light and low voltage electron beams (1–5 kV), the LaInO3: Sm3+, LaInO3: Pr3+ and LaInO3: Tb3+ phosphors show the characteristic emissions of Sm3+ (4G5/26H5/2,7/2,9/2 transitions, yellow), Pr3+ (3P03H4, 3P13H5, 1D23H4 and 3P03F2 transitions, blue–green) and Tb3+ (5D47F6,5,4,3 transitions, green) respectively. The corresponding luminescence mechanisms are discussed. These phosphors have potential applications in field emission displays.  相似文献   

12.
The ground state rotational spectra often isotopic species of trimethylamineborane, (CH3)3N10BH3, (CH3)3N11BH3, (CH3)3N10BD3, (CH3)3N11BD3, (CH3)3N11BD2H, (CD3)3N10BH3, (CD3)3N11BH3, (CD3)3N10BD3, (CD3)3N11BD3 and (13CH3)(12CH3)2N11BH3, have been measured and the effective moments of inertia obtained. The utilization of Kraitchman's equations leads to an rs value of the B-H distance of 1.211±0.003 Å and a NBH angle of 105.32±0.16°. By a least squares fit of the rotational constants the following structural parameters were obtained: r(NC) = 1.495 Å, r(BN) = 1.609 Å, and ∠BNC = 110.9°. The value of the dipole moment was found to be 4.59±0.13 D. A lower limit to the barrier to internal rotation of the BH3 group was determined to be 3.4 kcal/mole.  相似文献   

13.
The synthesis of imido analogues of p-block oxoanions is a relatively new area of investigation. This review summarises synthetic routes towards such anions and gives an analysis of the structure and reactivity of complexes containing the following anions, [B(NR)3]3−; [C(NR)3]2−; [Si(NR)4]4−; [P(NR)4]3−; [HP(NR)3]2−; [As(NR)3]3−; [Sb(NR)3]3−; [S(NR)3]2−; [Se(NR)3]2−; [Te(NR)3]2− and [S(NR)4]2−.  相似文献   

14.
Concentration‐optimized CaSc2O4:0.2 % Ho3+/10 % Yb3+ shows stronger upconversion luminescence (UCL) than a typical concentration‐optimized upconverting phosphor Y2O3:0.2 % Ho3+/10 % Yb3+ upon excitation with a 980 nm laser diode pump. The 5F4+5S25I8 green UCL around 545 nm and 5F55I8 red UCL around 660 nm of Ho3+ are enhanced by factors of 2.6 and 1.6, respectively. On analyzing the emission spectra and decay curves of Yb3+: 2F5/22F7/2 and Ho3+: 5I65I8, respectively, in the two hosts, we reveal that Yb3+ in CaSc2O4 exhibits a larger absorption cross section at 980 nm and subsequent larger Yb3+: 2F5/2→Ho3+: 5I6 energy‐transfer coefficient (8.55×10?17 cm3 s?1) compared to that (4.63×10?17 cm3 s?1) in Y2O3, indicating that CaSc2O4:Ho3+/Yb3+ is an excellent oxide upconverting material for achieving intense UCL.  相似文献   

15.
In this study, we investigated the effects of four inorganic anions (Cl, SO42−, H2PO4/HPO42−, and HCO3/CO32−) on titanium dioxide (TiO2)-based photocatalytic oxidation of aqueous ammonia (NH4+/NH3) at pH  9 and ∼10 and nitrite (NO2) over the pH range of 4–11. The initial rates of NH4+/NH3 and NO2 photocatalytic oxidation are dependent on both the pH and the anion species. Our results indicate that, except for CO32−, which decreased the homogeneous oxidation rate of NH4+/NH3 by UV-illuminated hydrogen peroxide, OH scavenging by anions and/or direct oxidation of NH4+/NH3 and NO2 by anion radicals did not affect rates of TiO2 photocatalytic oxidation. While HPO42− enhanced NH4+/NH3 photocatalytic oxidation at pH  9 and ∼10, H2PO4/HPO42− inhibited NO2 oxidation at low to neutral pH values. The presence of Cl, SO42−, and HCO3 had no effect on NH4+/NH3 and NO2 photocatalytic oxidation at pH  9 and ∼10, whereas CO32− slowed NH4+/NH3 but not NO2 photocatalytic oxidation at pH  11. Photocatalytic oxidation of NH4+/NH3 to NO2 is the rate-limiting step in the complete oxidation of NH4+/NH3 to NO3 in the presence of common wastewater anions. Therefore, in photocatalytic oxidation treatment, we should choose conditions such as alkaline pH that will maximize the NH4+/NH3 oxidation rate.  相似文献   

16.
The gas-phase clustering reactions of proton in propanol and acetone, and chloride ions in acetone were investigated. The −ΔHn−1,n values obtained for clustering reactions (n−1,n) were as follows: H+ (C3H7OH)n−1 + C3H7OH ⇄ H+ (C3H7OH)n, (2, 3) 18.9 kcal mol−1, (3, 4) 14.2 kcal mol−1, (4, 5) 11.7 kcal mol−1; H+ (CH3COCH3)2 + CH3COCH3 ⇄ H+ (CH3COCH3)3, 14.2 kcal mol−1; and Cl + CH3COCH3 ⇄ Cl (CH3COCH3), 12.4 kcal mol.−1. For clustering reactions, Cl (CH3COCH3n−1 + CH3COCH3 ⇄ Cl (CH3COCH3)n where n ≥ 2, the equilibria could not be established; probably due to the isomerization of ligand acetone molecules from the keto to enol form.  相似文献   

17.
The influence of differently substituted cyclopentadienyl CpR ligands on the reaction outcome of [CpRFe(CO)2]2 (CpR = C5Me5, EtC5Me4, 1,3-Bu2tC5H3) with As4 is examined. For C5Me5 and EtC5Me4, the pentaarsaferrocene derivatives [CpRFe(η5-As5)] are formed together with [(CpRFe)3As6] and [(CpRFe)3As6{(η3-As3)Fe}], while for 1,3-Bu2tC5H3 only [(CpRFe)3As6] is formed. The reaction of [(Me5C5Fe)3As6{(η3-As3)Fe}] with Tl+ leads to [{(Me5C5Fe)3As6Fe}2(μ,η33-As3)]2+ representing an unexpected dicationic cluster.  相似文献   

18.
Molecular structures and energies have been calculated in the MNDO approximation, for P4S3 and its molecular ion P4S3+, and for the mass spectral fragment pairs: (P3S3+ + P), (P3S2+ + PS), (P3S+ + PS2), (P2S3+ + P2), (P2S2+ + P2S), (P2S+ + P2S2), (P2S2), PS3+ + P3), (PS2+ + P3S), (PS+ + P3S2), and (PS+ + P2S + PS). Three distinct energy minima were found for each of P2S2+ and P2S2, and two minima for each of P2S+, P2S, PS3+, PS3+, PS2+, PS2, P3+ and P3. The fragments arising from P4 and P4+ were also investigated. The structures are discussed in terms of the Jahn—Teller effect, whose predictions are fulfilled without exception.  相似文献   

19.
《Polyhedron》1987,6(9):1717-1719
Determination of the stability constants for the system Co(II)/SO32−, was made potentiometrically using SO32−/HSO3 as buffers. Computation and matrix methods lead to the following overall stepwise constants: β1 = 4.3 x 102 M−1, β2 = 2.2 x 104 M−2 and β3 = 3.0 x 106 M−3. The partial constants (M−1) obtained were: K1 = 4.3 x 102, K2 = 51 and K3 = 1.4 x 102. The HSO3 competition was eliminated by using the extrapolation procedure.  相似文献   

20.
A series of five complexes that incorporate the guanidinium ion and various deprotonated forms of Kemp’s triacid (H3KTA) have been synthesized and characterized by single‐crystal X‐ray analysis. The complex [C(NH2)3+] ? [H2KTA?] ( 1 ) exhibits a sinusoidal layer structure with a centrosymmetric pseudo‐rosette motif composed of two ion pairs. The fully deprotonated Kemp’s triacid moiety in 3 [C(NH2)3+] ? [KTA3?] ( 2 ) forms a record number of eighteen acceptor hydrogen bonds, thus leading to a closely knit three‐dimensional network. The KTA3? anion adopts an uncommon twist conformation in [(CH3)4N+] ? 2 [C(NH2)3+] ? [KTA3?] ? 2 H2O ( 3 ). The crystal structure of [(nC3H7)4N+] ? 2 [C(NH2)3+] ? [KTA3?] ( 4 ) features a tetrahedral aggregate of four guanidinium ions stabilized by an outer shell that comprises six equatorial carboxylate groups that belong to separate [KTA3?] anions. In 3 [(C2H5)4N+] ? 20 [C(NH2)3+] ? 11 [HKTA2?] ? [H2KTA?] ? 17 H2O ( 5 ), an even larger centrosymmetric inner core composed of eight guanidinium ions and six bridging water molecules is enclosed by a crust composed of eighteen axial carboxyl/carboxylate groups from six HKTA2? anions.  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