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 484 毫秒
1.
Covalent functionalization of a zigzag boron nitride nanotube (BNNT) with acetylene has been investigated by density functional theory in terms of energetic, geometric, and electronic properties. It has been found that the most stable functionalized BNNT is the one in which the acetylene is diffused into the tube wall so that two heptagonal and two pentagonal rings are formed, releasing energy of 1.54 eV. In addition, the effect of substituting the hydrogen atoms of C2H2 by different functional groups including –F, –CH2F, –CN, and –OCH3 on the geometric and electronic properties of the BNNT has been investigated. The reaction energies are found to be in the range of ?1.03 to ?3.13 eV so that their relative magnitude order is as follows: C2F2 > (OCH3)2C2 > C2H2 > (CH2F)2C2 > (CN)2C2, suggesting that the functionalization energy is increased by increasing the electron donating character of the functional groups. Overall, chemical modification of BNNT by the studied groups results in little changes in electronic properties of the tube and may be an effective way for the purification of BNNTs.  相似文献   

2.
The kinetics of oxidation of a series of substituted 4-oxobutanoic acids (Y–C6H4COCH2CH2COOH: Y = H, OCH3, CH3, C6H5, Cl, Br or NO2) by N-bromophthalimide have been studied in aqueous acetic acid medium at 30 °C. The total reaction is second-order, first-order each in oxidant and substrate. The oxidation rate increases linearly with [H+], establishing the hypobromous acidium ion, H2O+Br, as the reactive species. A variation in ionic strength has no effect on the reaction rate. The order of reactivity among the studied 4-oxoacids is: 4-methoxy > 4-methyl > 4-phenyl > 4-H > 4-Cl > 4-Br > 3-NO2. The effect of changes on the electronic nature of the substrate reveals that there is a development of positive charge in the transition state. The activation parameters have been computed from Arrhenius and Eyring plots. Based on the kinetic results, a suitable mechanism has been proposed.  相似文献   

3.
Signed values of all intra‐ring 2,3,4J(C,C) couplings in nine monosubstituted benzenes (C6H5‐X where X = F, Cl, Br, CH3, OCH3, Si(CH3)3, C ≡ N, NO, NO2) are experimentally determined as well as nine couplings to substituent carbons. It is confirmed that while all the vicinal intra‐ring 3J(C,C) are positive and all geminal 2J(C2,C4) are negative, both signs are found for geminal 2J(C1,C3) couplings. All the determined signs agree with those already predicted by theoretical calculations. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
To enable a comparison between a C—H…X hydrogen bond and a halogen bond, the structures of two fluorous‐substituted pyridinium iodide salts have been determined. 4‐[(2,2‐Difluoroethoxy)methyl]pyridinium iodide, C8H10F2NO+·I, (1), has a –CH2OCH2CF2H substituent at the para position of the pyridinium ring and 4‐[(3‐chloro‐2,2,3,3‐tetrafluoropropoxy)methyl]pyridinium iodide, C9H9ClF4NO+·I, (2), has a –CH2OCH2CF2CF2Cl substituent at the para position of the pyridinium ring. In salt (1), the iodide anion is involved in one N—H…I and three C—H…I hydrogen bonds, which, together with C—H…F hydrogen bonds, link the cations and anions into a three‐dimensional network. For salt (2), the iodide anion is involved in one N—H…I hydrogen bond, two C—H…I hydrogen bonds and one C—Cl…I halogen bond; additional C—H…F and C—F…F interactions link the cations and anions into a three‐dimensional arrangement.  相似文献   

5.
The melting temperature, melting enthalpy, and specific heat capacities (C p) of 5′-deoxy-5′-iodo-2′,3′-O-isopropylidene-5-fluorouridine (DIOIPF) were measured using DSC-60 Differential Scanning Calorimetry. The melting temperature and melting enthalpy were obtained to be 453.80 K and 33.22 J g?1, respectively. The relationship between the specific heat capacity and temperature was obtained to be C p/J g?1 K?1 = 2.0261 – 0.0096T + 2 × 10?5 T 2 at the temperature range from 320.15 to 430.15 K. The thermal decomposition process was studied by the TG–DTA analyzer. The results showed that the thermal decomposition temperature of DIOIPF was above 487.84 K, and the decomposition process can be divided into three stages: the first stage is the decomposition of impurities, the mass loss in the second stage may be the sublimation of iodine and thermal decomposition process of the side-group C4H2O2N2F, and the third stage may be the thermal decomposition process of both the groups –CH3 and –CH2OCH2–. The obtained thermodynamic basic data are helpful for exploiting new synthetic method, engineering design, and commercial process of DIOIPF.  相似文献   

6.
Abstract

Electrophilic trisubstituted ethylenes, ring‐substituted ethyl 2‐cyano‐3‐phenyl‐2‐propenoates, RC6H4CH?C(CN)CO2C2H5 (where R is 2‐CH3, 3‐CH3, 4‐CH3, 2‐OCH3, 3‐OCH3, and 4‐OCH3) were prepared and copolymerized with styrene (ST). The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring‐substituted benzaldehydes and ethyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C NMR. All the ethylenes were copolymerized with ST (M1) in solution with radical initiation (AIBN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C NMR. The order of relative reactivity (1/r 1) for the monomers is 3‐OCH3 (0.88)?>?4‐CH3 (0.71)?>?2‐OCH3 (0.68)?>?3‐CH3 (0.55)?>?2‐CH3 (0.47)?>?4‐OCH3 (0.40). Higher T g of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the TSE structural unit. Gravimetric analysis indicated that the copolymers decompose in the 257–287°C range.  相似文献   

7.
Tetrahydroberberine (systematic name: 9,10‐dimethoxy‐5,8,13,13a‐tetrahydro‐6H‐benzo[g][1,3]benzodioxolo[5,6‐a]quinolizine), C20H21NO4, a widely distributed naturally occurring alkaloid, has been crystallized as a racemic mixture about an inversion center. A bent conformation of the molecule is observed, with an angle of 24.72 (5)° between the arene rings at the two ends of the reduced quinolizinium core. The intermolecular hydrogen bonds that play an apparent role in crystal packing are 1,3‐benzodioxole –CH2...OCH3 and –OCH3...OCH3 interactions between neighboring molecules.  相似文献   

8.
The salts 3‐[(2,2,3,3‐tetrafluoropropoxy)methyl]pyridinium saccharinate, C9H10F4NO+·C7H4NO3S, (1), and 3‐[(2,2,3,3,3‐pentafluoropropoxy)methyl]pyridinium saccharinate, C9H9F5NO+·C7H4NO3S, (2), i.e. saccharinate (or 1,1‐dioxo‐1λ6,2‐benzothiazol‐3‐olate) salts of pyridinium with –CH2OCH2CF2CF2H and –CH2OCH2CF2CF3meta substituents, respectively, were investigated crystallographically in order to compare their fluorine‐related weak interactions in the solid state. Both salts demonstrate a stable synthon formed by the pyridinium cation and the saccharinate anion, in which a seven‐membered ring reveals a double hydrogen‐bonding pattern. The twist between the pyridinium plane and the saccharinate plane in (2) is 21.26 (8)° and that in (1) is 8.03 (6)°. Both salts also show stacks of alternating cation–anion π‐interactions. The layer distances, calculated from the centroid of the saccharinate plane to the neighbouring pyridinium planes, above and below, are 3.406 (2) and 3.517 (2) Å in (1), and 3.409 (3) and 3.458 (3) Å in (2).  相似文献   

9.
Density‐functional theory calculations of a series of organic biradicals on the basis of the N,N′‐dioxy‐2,6‐diazaadamantane core with different substituents at carbon atoms adjacent to the nitroxyl groups have been performed by the UB3LYP/6‐311++G(2d,2p) method. Using the breaking symmetry approach, the values of the exchange interaction parameter, J, between the radical centers are calculated. It is shown that the intramolecular exchange interaction for the most part is ferromagnetic in nature, but the J parameter gradually decreases, changing its sign to antiferromagnetic interaction for the last substituent in the following sequence: CF3(CH3)COH > CH2F(H)COH > CH2OH > H > CBr3 > CH2F > CCl3 > CF3 > CH2Br > CH2Cl > CH3 > C2H5 > C3H7 > i‐C4H9 > F > Br > OCH3 > Cl > CH2C6H5. The calculations at the UHSEH1PBE/6‐311++G(2d,2p) level with the most of substituents show nearly the same variation sequence for the J parameter. It is concluded that spin polarization effects in the diazaadamantane cage and a direct through‐space antiferromagnetic exchange interaction between the nitroxyl groups are the main mechanisms contributing to the exchange interaction parameter value in the studied series of compounds. The exchange coupling constant, J, depends on the electronic effects and geometry of the substituents, as well as on their specific interactions with the nitroxyl radical groups.  相似文献   

10.
The dithiocarbene complex W(CO)5[C(SCH3)2 reacts with tertiary phosphines, PPh2CH3, PPh(CH3)2, P(C2H5)3 and P(OCH3)3 to form the phosphorane complexes W(CO)5[CH3S)2C-PR3] and with HPPh2 to form the phosphine complex W(CO)5[PPh2[CH(SCH3)2]. Kinetic studies of both types of reactions show that their rates are first order each in W(CO)5[C(SCH3)2] and in the phosphorus ligand. A mechanism involving rate determining phosphorus attack at the carbene carbon followed by rapid rearrangement to the product is consistent with this rate law. Rate constants for the reactions increase with increasing nucleophilicities of the phosphines: P(OCH3)3 < PPh2H < PPh2CH3 ? PPh(CH3)2 < P(C2H5)3. The ΔH values decrease (P(OCH3)3 > PPh2H > PPh2(CH3) > PPh(CH3)2 > P(C2H5)3) as the nucleophilicities of the phosphines increase. The ΔS values (≈-30 e.u.) remain essentially constant for all the reactions. The cyclic dithiorcarbenes W(CO)5[CS(CH2)nS], wheren- 3 or 4, react with PPh2(CH3) to form the cyclic phosphorane complexes, W(CO)5[S(CH2)nSC-PPh2(CH3)]. The 6- and 7- membered cyclic dithiocarbenes also react with PPh2H to form the phosphine complexes, W(CO)5 {PPh2- [CS(CH2)nS(H)]}.  相似文献   

11.
Reaction of Ti(OCH2CH2OR)4 (R?CH3 and C2H5) with 8‐hydroxyquinoline in benzene at room temperature resulted in the formation of Ti(C9H6NO)2(OCH2CH2OR)2, characterized by IR, 1H‐NMR, UV and mass spectroscopies. The molecular structure of Ti(C9H6NO)2(OCH2CH2OCH3)2 has been determined by single‐crystal X‐ray structure analysis. The geometry at titanium is a distorted octahedron, with the nitrogen atoms of quinolinate occupying the trans position with respect to oxygens of the 2‐methoxyethoxy groups. The prepared quinolinate derivatives of titanium alkoxides are very stable towards hydrolysis and harsh conditions are required for hydrolytic cleavage. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

12.
Fluorine Kα X-ray emission spectra have been measured and interpreted using UV photoelectron and X-ray photoelectron spectral data and the results of quantum-chemical calculations, for a series of fluorine-containing organic molecules: CH3F, n-C5F12, polytetrafluoroethylene, tetrafluoroethylene, 4-XC6H4F (X = H, F, NH2, NO2), 1,3-difluorobenzene, 1,2,4,5-tetrafluorobenzene, 1,4-difluorobenzene, C6F5X (X = H, F, SCH3, OCH3, CN, NO2, C6F5, P(OCH3)2), pentafluoropyridine, octafluoronaphthalene and 2,4-dinitrofluorobenzene, all in solid or gaseous states. It has been concluded that the fluorine 2pAO contribution to the highest occupied π-orbitals of the benzene ring and π-orbital of the ethylene bond is small: it is somewhat higher for a system of lower-lying π-orbitals and the highest for σ-orbitals. CH3F is assumed to have hyperconjugation.  相似文献   

13.
Cyclopentadienyl ligands bearing hydrophilic tentacles are obtained by treating Cp′SiMe2Cl (Cp′ Cp or C5Me4H) with H(OCH2CH2)nOMe (n = 2 or 3) in the presence of pyridine; the products Cp′SiMe2(OCH2CH2)nOMe can be deprotonated by potassium in toluene. Reaction of the potassium salt of CpSiMe2(OCH2CH2)2OMe with Me2SiCl2 yields C5H4(SiMe2Cl)(OCH2CH2)2OMe, which can be treated with H(OCH2CH2)2OMe to give C5H4[(OCH2CH2)2OMe]2.  相似文献   

14.
The structural effects of amineimide derivatives on photobase generation and the use of the resultant base for thermal curing of an epoxide/thiol system are investigated. The results of UV spectral change and gas chromatographic‐mass spectrometric analysis indicated that amineimide derivatives undergo photolysis by UV irradiation and generate bases. The order of conversion of the photolysis for the functional groups introduced to amineimide derivatives was NO2 > N(CH3)2 > CN > OCH3 > H. By using aminimide derivatives with NO2 and N(CH3)2 groups, the curing of the epoxide/thiol system was shifted to lower temperature after UV irradiation. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4045–4052, 2002  相似文献   

15.
A series of binuclear Cu(II) complexes with bisbenzoylhydrazones of 2,6-diformyl-4-R-phenols (R = OCH3, CH3, CH2Ph, Cl, Br, CO2CH3) with H2LCu2(NO3)2(OCH3) composition, where H2Lare monodeprotonated bishydrazones, was synthesized. The structure of the obtained complexes was established using IR spectro-scopy and conductometric and magnetochemical data. The title complexes were shown to exhibit exchange interaction of the antiferromagnetic type. Results of quantum-chemical calculations of structures modeling the state of a ligand in the complex were used to discuss the effect of substituent R nature on magnetic exchange in the complexes.  相似文献   

16.
Using silyl protected organic hydroxo compounds substitution of fluorine in IF5 is successful.Reacting IF5 with Si(OCH3)4 in CH3CN or SO2 using different molar ratios it was shown that in the series IF5?n(OCH3)n only the first member IF4(OCH3) (n=1) is stable enough to be isolated. The product in solution with n=2 bismutates to products with n=1 and n=3 if isolated as solids. The last one decomposes to the new oxo compound IF2O(OCH3) under elimination of CH3OCH3. With n=4,5 only redox reaction products could be isolated.IF2O(OCH3) can also be obtained by treating IF4(OCH3) with (CH3)6Si2O. Similarly reaction of IF5 with the disiloxane represents a new method to win IOF3. Excess of the oxygen transfer reagent leads to formation of IO2F and I2O5. An other oxo compound, IO(CH3COO)3, can be prepared by disolving IF5, IOF3 or IO2F in acetic acid anhydride.Reactions of IF5 with trimethylsilyl protected fluorinated benzoic acids RfCOOSi(CH3)3 (Rf = C6F5, 4HC6F4) appeared to be independent of the educts molar ratios because the only products are IF(RfCOO)4.In order to stabilize iodine (V) derivates with bifunctional chelating oxo ligands we applicated bis(trimethylsilyl) pinacolate, and in smooth reactions we yielded IF3[OC(CH3)2C(CH3)2O] and IF[OC(CH3)2  C(CH3)2O]2, in which iodine is part of five membered heterocyclic rings. The 19F-nmr-spectra are consistent with the diolate occupying the axiale and equatorial positions.An extension of the silyl method is the new synthesis of C6F5IF4 which could be obtained in the smooth reaction of IF5 with stochiometric amounts of Si(C6F5)4.  相似文献   

17.
Three new monophosphine-substituted iron carbonyl cluster complexes [(μ-PDT)Fe2(CO)5L] [(PDT = SCH2CH2CH2S, L = P(CH2Ph)3, 1; P(C6H11)3, 2; PPh2(PhMe-p), 3)], which can be regarded as active site mimics for [FeFe]-hydrogenase, have been prepared in 40–70 % yields by reactions of the parent complex (μ-PDT)Fe2(CO)6 (A) with monophosphine ligands in the presence of the decarbonylating agent Me3NO·2H2O. All three complexes were characterized by elemental analysis and spectroscopic techniques, as well as by X-ray crystallography for complex 1. The IR spectra of the complexes reveal that the electron-donating abilities of the different monophosphine ligands follow the order PPh2(PhMe-p) > P(C6H11)3 > P(CH2Ph)3.  相似文献   

18.
Small-angle neutron scattering (SANS) data have been obtained for (i) a series of solutions of C m H2m+1(OCH2–CH2)2SO4Na, for m=18, 16, and 14; (ii) an approximately 0.07M solution of C14H29(OCH2–CH2)2SO4Na to which different amounts of NaCl were added; and (iii) a series of solutions of variable concentration of C12H25(OCH2–CH2)SO4Na. The increase of the number of carbon atoms of the hydrocarbon chain produces a noticeable increase of the aggregation number of the micelles, while the salt tolerance decreases with increasing m. All the data can be described in terms of a monodispersed, charged, hard-spheres model interacting via a screened Coulombic potential, except the run at highest salt concentration, for which an ellipsoid model gives better results.  相似文献   

19.
The values of the fraction of ionizes phenyl salicylate, fPS-, obtained from initial absorbance measurement of phenyl salicylate at 350 nm, remain unchanged with the increase in [CH3CO2Na] from 0.0 to 0.7 M at 0.01 M NaOH (fPS- ≈ 0.70) and 0.02 M NaOH (fPS- ≈ 0.93). The values of fPS- decrease from ~ 1.0 to 0.90 and ~ 1.0 to 0.84 with the increase in respective [CH3CO2Na] and [NaBr] from 0.0 to 0.6 M at 0.01 M NaOH, 0.02 M C12E23(=C12H25(OCH2CH2)23OH) and 0.01 M CTABr (=C16H33NMe3Br).  相似文献   

20.
The accelerated formation of 2,3-diphenylquinoxalines in microdroplets generated in a nebulizer has been investigated by competition experiments in which equimolar quantities of 1,2-phenylenediamine, C6H4(NH2)2, and a 4-substituted homologue, XC6H3(NH2)2 [X = F, Cl, Br, CH3, CH3O, CO2CH3, CF3, CN or NO2], or a 4,5-disubstituted homologue, X2C6H2(NH2)2 [X = F, Cl, Br, or CH3], compete to condense with benzil, (C6H5CO)2. Electron-donating substituents (X = CH3 and CH3O) accelerate the reaction; in contrast, electron-attracting substituents (X = F, Cl, Br and particularly CO2CH3, CN, CF3 and NO2) retard it. A structure–reactivity relationship in the form of a Hammett correlation has been found by analyzing the ratio of 2,3-diphenylquinoxaline and the corresponding substituted-2,3-diphenylquinoxaline, giving a ρ value of −0.96, thus confirming that the electron density in the aromatic ring of the phenylenediamine component is reduced in the rate-limiting step in this accelerated condensation. This correlation shows that the phenylenediamine acts as a nucleophile in the reaction.  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