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.
The betaines1b–d were prepared4 by systematic variation of the alkyl groups and were reacted with trifluoroacetic acid anhydride (TFA) to give the diacyl-ylides2b,c. The betain1d andTFA afford the trifluoroacetate3d 5. The salts3b,c, which result from hydrolysis of2b,c as well as3d (X=I) can be transformed in 75 to 83% yield into the monoacyl-ylides4b–d with the help of silver oxide. Aqueous solutions of4a–d exhibit alkalinepH, which points to the formation of the corresponding ammonium bases. In the case of4b,c the bases5b,c could be isolated. It can be shown, that4b,c and5b,c, respectively, undergo a reversible addition or elimination of one mole wather with great ease.  相似文献   

2.
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   

3.
(E)-11H-Bisbenzo[a]fluorenylidene (E-6) was synthesized by Barton’s double extrusion diazo-thione coupling method from 11H-benzo[a]fluoren-11-thione (11) and 11-diazo-11H-benzo[a]fluorene (13). The reaction is probably thermodynamically controlled; in the event that the less stable Z -6 is also formed, it would rapidly undergo Z → E diastereomerization to give E -6. The B3LYP/6-311G(d,p) calculated diastereomerization barrier for Z -6 → E -6 is ΔG 298 = 57.0 kJ/mol (13.6 kcal/mol). The calculated equilibrium constant K eq(E -6 → Z -6) = 92:8 (at 298 K) is indicative of a marked diastereoselectivity of the reaction leading to E -6. The structure of E-6 was established by 1H-NMR and 13C-NMR spectroscopies and by X-ray analysis. PAE E-6 crystallizes in the monoclinic space group C2/c. The unit cell of the crystal structure E -6 contains eight molecules, arranged as four pairs of enantiomers. PAE E -6 adopts a twisted conformation with the pure twist of the central C11=C11′ bond ω = 39°. The dihedral angle ν in E -6 is 60.6°, which is significantly higher than the respective dihedral angle in PAEs Z -6, 2, E -7, Z -7, 14, and 15. The large syn-pyramidalization angles at C11 and C11′ (χ = 12.6° and 14.8°) of E-6 indicates the enhanced strain in the fjord regions of the molecule. The enhanced twist is primarily attributed to the double benzo[a]annelation of the bifluorenylidene moiety at the fjord regions. The B3LYP/6-311G(d,p) calculated structure of E -6 is in a very good agreement with the experimental X-ray structure. PAE E -6 adopts a twisted conformation in solution, with the downfield chemical shift of H1/H1′ (8.31 ppm); H10/H10′ (δ = 7.20 ppm) and H9/H9′ (δ = 6.86 ppm) in E -6 are positioned above the planes of the opposing naphthalene rings. PAEs E -6 and Z -6 are significantly higher in energy than their corresponding benzo[b]annelated isomers E -7 and Z -7.  相似文献   

4.
A high-performance liquid chromatographic method has been developed for the simultaneous analysis of the flavonols myricitrin (1), avicularin (2), and juglanin (3) in rat plasma and urine after oral administration of the total flavonoids from Polygonum aviculare. Samples were prepared by solid-phase extraction then separated on a C18 reversed-phase column by use of a mobile-phase gradient prepared from methanol and aqueous formic acid solution. The flow rate was 1 mL min?1. Detection was performed at 254 nm. The calibration range was 11–1,100 μg mL?1 for both 2 and 3 in plasma; in urine the calibration ranges for 1, 2, and 3 were 32–1,600, 11–1,100, and 22–1,100 μg mL?1, respectively. Intra-day and inter-day RSD were less than 4.33 and 3.62% for 2 and 3, respectively, in plasma, and no more than 4.03 and 2.22% for all the analytes in urine. The analytical sensitivity and selectivity of the assay enabled successful application to pharmacokinetic studies of flavonols 13 in rats.  相似文献   

5.
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   

6.
Three novel thiaazacrown ethers 1, 2 and 3 were synthesized in a simple way and in high yield. The complex formation between Ag+, Cu2+, Zn2+, Pb2+, Hg2+ and Cd2+ metal cations with thiaazacrown ethers 1, 2 and 3 have been studied in acetonitrile:chloroform (1:1) binary solvent system using conductometric technique. The conductance data show that the stochiometry of the complexes with Ag+, Cu2+ and Zn2+ cations is 1:1 (L:M), but in the case of Pb2+ and Hg2+ cations, a 1:2 (L:M) complex is formed in solutions. The formation constants of the resulting 1:1 complexes were determined from the molar conductance-mole ratio data at 25 °C. It was found that the stability constants of 1-Ag2+, 2-Ag+ and 3-Ag+ complexes are higher than those of their corresponding Zn2+ and Cu2+ complexes and found to vary in order 2 for Ag+.  相似文献   

7.
Novel comparison of the structural, electronic and energy aspects of lanthanide complexes of model phosphoramides (PAs) with those of phosphine oxides (POs), phosphate esters (PEs) and phosphoryl trihalides (PHs) has been carried out by ab initio and DFT calculations. Atoms in Molecules (AIM) and Natural Bonding Orbital (NBO) analyses were performed to understand the electronic structure of ligands L and related complexes, L–Ln3+. NBO analysis indicates that the negative charge on phosphoryl oxygen (OP) and the p character of the phosphoryl lone pair, Lp(OP), increase in the order PH < PE < PO < PA. Positive charge of the lanthanide cation in PA complexes is less than those of PH, PE and PO complexes, due to the more intense ligand to metal charge transfer (LMCT). The metal–ligand distance decreases in the order PH > PE > PO > PA, which is confirmed by the results of AIM analysis. Charge density at the bond critical point of L–Ln3+ follows the sequence PH < PE < PO < PA. The results of the Energy Decomposition Analysis (EDA) indicate that the donative interaction and LMCT increases in order PH < PO < PE < PA. The effect of basis set superposition error (BSSE) on the L···Ln3+ interaction energies was also studied in detail at DFT, MP2 and CCSD(T) levels using the counterpoise (CP) method. Trends in the CP-corrected L–Ln3+ bond energies are in good accordance with the optimized OP···Ln3+ distances. The results show that the difference between CP-corrected and uncorrected interaction energies in PA complexes is larger than those in the others, because PAs are more deformable. It is depicted that PAs are comparable with POs in lanthanide complexation.  相似文献   

8.
The aim of this work was to synthesize deuterated and tritiated analogues of highly oxidized lupane derivatives known from our group. We selected compounds that previously showed very high cytotoxic activity on multiple cancer cell lines in order to further investigate the mechanism of their action. From starting material (compounds 14), we obtained benzyl platanate (5) and its reaction with deuteromethyltriphenylphosphonium iodide gave deuterated compound 6. Following benzyl deprotection gave free acid 7 and oxidation with SeO2 gave 30-oxo-[29-2H2]lup-20(29)-en-28-oic acid (8), which is one of the most active compounds synthesized in our group to date (IC50 6 μmol/L on CEM cell line). The alkylation of benzyl 2-hydroxy-3-oxolupa-1,20(29)-dien-28-oate (9) with methyliodide or deuteromethyliodide followed by a series of deprotection and hydrogenation steps gave compounds 1014, where 2β-[31-2H3]methoxy-3-oxolupan-20(29)-en-28-oic acid (13) is especially interesting, it showed lower activity on CEM cell line (IC50 10 μmol/L) however, it was very active against Ph1—positive human leukemia BV-173 (IC50 0.91 μmol/L) and against human myelogenous leukemia K562 (IC50 0.52 μmol/L). Selectively labelled [3α-2H] and [3α-3H] methyl 3β-acetoxy-21,22-dioxolup-18-en-28-oates 24, 25 were prepared in three steps by reduction of corresponding 3-oxo derivatives and they showed moderate activity on CEM cell line (IC50 10 μmol/L). In total, 11 labelled compounds (68, 11, 14, 18, 19, 21, 22, 24 and 25) have not been reported before.  相似文献   

9.
From extraction experiments and $ \gamma $ -activity measurements, the extraction constants corresponding to the general equilibrium Eu3+(aq) + 3 A?(aq) + L(nb) $ \Leftrightarrow $ EuL3+(nb) + 3A?(nb) taking place in the two-phase water–nitrobenzene system ( $ {\text{A}}^{ - } = {\text{CF}}_{ 3} {\text{SO}}_{3}^{ - } $ ; L = electroneutral receptors denoted by 1, 2, and 3 – see Scheme 1; aq = aqueous phase, nb = nitrobenzene phase) were evaluated. Further, the stability constants of the EuL3+ complexes in nitrobenzene saturated with water were calculated; they were found to increase in the series of 3 < 2 < 1.
Scheme 1
Structural formulas of N,N,N′,N′,N″,N″-hexacyclohexyl-4,4′,4″-propylidynetris(3-oxabutyramide) (1), bis[(12-crown-4)methyl] dodecylmethylmalonate (2), and bis[(benzo-15-crown-5)-4′-ylmethyl] pimelate (3)  相似文献   

10.
From extraction experiments and $ \gamma $ -activity measurements, the exchange extraction constants corresponding to the general equilibrium M2+(aq) + Sr2+(nb) $ \Leftrightarrow $ M2+(nb) + Sr2+(aq) taking place in the two-phase water–nitrobenzene system (M2+ = Mg2+, Ca2+, Ba2+, Pb2+, Cu2+, Zn2+, Cd2+, $ {\hbox{UO}}_{2}^{2 + } $ , Mn2+, Co2+, Ni2+; 1 = macrocyclic lactam receptor–see Scheme 1; aq = aqueous phase, nb = nitrobenzene phase) were evaluated. Moreover, the stability constants of the M2+ complexes in nitrobenzene saturated with water were calculated; they were found to increase in the following cation order: Mg2+ < Co2+ < Cu2+, Mn2+, Ni2+ < Cd2+ < Ca2+ < Ba2+, Zn2+ < Pb2+ <  $ {\hbox{UO}}_{2}^{2 + } $ .
Scheme 1
Structural formula of 2,18-dichloro-9,10,11,12-tetrahydro-6H, 20H-dibenzo[l,o][1,11,4,8]dioxadiazacyclohexadecine-7,13(8H, 14H)-dione (abbrev. 1)  相似文献   

11.
4-Alkylaminopyridinethiones · HCl (1 · HCl) react with bis-trichlorethylmalonate (3) predominantly to 5-alkylamino-4H-thiopyrano [2,3-b]pyridine-4-ones (6). With alcohols in the presence of acids at 25°C6 undergoes an alcoholysis to the corresponding alkyl-3-(2-thioxo-3-pyridyl)propionates (9). On heating in dilute alkali6 is hydrolysed via 4-alkylamino-2-thioxopyridyl-propylketones (11) to the tautomers, 4-hydroxy-2-thioxopyridylpropylketone (12 A) and 2-thioxo-3-(1-hydroxybutenyl)-4-piperidon (12 B), resp. On refluxing with alkali the ethyl-pyridylpropionate9 a is cyclisized to the 1-alkyl-1,6-naphthyridine-2(1H)-one (4 a), but boiling in ethanolic acid hydrolyses9 a via the pyridylpropionic acid10 to 4-alkyl-aminopyridylpropylketone (11 a). The latter can be transformed via the tautomers12 A,B and 2-methylthio-3-pyridylpropylketone (13) to the 4-hydroxy-3-butyrylpyridone (14 A) and its tautomer, 3-(1-hydroxy-butenyl)-piperidine-2,4-diones (14 B) resp. The structure of14 A,B is established by reaction of 4-isopropylamino-2(1H)-pyridone (2) with butanoylchloride to the 4-isopropylamino-3-butyrypyridone (15) and hydrolysis of15 to the tautomers14 A,B.  相似文献   

12.
The interaction between chemosensor, N-(2-methyl-1,3-dioxo-indan-5-yl)-benzamide (1) and different halide ions (F ? , Cl? and Br?) has been investigated using density functional theory (DFT). A clear insight of the sensor anion binding process has been presented. Our calculations revealed that the observed colorimetric and fluorescent signals are induced due to the ground state deprotonation of the sensor molecule caused by F? which has two times higher binding affinity than other halide ions (Cl? and Br?). Derivatives of system 1 have been made to find a better sensor with higher binding affinity and longer wavelength of absorption. All the derivatives are better sensors than the parent 1 except 4-methyl-N-(2-methyl-1,3-dioxo-indan-5-yl)-benzamide (2). Among these derivatives, trimethyl-[4-(2-methyl-1,3-dioxo-indan-5-ylcarbamoyl)-phenyl]-ammonium (8) and (5-benzoylamino-1,3-dioxo-indan-2-yl)-trimethyl-ammonium (9) showed a change to higher binding energies of about 58 Kcal/mol and longer absorption wavelengths of 53 nm after deprotonation process than the parent system 1 which is highly demanded in selective chemical sensing. Systems 8, 9 and their deprotonated zwitterionic forms (8z and 9z) have also been studied for their nonlinear optical responses. Systems 8, 9 showed significantly good first hyperpolarizability (β) of 84 × 10?30 and 40 × 10?30 esu, respectively. These β values increase in zwitterionic states up to 216 × 10?30 and 109 × 10?30 esu, respectively after deprotonation with F?, representing a new signal of deprotonation.  相似文献   

13.
The effect of energetic γ-radiation on 1H NMR, electronic absorption, ESR spectra, differential thermal analysis (DTA) and solid state dc electrical conductivity of the ligand N-phenyl-2-(2-(phenylamino)acetyl)hydrazine carbothioamide (H2L) and its copper(II) complexes; Cu(HL)(OAc)H2O, Cu(HL)BrH2O and Cu(H2L)2(NO3)2?3H2O before and after γ-irradiation (hereafter referred to as (B), (B 1 ), (B 2 ), (B 3 ) and (A), (A 1 ), (A 2 ), (A 3 ), respectively) has been studied. Electronic spectral bands of the complexes after irradiation exhibited some better resolved shapes with a remarkably higher absorbance, ESR spectrum of complex Cu(HL)BrH2O (B 2 ) before irradiation showed isotropic spectrum with g iso = 2.075 however, after irradiation (A 2 ) displayed axial ESR spectrum with g  > g  > 2.0023 and d (x2?y2) ground state. DTA of the compounds reveals that γ-irradiation induced generation of new peaks as well as changes in the peak intensities. Solid state dc electrical conductivity for complexes was investigated before and after γ-irradiation. Complexes were found to be semiconductors, the activation energies (E a) were calculated for the complexes by using the Arrhenius plot.  相似文献   

14.
A glycol ether modified precursor, [Nb{O(CH2CH2O)2}(OPri)3] (A) was prepared by the reaction of Nb(OPri)5 with O(CH2CH2OH)2 in 1:1 molar ratio in anhydrous benzene. Further reactions of A with a variety of internally functionalized oximes in different molar ratios, yielded heteroleptic complexes of the type, [Nb{O(CH2CH2O)2}(OPri)3?n{ON = C(CH3)(Ar)}n] (1–9) {where Ar = C4H3O-2, n = 1 [1], n = 2 [2], n = 3 [3]; C4H3S-2, n = 1 [4], n = 2 [5], n = 3 [6]; C5H4N-2, n = 1 [7], n = 2 [8], n = 3 [9]}. All the above derivatives have been characterized by elemental analyses, FT-IR, NMR (1H, 13C {1H}) and FAB mass studies. Spectral studies of 1–9 suggest the presence of mono- and bi-dentate mode of oxime moieties, in the solution and in the solid states, respectively. FAB mass studies indicate monomeric nature for 3 and dimeric nature for A. TG curves of A and 6 show their low thermal stability. Soft transformation of A and 3 to pure niobia, a and b, respectively have been carried out by sol–gel technique. The XRD patterns of niobia a and b suggest the formation of nano-size crystallites of average size of 10.8 and 19.5 nm, respectively. The XRD patterns also indicate the formation of monoclinic phase of the niobia in both the cases. Absorption spectra of a and b suggest energy band gaps of 4.95 and 4.39 eV, respectively.  相似文献   

15.
Mononuclear copper(II) complexes of 1,2,4-triazole-based Schiff base macrocyclic hydrazones, III and IV, have been reported. The prepared amorphous complexes have been characterized by spectroscopic methods, electron spray ionization mass spectrometry, and elemental analysis data. Electrochemical studies of the complexes in DMSO show only one quasi-reversible reduction wave at +0.43 V (ΔE = 70 mV) and +0.42 V (ΔE = 310 mV) for III and IV, respectively, which is assigned to the Cu(II) → Cu(I) reduction process. Temperature dependence of magnetic susceptibilities of III and IV has been measured within an interval of 2–290 K. The values of χM at 290 K are 1.72 × 10?3 cm3 mol?1 and 1.71 × 10?3 for III and IV, respectively, which increases continuously upon cooling to 2 K. EPR spectra of III and IV in frozen DMSO and DMF were also reported. The trend g|| > g⊥ > ge suggests the presence of an unpaired electron in the dx2?y2 orbital of the Cu(II) in both complexes. Furthermore, spectral and antimicrobial properties of the prepared complexes were also investigated.  相似文献   

16.
Two βCD dimers (linked by succinic acid, 2, or ethylenediaminetetraacetic acid, EDTA, 3, bridges) and a negatively charged monomer derivative of βCD, 1, have been synthesized and their ability to solubilize cholesterol in aqueous solution was studied. The three compounds exhibit a great capacity in solubilizing cholesterol as, for instance, concentrations up to 6 mM of cholesterol were measured in the presence of 25 mM of 3. The phase-solubility diagrams of the two dimers exhibit A L type profiles while the monomer 1 follows an A P isotherm. The cholesterol/dimer complexes have 1:1 stoicheiometries while monomer 1 forms two complexes with molar ratios of 1:1 and 1:2 (cholesterol/1). The equilibrium constants are K 1:1 = (5.9 ± 0.3) × 104 M?1 and K 1:1 = (8.8 ± 0.2) × 104 M?1 for 2 and 3, respectively, and K 1:1 = 73 ± 19 M?1 and K 1:2 = 204 ± 65 M?1 for 1. The comparison of K 1:1(3) with the product K 1:1 × K 1:2 (1) reveals that a chelate effect in binding the cholesterol by 3 exists. The structure of the cholesterol/3 complex was studied by ROESY experiments and by molecular dynamics simulations.  相似文献   

17.
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   

18.
Four phenylpropanoids, (E)-p-coumaryl alcohol (1), 3,4-dihydroxycinnamyl alcohol (2), sachaliside 1 (3), and coniferin (4) have been isolated from the rhizome of Pinellia ternata. Their structures were elucidated by spectroscopic methods. Compounds 13 were isolated from the genus Pinellia for the first time. Compound 4 was isolated from this plant for the first time. A rapid, sensitive, and accurate reversed-phase high-performance liquid chromatographic method with UV detection at 260 nm was established for simultaneous separation and determination of the four phenylpropanoids in nineteen batches of dried rhizomes of P. ternata. Compounds were separated on a 250 × 4.6 mm C18 column with methanol–acetonitrile–water–phosphoric acid, 20:5:75:0.3, as mobile phase. The amounts of 14 in the rhizome of P. ternata could be easily determined within 30 min. The linear calibration ranges for 14 were 0.05–137.50, 0.66–1050.00, 0.06–30.00, and 0.05–67.50 μg mL?1, respectively. Recovery of 14 was 97.43–103.73%, with RSD from 0.12 to 1.62%. Limits of quantification for 14 were 50, 660, 60, and 50 ng mL?1, respectively. The method was successfully used for phytochemical analysis of phenylpropanoids from the rhizome of P. ternata.  相似文献   

19.
A series of acyclic Schiff base podands 14?C19 with lipophilic amide and ester end groups were synthesized in good yield and in a simple way. Their transition metal ions complexation was studied using conductometric method in acetonitrile (AN) at 25 °C. Schiff base podands 14?C16 showed a continuous decrease in the molar conductances in their complexation with Hg2+, Pb2+, Cu2+, Zn2+ and Cd2+ which begins to level off at a mole ratio of 1:1 crown-to-metal indicating the formation of a stable 1:1 complexes. The order of the stability constants of the metal ions studied with the Schiff base podands 14, 15 and 16 is: Hg2+ > Pb2+ > Cu2+ > Zn2+ > Cd2+ > Ag+. Metal ion complexation by acyclic diamide or diester podands involves presumably the oxygen atoms of the carbonyl groups in addition to the nitrogen atoms of the imino groups.  相似文献   

20.
In light of the very recent significant discrepancies on the global isomer of the sept-atomic molecule OB6, we performed a detailed potential energy surface survey of OB6 covering various isomeric forms. We showed that at the CCSD(T)/6-311+G(2df)//B3LYP/6-311+G(d) level, the planar knife-like isomer 01 with a –BO moiety has the lowest energy, followed by the planar belt-like isomer 02 at 22.6 kcal/mol. Another isomer 05 at 33.1 kcal/mol can be viewed as the direct O-adduct of the pentagonal pyramid B6. Kinetically, the three isomers 01, 02 and 05 all have considerable barriers (19–29 kcal/mol) (obtained at B3LYP/6-311+G(d) level) against isomerization. However, other isomers either have much higher energy or possess much smaller conversion barriers and are thus of little likeliness for isolation. Moreover, though being isoelectronic to the well-known CB 6 2? molecule, OB6 does not have any kinetically stabilized wheel-like isomers with O or B centers. The three OB6 isomers 01, 02 and 05 await future laboratory studies. The detailed results reported in this paper are expected to provide useful information for understanding the growing process of boron oxides, O-doping and oxidation mechanism of boron clusters.  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