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.
Electron flood guns used for charge compensation in secondary ion mass spectrometry (SIMS) cause chemical degradation. In this study, the effect of electron flood gun damage on argon cluster depth profiling is evaluated for poly(vinylcarbazole), 1,4-bis((1-naphthylphenyl)amino)biphenyl and Irganox 3114. Thin films of these three materials are irradiated with a range of doses from a focused beam of 20 eV electrons used for charge neutralization. SIMS chemical images of the irradiated surfaces show an ellipsoidal damaged area, approximately 3 mm in length, created by the electron beam. In depth profiles obtained with 5 keV Ar2000 + sputtering from the vicinity of the damaged area, the characteristic ion signal intensity rises from a low level to a steady state. For the damaged thin films, the ion dose required to sputter through the thin film to the substrate is higher than for undamaged areas. It is shown that a damaged layer is formed and this has a sputtering yield that is reduced by up to an order of magnitude and that the thickness of the damaged layer, which increases with the electron dose, can be as much as 20 nm for Irganox 3114. The study emphasizes the importance of minimizing the neutralizing electron dose prior to the analysis. Figure
?  相似文献   

2.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium Cs+(aq) + 1·Na + (nb) = 1·Cs+(nb) + Na+(aq) taking place in the two-phase water-nitrobenzene system (1 = lithium ionophore VIII; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex (Cs+, 1·Na+) = ?0.5 ± 0.1. Further, the stability constant of the 1·Cs+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb (1·Cs+) = 4.8 ± 0.2. Finally, by using quantum mechanical DFT calculations, the most probable structure of the cationic complex species 1·Cs+ was derived. In the resulting complex, the “central” cation Cs+ is bound by six bond interactions to the corresponding six oxygen atoms of the parent ligand 1.  相似文献   

3.
Both the singlet and triplet potential energy surfaces (PESs) of the NH (X3Σ?) + HCNO reaction have been investigated at the BMC-CCSD level based on the UB3LYP/6-311++G(d, p) structures. The results show that the title reaction is more favorable through the singlet potential energy surface than the triplet one. For the singlet potential energy surface of the NH (X3Σ?) + HCNO reaction, the most feasible association of NH (X3Σ?) with HCNO is found to be a non-barrier nitrogen-to-carbon attack forming the adduct a (trans-HNCHNO), which can isomerize to the adduct b (cis-HNCHNO). The most feasible channel is that the 1, 3-H shift with N2–H2 and C–N1 bonds cleavage associated with the N1–H2 bond formation of adduct a leads to the product P 1 (HCN + HNO). Moreover, P 2 (HNC + HNO) should be the competitive product. The other products, including P 3 (NH2 + NCO) and P 4 (N2H2 + CO), are minor products. The product P 1 can be obtained through two competitive channels Path 1: R  a  P 1 and Path 3: R  b  d  P 1 , whereas the product P 2 can be formed through Path 2: R  b  d  P 2 . At high temperatures, the nitrogen-to-nitrogen approach may become feasible. For the triplet potential energy surface of the NH (X3Σ?) + HCNO reaction, the Path 10: R  3 a  3 a 1  P 1 should be the most feasible pathway due to the less reaction steps and lower barriers. These conclusions will have impacts on further experimental investigations.  相似文献   

4.
From extraction experiments and $ \gamma $ -activity measurements, the exchange extraction constant corresponding to the equilibrium Tl+ (aq) + 1·Cs+ (org) ? 1·Tl+ (org) + Cs+ (aq) taking place in the two-phase water–phenyltrifluoromethyl sulfone (abbrev. FS 13) system (1 = calix[4]arene-bis(t-octylbenzo-18-crown-6); aq = aqueous phase, org = FS 13 phase) was evaluated as log K ex (Tl+, 1·Cs+) = 1.7 ± 0.1. Further, the extraordinarily high stability constant of the 1·Tl+ complex in FS 13 saturated with water was calculated for a temperature of 25 °C: log β org(1·Tl+) = 13.1 ± 0.2. Finally, by using quantum mechanical DFT calculations, the most probable structure of the cationic complex species 1·Tl+ was derived. In the resulting 1·Tl+ complex, the “central” cation Tl+ is bound by eight bond interactions to six oxygen atoms from the respective 18-crown-6 moiety and to two carbons of the corresponding two benzene rings of the parent receptor 1 via cation–π interaction.  相似文献   

5.
A method for depth profiling chromium in the surface and near surface regions of materials using the resonance at 1,005 keV in 52Cr(p,γ)53Mn nuclear reaction is presented. The detection sensitivity, depth resolution and probing depth of the resonance in Si are determined to be about 3 at.%, 25 nm and 2.5 µm respectively from the excitation function of the reaction constructed in 0.90–1.2 MeV proton energy region by measuring 378 keV prompt γ-rays from 53Mn nuclei. The reaction is interference free. These features make the approach attractive for profiling chromium in mid as well as high Z matrices.  相似文献   

6.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium Cs+(aq) + A?(aq) + 1(nb) ? 1·Cs+(nb) + A?(nb) taking place in the two-phase water–nitrobenzene system (A? = picrate, 1 = nonactin; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex (1·Cs+,A?) = 2.8 ± 0.1. Further, the stability constant of the 1·Cs+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb (1·Cs+) = 4.7 ± 0.1. Finally, by using quantum–mechanical DFT calculations, the most probable structure of the resulting cationic complex species 1·Cs+ was derived.  相似文献   

7.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium Cs+(aq) + I?(aq) + 1(nb) ? Cs+(nb) + I?(nb) taking place in the two–phase water–nitrobenzene system (1 = 1,3-alternate-25,27-bis(1-octyloxy)calix[4]arene-crown-6; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex (Cs+, I?) = 2.9 ± 0.1. Further, the stability constant of the Cs+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log βnb (Cs+) = 8.8 ± 0.1. Finally, by using quantum–mechanical DFT calculations, the most probable structure of the resulting cationic complex species Cs+ was derived.  相似文献   

8.
From extraction experiments and γ-activity measurements, the exchange extraction constant corresponding to the equilibrium Ag+ (aq) + Cs+(org) ? Ag+ (org) + Cs+ (aq) taking place in the two-phase water–phenyltrifluoromethyl sulfone (FS 13) system (1 = calix[4]arene-bis(t-octylbenzo-18-crown-6); aq = aqueous phase, org = FS 13 phase) was evaluated as logK ex (Ag+, Cs+) = ?1.5 ± 0.1. Further, the stability constant of the Ag+ complex in FS 13 saturated with water was calculated for a temperature of 25 °C: log β org(Ag+) = 10.1 ± 0.2. Finally, by using quantum mechanical DFT calculations, the most probable structure of the cationic complex species Ag+ was derived. In the resulting Ag+ complex, the “central” cation Ag+ is bound by eight bond interactions to six oxygen atoms from the respective 18-crown-6 moiety and to two carbons of the corresponding two benzene rings of the parent ligand 1 via cation-π interaction.  相似文献   

9.
From extraction experiments and γ-activity measurements, the exchange extraction constants corresponding to the general equilibrium M+ (aq) + 1·Na+ (org) $ \Leftrightarrow $ 1·M+ (org) + Na+ (aq) taking place in the two-phase water–phenyltrifluoromethyl sulfone (abbrev. FS 13) system (M+ = Li+, H3O+, NH4 +, Ag+, Tl+, K+, Rb+, Cs+; 1 = benzo-18-crown-6; aq = aqueous phase, org = FS 13 phase) were evaluated. Further, the stability constants of the 1·M+ complexes in FS 13 saturated with water were calculated; they were found to increase in the series of $ {\text{Cs}}^{ + } \, < \,{\text{Rb}}^{ + } \, < \,{\text{H}}_{ 3} {\text{O}}^{ + } \, < \,{\text{Ag}}^{ + } \, < \,{\text{Li}}^{ + } \, < \,{\text{NH}}_{4}^{ + } \, < \,{\text{K}}^{ + } \, < \,{\text{Tl}}^{ + } $ .  相似文献   

10.
Complexation of alkali metal cations with 5,11,17,23-tetra-tert-butyl-26,28,25,27-tetrakis(O-methyl-D-α-phenylglycylcarbonylmethoxy)calix[4]arene (L) was studied by means of spectrophotometric, conductometric and potentiometric titrations at 25 °C. The solvent effect on the binding ability of L was examined by using two solvents with different affinities for hydrogen bonding, viz. methanol and acetonitrile. Despite the presence of intramolecular NH···O=C hydrogen bonds in L, which need to be disrupted to allow metal ion binding, this calix[4]arene amino acid derivative was shown to be an efficient binder for smaller Li+ and Na+ cations in acetonitrile (lg K LiL  > 5, lg K NaL  = 7.66), moderately efficient for K+ (lg K KL  = 4.62), whereas larger Rb+ and Cs+ did not fit in its hydrophilic cavity. The complex stabilities in methanol were significantly lower (lg K NaL  =  4.45, lg K KL  = 2.48). That could be explained by different solvation of the cations and by competition between the cations and methanol molecules (via hydrogen bonds) for amide carbonyl oxygens. The influence of cation solvation on complex stability was most pronounced in the case of Li+ for which, contrary to the quite stable LiL + complex in acetonitrile, no complexation was observed in methanol under the conditions used.  相似文献   

11.
Aryl/pyridyl oxadiazole chromophores 6, 8 and 10, carrying N-phenyl aza-18-crown-6 have been synthesized as new photo-induced charge transfer (PCT) probes. While, the absorption spectra of the hosts experienced a slight negative solvatochromism, however the emission bands were dramatically red shifted (Stokes shifts up to 178 nm) in solvents of increasing polarity. Among the metal ions tested, Li+, Na+, K+ and Mg2+ did not appreciably perturbed the optical properties of the hosts. On the other hand, Ba2+ and to a lesser extent Ca2+ induced marked blue shifts in both the absorption and emission spectra of the hosts. The magnitude of cation induced spectral blue shifts corresponded with the increasing acceptor strength of the attached aryl/pyridyl groups in the host molecules. The blue shifts and the stability constants were found to follow the order Ba2+ > Ca2+ ? Mg2+ > Na+ > Li+ > K+. Competitive experiments performed with a matrix of ions also revealed superior binding affinity of Ba2+ with all the hosts examined. Noteworthily, the deep yellow solution (λmax, 386 nm) of the host 10 was completely bleached (λmax, 320 nm), in the company of Ba2+ thereby allowing the naked eye detection of this ion.  相似文献   

12.
In X-ray photoelectron spectroscopy (XPS) Ar+ ion sputtering is usually used for depth profiling. However, for such samples as organic coatings, this is not feasible because of degradation. Also, measurement of a depth profile on a conventionally prepared cross-section is not possible if, for example, sample thickness is below the smallest available measurement spot size of the XPS system. In our approach we used a rotary microtome to cut samples under a shallow tilting angle of 0.5° to obtain an extended cross-section suitable for XPS investigations. We also used liquid nitrogen cooling to ensure an exposed area of higher quality: topography measurements with a novel optical 3D microscope and by atomic force microscopy revealed the linearity of the inclined sections. With our cryo ultra-low-angle microtomy (cryo-ULAM) preparation technique we were able to determine, by XPS, elemental and chemical gradients within a 25 μm thick polyester-based organic coating deposited on steel. The gradients were related to, for example, depletion of the crosslinking agent in the sub-surface region. Complementary reflection electron energy-loss spectroscopy measurements performed on the cryo-ULAM sections also support the findings obtained from the XPS depth profiles.
Figure
Top view of the sample as imaged in 3D by FF-OCM is shown in a). In b) a side view of the 3D sample structure simultaneously exhibiting the coating surface and substrate-coating interface is depicted. Compositional XPS-depth profile of the coating prepared by cryo-ULAM is shown in c)  相似文献   

13.
The first examples of two d 10 metal coordination polymers based on 1,5-naphthalenedisulfonic acid and benzimidazole, namely, [Cd(bim)4(1,5-nds)] 1 and [Ag(bim)2]·½1,5-nds 2 (bim = benzimidazole, 1,5-nds = 1,5-naphthalenedisulfonic acid) were synthesized successfully under solvothermal conditions. Their structures were determined by single-crystal X-ray diffraction analyses and further characterized by elemental analyses, IR, TG-DSC, PXRD, UV–Vis, and DFT. 1 exhibits a 3D supramolecular structure which started from 1D helical chain. 2 also shows a 3D architecture which was assembled via π–π interactions and hydrogen bonds. The hydrogen bonds formed by 1,5-nds play an important role in increasing the dimensionality of architectures of 1 (from 1D to 3D) and 2 (from 0D to 1D). Moreover, both 1 and 2 show excellent luminescent properties, 335 nm for 1 (λ ex = 286 nm) and 341 nm and 389 nm for 2 (λ ex = 234 nm), and may be the suitable candidates of fluorescent materials. From UV–Vis spectra, the remarkable absorption peaks can be found at 276 nm for 1 and 271 nm for 2. The simulated UV–Vis spectra (290.30 nm for 1 and 271.64 nm for 2) by Gaussian 09 DFT correspond well with the experimental results. From the electron density distribution of frontier molecular orbitals, it is known that the two absorption bands are attributed to the intra-ligand electron transition from HOMO?26 to LUMO and from HOMO?3 to LUMO+1 for 1, from HOMO?1 to LUMO+18 and LUMO+19 for 2.  相似文献   

14.
Non-aqueous reactions of aluminum isopropoxide with 8-hydroxyquinoline (Hq = HONH6C9) in 1:1, 1:2 and 1:3 molar ratios in anhydrous benzene yield complexes of the type [qnAl(OPri)3?n] {where n = 1 (1), n = 2 (2), n = 3 (3)}. Progress of the reactions were monitored by estimating liberated 2-propanol in benzene-2-propanol azeotrope by oxidimetric method. All the products were fluorescent green powders, sparingly soluble in CHCl3. They were characterized by elemental analysis, FT-IR and (1H, 13C and 27Al) NMR studies. The ESI mass spectral studies indicate dimeric nature for (1) and (2) and monomeric nature for the compound (3). The XRD spectra of (13) showed crystalline nature with the average particle size of 45, 32 and 27 nm respectively, as evaluated from DebyeScherrer equation. The XRD spectrum of (3) also suggests the formation of β-crystalline polymorphs of Alq3. The SEM images appear to indicate granular morphology for (1) and formation of cylindrical shaped rods for (2) and (3). Sol–gel hydrolysis of (1), (2) or (3) in presence of a strong acid as well as of the precursor, Al(OPri)3,without acid or base catalyst, followed by sintering at 950 °C yielded tetragonal primitive phase of nano-sized δ-alumina in all the cases, as reflected by their powder X-ray diffraction pattern. The IR, SEM and EDX studies also support the formation of transition alumina.  相似文献   

15.
From extraction experiments and γ-activity measurements, the exchange extraction constant corresponding to the equilibrium Ca2+(aq) + 1·Sr2+(nb) ? 1·Ca2+(nb) + Sr2+(aq) taking place in the two-phase water–nitrobenzene system (1 = beauvericin; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex(Ca2+, 1·Sr2+) = 1.1 ± 0.1. Further, the stability constant of the 1·Ca2+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb(1·Ca2+) = 10.1 ± 0.2. Finally, by using quantum mechanical density functional level of theory calculations, the most probable structures of the non-hydrated 1·Ca2+ and hydrated 1·Ca2+·H2O complex species were predicted.  相似文献   

16.
The radical-molecule reaction of C2Cl3 with NO2 is explored at the B3LYP/6-311G(d,p) and CCSD(T)/6-311+G(d,p) (single-point) levels. On the singlet potential energy surface (PES), the association between C2Cl3 and NO2 is found to be carbon-to-nitrogen attack forming the adduct C2Cl3NO2 (1) without any encounter barrier, followed by isomerization to C2Cl3ONO (2). Starting from 2, the most feasible pathway is the N–O1 bond cleavage which lead to P 1 (C2Cl3O + NO). Much less competitively, 2 transforms to the three-membered ring isomer c-OCCl2C–ClNO (4 a ) which can easily interconvert to c-OCCl2C–ClNO 4 b . Then 4 (4 a , 4 b ) takes direct C1–C2 and C2–O1 bonds cleavage to give P 2 (COCl2 + ClCNO). The lesser competitive channel is the 4 a isomerizes to the four-membered ring intermediate O-c-CNClOCCl2 (5) followed by dissociation to P3 (CO + ClNOCCl2). The concerted 1,2-Cl shift along with C1–O1 bond rupture of 4 b to form ONC(O)CCl3 (6) followed by dissociation to P 4 (ClNO + OCCCl2) is even much less feasible. Moreover, some of P 3 and P 4 can further dissociate to P 5 (ClNO + CO + CCl2). Compared with the singlet pathways, the triplet pathways may have less contribution to the title reaction. Our results are in marked difference from previous theoretical studies which showed that two initial adducts C2Cl3–NO2 and C2Cl3–ONO are obtained. Moreover, in the present paper we focus our main attentions on the cyclic isomers in view of only the chain-like isomers are considered by previous studies. The present study may be helpful for understanding the halogenated vinyl chemistry.  相似文献   

17.
The complex triplet potential energy surface for the reaction of HCNO with NH is investigated at the G3B3 level using the B3LYP/6-311++G(d,p), and QCISD/6-311++G(d,p) geometries. Various possible isomerization and dissociation pathways are probed. The initial association between HCNO and NH is found to be carbon to nitrogen attack leading to HNCHNO 2a, which can convert to 2b, 2c, and 2d. Subsequently, 1,4-H-shift of 2a to form NCHNOH 3a followed by dissociation to P 2 (1HCN + 3HON) is the most feasible pathway. Much less competitively, 2d undergoes successive 1,3-H-shift and C-N cleavage to form HNCNOH 8b, and then to product P 3 (1HNC + 3HON), the second feasible pathway. 8b can alternatively isomerize to 8c followed by N–O bond rupture to generate P 6 (2OH + 2HNCN), the lesser followed feasible pathway. In addition, 2b takes continuously 1,3- and 1,2-H-shift to form NC(H)NHO 6a, then to ONHCNH 7a which can convert to 7b. Eventually, 7b may take C-N bond fission to produce P 5 (1HNC + 3HNO), the least feasible pathway. The present paper may be helpful for future experimental identification of the product distributions for the title reaction, and may be helpful to deeply understand the mechanism of the title reaction.  相似文献   

18.
By using extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium Sr2+(aq) + 2A?(aq) + 1(nb) ? 1·Sr2+(nb) + 2A?(nb) occurring in the two-phase water–nitrobenzene system (A? = picrate, 1 = antamanide; aq = aqueous phase, nb = nitrobenzene phase) was determined as log K ex (1·Sr2+, 2A?) = ?0.3 ± 0.1. Further, the stability constant of the 1·Sr2+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb (1·Sr2+) = 8.8 ± 0.1. Finally, applying quantum mechanical density functional level of theory calculations, the most probable structure of the cationic complex species 1·Sr2+ was derived. In the resulting complex, the “central” cation Sr2+ is bound by six bond interactions to the corresponding six oxygen atoms of the parent ligand 1. The interaction energy of the considered 1·Sr2+ complex was found to be ?1,114.9 kJ/mol, confirming the formation of this cationic species as well.  相似文献   

19.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium Sr2+(aq) + 2A?(aq) +1(nb) ? 1·Sr2+(nb) + 2A?(nb) taking place in the two-phase water–nitrobenzene system (A? = picrate, 1 = beauvericin; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex(1·Sr2+,2A?) = ?0.6 ± 0.1. Further, the stability constant of the 1·Sr2+ complex in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb(1·Sr2+) = 8.5 ± 0.1. Finally, by using quantum-mechanical DFT calculations, the most probable structure of the resulting cationic complex 1·Sr2+ was derived.  相似文献   

20.
The exchange extraction constants corresponding to the general equilibrium C+(aq) + Cs+(nb) ? C+ (nb) + Cs+(aq) occurring in the two-phase water–nitrobenzene system (C+ = protonated α-amino acid methyl ester, 1 = hexaethyl p-tert-butylcalix[6]arene hexaacetate; aq = aqueous phase, nb = nitrobenzene phase) were evaluated on the basis of extraction experiments and γ-activity measurements. Further, the stability constants of the C+ cationic complex species in nitrobenzene saturated with water were calculated; they were found to increase in the following cation order: protonated l-tryptophan methyl ester < protonated l-phenylalanine methyl ester < protonated l-leucine methyl ester < protonated l-methionine methyl ester < protonated l-valine methyl ester.  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