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.
Cavity ring‐down (CRD) techniques were used to study the kinetics of the reaction of Br atoms with ozone in 1–205 Torr of either N2 or O2, diluent at 298 K. By monitoring the rate of formation of BrO radicals, a value of k(Br + O3) = (1.2 ± 0.1) × 10−12 cm3 molecule−1 s−1 was established that was independent of the nature and pressure of diluent gas. The rate of relaxation of vibrationally excited BrO radicals by collisions with N2 and O2 was measured; k(BrO(v) + O2 → BrO(v − 1) + O2) = (5.7 ± 0.3) × 10−13 and k(BrO(v) + N2 → BrO(v − 1) + N2) = (1.5 ± 0.2) × 10−13 cm3 molecule−1 s−1. The increased efficiency of O2 compared with N2 as a relaxing agent for vibrationally excited BrO radicals is ascribed to the formation of a transient BrO–O2 complex. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 125–130, 2000  相似文献   

2.
A series of indole- and carbazole-substituted pyridinium iodide salts has been synthesized and characterized. X-ray analysis revealed that the iodide salt of the indole-substituted cation (E)-4-(1H-indol-3-yl­vinyl)-N-methyl­pyridinium (IMPE+), C16H15N2+·I, (I), has two polymorphic modifications, (Ia) and (Ib), and a hemihydrate structure, C16H15N2+·I·0.5H2O, (II). Until now, only one crystal modi­fication has been identified for the (E)-4-(9-ethyl-9H-carbazol-3-yl­vinyl)-N-methyl­pyridinium (ECMPE+) iodide salt, C22H21N2+·I, (III). Crystals of (Ia) and (Ib) comprise stacks of antiparallel cations with iodide anions located in the channels between the stacks. Due to the presence of the water mol­ecules, the packing in (II) is quite different to that found in (Ia) and (Ib), and positional disorder involving a statistical superposition of two rotamers of IMPE+, with different orientations of the indole fragment, was found. Crystals of (III) contain two independent ECMPE+ rotamers with different orientations of their carbazole substituents. The cations are packed in stacks, with the iodide anions located in the channels between the stacks. In (III), the iodide was found to be disordered over two sites, with occupancies of 0.83 and 0.17.  相似文献   

3.
Despite the high profile of amphetamine, there have been relatively few structural studies of its salt forms. The lack of any halide salt forms is surprising as the typical synthetic route for amphetamine initially produces the chloride salt. (S)‐Amphetamine hydrochloride [systematic name: (2S)‐1‐phenylpropan‐2‐aminium chloride], C9H14N+·Cl, has a Z′ = 6 structure with six independent cation–anion pairs. That these are indeed crystallographically independent is supported by different packing orientations of the cations and by the observation of a wide range of cation conformations generated by rotation about the phenyl–CH2 bond. The supramolecular contacts about the anions also differ, such that both a wide variation in the geometry of the three N—H...Cl hydrogen bonds formed by each chloride anion and differences in C—H...Cl contacts are apparent. (S)‐Amphetamine hydrobromide [systematic name: (2S)‐1‐phenylpropan‐2‐aminium bromide], C9H14N+·Br, is broadly similar to the hydrochloride in terms of cation conformation, the existence of three N—H...X hydrogen‐bond contacts per anion and the overall two‐dimensional hydrogen‐bonded sheet motif. However, only the chloride structure features organic bilayers and Z′ > 1.  相似文献   

4.
This paper overviews three living cationic polymerization systems (for styrene, p-methoxystyrene, and isobutyl vinyl ether) that are, in common, featured by: (i) specifically in nonpolar solvents, the use of the hydrogen halide/metal halide initiating systems (HX/MXn; X: I, Br, Cl; MXn: ZnX2, SnCl4), which generate a living growing carbocation stabilized by a nucleophilic counteranion (X…MXn); (ii) specifically in polar solvents, the use of externally added ammonium salts (nBu4N+Y; Y: I, Br, Cl), which permit the generation of living species from HX/MXn by providing nucleophilic halogen anions Y, either the same as or different from the halogen X in HX.  相似文献   

5.
A new type of single-ion conductor with fixed cation was synthesized by spontaneous anionic polymerization of 4-vinylpyridine in the presence of short polyethylene oxide ( PEO ) chains as alkylating agents. These comblike polymers have low Tgs and are amorphous with the shorter PEO s. Their conductivities are unaffected by the nature of the anion ( Br , ClO 4, and tosylate) and are controlled by the free volume and the mobility of the pendant cation. By comparison of the results at constant free volume, it is shown that the charge density decreases with the increasing length of pendant PEO demonstrating that PEO acts only as a plasticizing agent. Best conductivity results (σ = 10−5 S cm−1 at 60°C) are obtained with PEO side chains of molecular weight 350. With this sample, the conductivity in the presence of various amounts of added salt (LiTFSI) was studied. A best value of 10−4 S cm−1 at 60°C is obtained with a molar ratio EO/Li of 10. It is shown that, over the range of examined concentrations (0.2–1.3 mol Li kg−1), the reduced conductivity σr/c increases linearly with increasing salt concentration showing that the ion mobility increases continuously. Such behavior is quite unusual since in this concentration range a maximum is generally observed with PEO systems. To interpret this result and by analogy with the behavior of this type of polymer in solution, it is proposed that the conformation of these polymers in the solid state is segregated with the P4VP skeleton more or less confined inside the dense coils surrounded by the PEO side chains. Under the influence of the increasing salt concentration, this microphase separation vanishes progressively: The LiTFSI salt exchanges with the tosylate anions and acts as a miscibility improver agent. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35 : 2719–2728, 1997  相似文献   

6.
Six ammonium carboxylate salts, namely cyclopentylammonium cinnamate, C5H12N+·C9H7O2, (I), cyclohexylammonium cinnamate, C6H14N+·C9H7O2, (II), cycloheptylammonium cinnamate form I, C7H16N+·C9H7O2, (IIIa), and form II, (IIIb), cyclooctylammonium cinnamate, C8H18N+·C9H7O2, (IV), and cyclododecylammonium cinnamate, C12H26N+·C9H7O2, (V), are reported. Salts (II)–(V) all have a 1:1 ratio of cation to anion and feature three N+—H...O hydrogen bonds forming one‐dimensional hydrogen‐bonded columns consisting of repeating R43(10) rings, while salt (I) has a two‐dimensional network made up of alternating R44(12) and R68(20) rings. Salt (III) consists of two polymorphic forms, viz. form I having Z′ = 1 and form II with Z′ = 2. The latter polymorph has disorder of the cycloheptane rings in the two cations, as well as whole‐molecule disorder of one of the cinnamate anions. A similar, but ordered, Z′ = 2 structure is seen in salt (IV).  相似文献   

7.
In the salt 1‐methylpiperazine‐1,4‐diium bis(dihydrogen phosphate), C5H13N22+·2H2PO4, (I), and the solvated salt 2‐(pyridin‐2‐yl)pyridinium dihydrogen phosphate–orthophosphoric acid (1/1), C10H9N2+·H2PO4·H3PO4, (II), the formation of O—H...O and N—H...O hydrogen bonds between the dihydrogen phosphate (H2PO4) anions and the cations constructs a three‐ and two‐dimensional anionic–cationic network, respectively. In (I), the self‐assembly of H2PO4 anions forms a two‐dimensional pseudo‐honeycomb‐like supramolecular architecture along the (010) plane. 1‐Methylpiperazine‐1,4‐diium cations are trapped between the (010) anionic layers through three N—H...O hydrogen bonds. In solvated salt (II), the self‐assembly of H2PO4 anions forms a two‐dimensional supramolecular architecture with open channels projecting along the [001] direction. The 2‐(pyridin‐2‐yl)pyridinium cations are trapped between the open channels by N—H...O and C—H...O hydrogen bonds. From a study of previously reported structures, dihydrogen phosphate anions show a supramolecular flexibility depending on the nature of the cations. The dihydrogen phosphate anion may be suitable for the design of the host lattice for host–guest supramolecular systems.  相似文献   

8.
4,4′‐Bipyrazolium [or 4‐(1H‐pyrazol‐4‐yl)pyrazolium] bromide monohydrate, C6H7N4+·Br·H2O, and 4,4′‐bipyrazolium perchlorate monohydrate, C6H7N4+·ClO4·H2O, have closely related layered structures involving tight stacks of antiparallel N—H⋯N hydrogen‐bonded polar bipyrazolium chains [N⋯N = 2.712 (3) and 2.742 (2) Å], which are crosslinked by hydrogen bonds with water mol­ecules and counter‐anions.  相似文献   

9.
The objective of this paper is to discuss: (i) the general approaches to living cationic polymerizations; (ii) the nature of the growing species thus generated. For the first, it is concluded that three general methods are currently available which involve the nucleophilic stabilization of the growing carbocations by (a) a suitable counteranion, (b) an added Lewis base, or (c) an added neutral salt. According to this view, a variety of initiating systems are classified. For the second, findings are presented for the recently developed living cationic polymerization of vinyl ethers by the HCl/SnCl4 initiating system in the presence of an added salt (nBu4N+Cl). The nature of the growing species therein is discussed on the basis of the steric structure of the living polymers, relative to nonliving counterparts, and the in-situ 13C NMR spectroscopic analysis of model reactions where the interaction of the growing end model [CH3CH(OR)Cl] with SnCl4 and the added salt is analyzed.  相似文献   

10.
The transient absorption bands (λmax = 330, 525 nm, kf = 5 × 109 dm3 mol−1 s−1) obtained on pulse radiolysis of N2O‐saturated neutral aqueous solution of 4,4′‐thiodiphenol (TDPH) are due to the reaction of TDPH with ·OH radicals and are assigned to phenoxyl radical formed on fast deprotonation of the solute radical cation. The reaction of specific one‐electron oxidants (Cl2·−, Br2·−, N3·, TI2+, CCl3OO·) with TDPH also produced similar transient absorption bands. The phenoxyl radicals are also produced on pulse radiolysis of N2‐saturated solution of TDPH in 1,2‐dichloroethane. The nature of transient absorption spectrum obtained on reaction of ·OH radicals with TDPH is not affected in acidic solutions, showing that OH‐adduct is not formed in neutral solutions. The oxidation potential for the formation of phenoxyl radical is determined to be 0.98 V. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 603–610, 1999  相似文献   

11.
A dinuclear Dy (III) complex [Dy2( L 1 )2(NO3)4]·2CH3CN ( 1 ) (H L 1 = 1,3-bis{[(E)-pyridin-2-ylmethylene]amino}propan-2-ol) was obtained via the reaction of 1,3-diamino-2-propanol, 2-pyridyaldehyde and Dy (NO3)3·6H2O at room temperature. In the structure of complex 1 , two Dy (III) ions are in the N4O6 coordination environment provided by NO3 and ( L 1 ), and both of these ions are in the sphenocorona configuration. [Dy2( L 2 )2(NO3)4] ( 2 ) [H L 2 = 2-(pyridin-2-yl)hexahydropyrimidin-5-ol] was obtained using the same reaction material only when the reaction temperature was changed to 60°C. Structural analysis of complex 2 showed that the two Dy (III) ions with the same coordination configuration are in the N3O6 coordination environment provided by NO3 and ( L 2 ) and are in the distorted spherical-capped square antiprism. Surprisingly, H L 2 with the parent of bipyridine was synthesized by the Schiff base reaction of 1,3-diamino-2-propanol with 2-pyridoxaldehyde followed by the ring-closing reaction catalyzed by Dy (III) ions. Magnetic measurements of the Dy (III) complexes revealed no obvious frequency-dependent behavior of complex 1 . In contrast, complex 2 showed an obvious frequency dependence (Ueff = 0.49 K and τ0 = 6.62 × 10−4 s) under the condition of zero field and a weak double relaxation behavior (Ueff = 9.25 K and τ0 = 9.70 × 10−4 s) at 1500 Oe.  相似文献   

12.
Styrene underwent unprecedented coordination–insertion copolymerization with naked polar monomers (ortho ‐/meta ‐/para ‐methoxystyrene) in the presence of a pyridyl methylene fluorenyl yttrium catalyst. High activity (1.26×106 g molY−1 h−1) and excellent syndioselectivity were observed, and high‐molecular‐weight copolymers (24.6×104 g mol−1) were obtained. The insertion rate of the polar monomers could be adjusted in the full range of 0–100 % simply by changing the loading of the polar styrene monomer. Strikingly, the copolymers had tapered, gradient, and even random sequence distributions, depending on the position of the polar methoxy group on the phenyl ring and thus on its mode of coordination to the active metal center, as shown by tracking the polymerization process and DFT calculations.  相似文献   

13.
The heats of formation and strain energies for saturated and unsaturated three- and four-membered nitrogen and phosphorus rings have been calculated using G2 theory. G2 heats of formation (ΔHf298) of triaziridine [(NH)3], triazirine (N3H), tetrazetidine [(NH)4], and tetrazetine (N4H2) are 405.0, 453.7, 522.5, and 514.1 kJ mol−1, respectively. Tetrazetidine is unstable (121.5 kJ mol−1 at 298 K) with respect to its dissociation into two trans-diazene (N2H2) molecules. The dissociation of tetrazetine into molecular nitrogen and trans-diazene is highly exothermic (ΔH298 = −308.3 kJ mol−1 calculated using G2 theory). G2 heats of formation (ΔHf298) of cyclotriphosphane [(PH)3], cyclotriphosphene (P3H), cyclotetraphosphane [(PH)4], and cyclotetraphosphene (P4H2) are 80.7, 167.2, 102.7, and 170.7 kJ mol−1, respectively. Cyclotetraphosphane and cyclotetraphosphene are stabilized by 145.8 and 101.2 kJ mol−1 relative to their dissociations into two diphosphene molecules or into diphosphene (HP(DOUBLE BOND)PH) and diphosphorus (P2), respectively. The strain energies of triaziridine [(NH)3], triazirine (N3H), tetrazetidine [(NH)4], and tetrazetine (N4H2) were calculated to be 115.0, 198.3, 135.8, and 162.0 kJ mol−1, respectively (at 298 K). While the strain energies of the nitrogen three-membered rings in triaziridine and triazirine are smaller than the strain energies of cyclopropane (117.4 kJ mol−1) and cyclopropene (232.2 kJ mol−1), the strain energies of the nitrogen four-membered rings in tetrazetidine and tetrazetine are larger than those of cyclobutane (110.2 kJ mol−1) and cyclobutene (132.0 kJ mol−1). In contrast to higher strain in cyclopropane as compared with cyclobutane, triaziridine is less strained than tetrazetidine. The strain energies of cyclotriphosphane [(PH)3, 21.8 kJ mol−1], cyclotriphosphene (P3H, 34.6 kJ mol−1), cyclotetraphosphane [(PH)4, 24.1 kJ mol−1], and cyclotetraphosphene (P4H2, 18.5 kJ mol−1), calculated at the G2 level are considerably smaller than those of their carbon and nitrogen analog. Cyclotetraphosphene containing the P(DOUBLE BOND)P double bond is less strained than cyclotetraphosphane, in sharp contrast to the ratio between the strain energies for the analogous unsaturated and saturated carbon and nitrogen rings. © 1997 John Wiley & Sons, Inc. Int J Quant Chem 62 : 373–384, 1997  相似文献   

14.
π‐Conjugated organic materials exhibit high and tunable nonlinear optical (NLO) properties, and fast response times. 4′‐Phenyl‐2,2′:6′,2′′‐terpyridine (PTP) is an important N‐heterocyclic ligand involving π‐conjugated systems, however, studies concerning the third‐order NLO properties of terpyridine transition metal complexes are limited. The title binuclear terpyridine CoII complex, bis(μ‐4,4′‐oxydibenzoato)‐κ3O,O′:O′′;κ3O′′:O,O′‐bis[(4′‐phenyl‐2,2′:6′,2′′‐terpyridine‐κ3N,N′,N′′)cobalt(II)], [Co2(C14H8O5)2(C21H15N3)2], (1), has been synthesized under hydrothermal conditions. In the crystal structure, each CoII cation is surrounded by three N atoms of a PTP ligand and three O atoms, two from a bidentate and one from a symmetry‐related monodentate 4,4′‐oxydibenzoate (ODA2−) ligand, completing a distorted octahedral coordination geometry. Neighbouring [Co(PTP)]2+ units are bridged by ODA2− ligands to form a ring‐like structure. The third‐order nonlinear optical (NLO) properties of (1) and PTP were determined in thin films using the Z‐scan technique. The title compound shows a strong third‐order NLO saturable absorption (SA), while PTP exhibits a third‐order NLO reverse saturable absorption (RSA). The absorptive coefficient β of (1) is −37.3 × 10−7 m W−1, which is larger than that (8.96 × 10−7 m W−1) of PTP. The third‐order NLO susceptibility χ(3) values are calculated as 6.01 × 10−8 e.s.u. for (1) and 1.44 × 10−8 e.s.u. for PTP.  相似文献   

15.
Cavity ring‐down UV absorption spectroscopy was used to study the kinetics of the recombination reaction of FCO radicals and the reactions with O2 and NO in 4.0–15.5 Torr total pressure of N2 diluent at 295 K. k(FCO + FCO) is (1.8 ± 0.3) × 10−11 cm3 molecule−1 s−1. The pressure dependence of the reactions with O2 and NO in air at 295 K is described using a broadening factor of Fc = 0.6 and the following low (k0) and high (k) pressure limit rate constants: k0(FCO + O2) = (8.6 ± 0.4) × 10−31 cm6 molecule−1 s−1, k(FCO + O2) = (1.2 ± 0.2) × 10−12 cm3 molecule−1 s−1, k0(FCO + NO) = (2.4 ± 0.2) × 10−30 cm6 molecule−1 s−1, and k (FCO + NO) = (1.0 ± 0.2) × 10−12 cm3 molecule−1 s−1. The uncertainties are two standard deviations. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 130–135, 2001  相似文献   

16.
Using 2‐amino­methyl‐1H‐benz­imidazole as the ligand, a new thio­cyanate‐bridged copper(II) complex, namely bis(2‐aminomethyl‐1H‐benz­imidazole‐κ2N2,N3)­di­thio­cyanato­copper(II),[Cu(NCS)2(C8H9N3)], has been synthesized and structurally characterized. The Cu atom is five‐coordinated and exhibits a distorted square‐pyramidal geometry. The thio­cyanate ions (NCS) act as either bridging or terminal ligands. The bridging NCS ligands connect neighboring Cu atoms, constructing chains, while the terminal NCS ligands form hydrogen bonds with amine H atoms, leading to a complicated network.  相似文献   

17.
The sodium salt of [immucillin‐A–CO2H] (Imm‐A), namely catena‐poly[[[triaquadisodium(I)](μ‐aqua)[μ‐(1S)‐N‐carboxylato‐1‐(9‐deazaadenin‐9‐yl)‐1,4‐dideoxy‐1,4‐imino‐d ‐ribitol][triaquadisodium(I)][μ‐(1S)‐N‐carboxylato‐1‐(9‐deazaadenin‐9‐yl)‐1,4‐dideoxy‐1,4‐imino‐d ‐ribitol]] tetrahydrate], {[Na2(C12H13N4O6)2(H2O)7]·4H2O}n, (I), forms a polymeric chain via Na+—O interactions involving the carboxylate and keto O atoms of two independent Imm‐A molecules. Extensive N,O—H...O hydrogen bonding utilizing all water H atoms, including four waters of crystallization, provides crystal packing. The structural definition of this novel compound was made possible through the use of synchrotron radiation utilizing a minute fragment (volume ∼2.4 × 10−5 mm−3) on a beamline optimized for protein data collection. A summary of intra‐ring conformations for immucillin structures indicates considerable flexibility while retaining similar intra‐ring orientations.  相似文献   

18.
N‐Heterocyclic carbene (NHC) based systems are usually exploited in the exploration of catalytic mechanisms and processes in organocatalysis, and homo‐ and heterogeneous catalysis. However, their molecular structures have not received adequate attention. The NHC proligand methylenebis(N‐butylimidazolium) has been synthesized as the acetonitrile solvate of the diiodide salt, C15H26N42+·2I·CH3CN [1,1′‐methylenebis(3‐butylimidazolium) diiodide acetonitrile monosolvate], and fully characterized. An interesting cation–anion connection pattern has been identified in the crystal lattice, in which three iodide anions interact simultaneously with the cisoid‐oriented cation. A Hirshfeld surface analysis reveals the predominance of hydrogen bonding over anion–π interactions. This particular arrangement is observed in different methylene‐bridged bis(imidazolium) cations bearing chloride or bromide counter‐anions. Density functional theory (DFT) calculations with acetonitrile as solvent reproduce the geometry of the title cation.  相似文献   

19.
Carbamazepine (CBZ) is well known as a model active pharmaceutical ingredient used in the study of polymorphism and the generation and comparison of cocrystal forms. The pharmaceutical amide dihydrocarbamazepine (DCBZ) is a less well known material and is largely of interest here as a structural congener of CBZ. Reaction of DCBZ with strong acids results in protonation of the amide functionality at the O atom and gives the salt forms dihydrocarbamazepine hydrochloride {systematic name: [(10,11‐dihydro‐5H‐dibenzo[b,f]azepin‐5‐yl)(hydroxy)methylidene]azanium chloride, C15H15N2O+·Cl}, dihydrocarbamazepine hydrochloride monohydrate {systematic name: [(10,11‐dihydro‐5H‐dibenzo[b,f]azepin‐5‐yl)(hydroxy)methylidene]azanium chloride monohydrate, C15H15N2O+·Cl·H2O} and dihydrocarbamazepine hydrobromide monohydrate {systematic name: [(10,11‐dihydro‐5H‐dibenzo[b,f]azepin‐5‐yl)(hydroxy)methylidene]azanium bromide monohydrate, C15H15N2O+·Br·H2O}. The anhydrous hydrochloride has a structure with two crystallographically independent ion pairs (Z′ = 2), wherein both cations adopt syn conformations, whilst the two hydrated species are mutually isostructural and have cations with anti conformations. Compared to neutral dihydrocarbamazepine structures, protonation of the amide group is shown to cause changes to both the molecular (C=O bond lengthening and C—N bond shortening) and the supramolecular structures. The amide‐to‐amide and dimeric hydrogen‐bonding motifs seen for neutral polymorphs and cocrystalline species are replaced here by one‐dimensional polymeric constructs with no direct amide‐to‐amide bonds. The structures are also compared with, and shown to be closely related to, those of the salt forms of the structurally similar pharmaceutical carbamazepine.  相似文献   

20.
In 1‐naphthylammonium iodide, C10H10N+·I, and naphthalene‐1,8‐diyldiammonium diiodide, C10H12N22+·2I, the predominant hydrogen‐bonding pattern can be described using the graph‐set notation R42(8). This is the first report of a structure of a diprotonated naphthalene‐1,8‐diyldiammonium salt.  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