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 812 毫秒
1.
Three azide based compounds were synthesized employing aliphatic amines as site blocking agents: [Ni(N3)(C6H16N2)2]ClO4 ( I ) [C6H16N2 = N,N′‐diethylethylenediamine (DEDA)], [Cu8(N3)16(C6H18N4)2] ( II ) [C6H18N4 = tris(2‐aminoethyl) amine (TREN)], and [Cu7(N3)14(C7H19N3)2] · 2H2O ( III ) [C7H19N3 = 3,3′‐diamino‐N‐methyldipropylamine (DMDA)]. The compounds I and II have one‐dimensional structure and III has a two‐dimensional structure. Compound I is a simple linear cationic Ni–azide chain and compound II has Cu6 azide units forming a column terminated by the copper‐metalloligand generated in‐situ during the course of the reaction. The charge compensation perchlorate anions occupy spaces in between the chains in I . Compound II packs in a herringbone arrangement, which is not commonly observed in low‐dimensional structures. Compound III has one‐dimensional copper‐azide chains connected through copper‐metalloligand forming the two‐dimensional structure. All the three compounds exhibit anti‐ferromagnetic behavior.  相似文献   

2.
The crystal structures of the monomeric palladium(II) azide complexes of the type L2Pd(N3)2 (L = PPh3 ( 1 ), AsPh3 ( 2 ), and 2‐chloropyridine ( 3 )), the dimeric [(AsPh4)2][Pd2(N3)4Cl2] ( 4 ), the homoleptic azido palladate [(PNP)2][Pd(N3)4] ( 5 ) and the homoleptic azido platinates [(AsPh4)2][Pt(N3)4] · 2 H2O ( 6 ) and [(AsPh4)2][Pt(N3)6] ( 7 ) were determined by X‐ray diffraction at single crystals. 1 and 2 are isotypic and crystallize in the triclinic space group P1. 1 , 2 and 3 show terminal azide ligands in trans position. In 4 the [Pd2(N3)4Cl2]2– anions show end‐on bridging azide groups as well as terminal chlorine atoms and azide ligands. The anions in 5 and 6 show azide ligands in equal positions with almost local C4h symmetry at the platinum and palladium atom respectively. The metal atoms show a planar surrounding. The [Pt(N3)6]2– anions in 7 are centrosymmetric (idealized S6 symmetry) with an octahedral surrounding of six nitrogen atoms at the platinum centers.  相似文献   

3.
Abstract

A homologous series of di(4-alkyloxybenzoates) of 4,4′-dimercaptobiphenyl: CH3(CH2) n-1O?C6H4?COS?C6H4?C6H4?SOC?C6H4?O(CH2) n-1CH3,n=1–7, has been synthesized and the thermotropic liquid-crystalline behaviour investigated. All compounds exhibit enantiotropic mesomorphism over a remarkable temperature range. While the mesophase thermal stability is moderately higher than that found for the corresponding oxygenated analogues, the smectic stability is definitely lower. In fact, all the compounds are nematic but smectic mesomorphism (SC) is observed for n = 7. Compounds with n = 6 or 7 exhibit enantiotropic highly ordered smectic (or disordered crystal) phases, probably SG in type.  相似文献   

4.
Neutral benzene-ammonia clusters, prepared in a supersonic expansion, were ionized using multiphoton ionization. The cluster ions were investigated with a time-of-flight mass spectrometer. The observed major cluster ions, under 355-nm laser irradiation, resulting from prompt intracluster ion-molecule reaction and fragmentation following ionization are (C6H6)m(NH3)nH+, m = 1–6, n = 1–4 and (C6H6)m+, m = 1–3. The results of isotopic labeling experiments clearly indicate that C6H6 does not participate in intracluster ion-molecule reactions to form (C6H6)m(NH3)nH+. A local maximum appears at n = 2 in the intensity distribution of (C6H6)m(NH3) nH+ for each value of m under all experimental conditions. This finding indicates that (C6H6)m(NH3)2H+ is more stable than any other (C6H5)m(NH3)mH+ (n = 1,3,4) for m = 1–6.  相似文献   

5.
Synthesis, Crystal Structures, and Vibrational Spectra of [Pt(N3)6]2– and [Pt(N3)Cl5]2–, 195Pt and 15N NMR Spectra of [Pt(N3)nCl6–n]2– and [Pt(15NN2)n(N215N)6–n]2–, n = 0–6 By ligand exchange of [PtCl6]2– with sodium azide mixed complexes of the series [Pt(N3)nCl6–n]2– and with 15N‐labelled sodium azide (Na15NN2) mixtures of the isotopomeres [Pt(15NN2)n(N215N)6–n]2–, n = 0–6 and the pair [Pt(15NN2)Cl5]2–/[Pt(N215N)Cl5]2– are formed. X‐ray structure determinations on single crystals of (Ph4P)2[Pt(N3)6] ( 1 ) (triclinic, space group P1, a = 10.175(1), b = 10.516(1), c = 12.380(2) Å, α = 87.822(9), β = 73.822(9), γ = 67.987(8)°, Z = 1) and (Ph4As)2[Pt(N3)Cl5] · HCON(CH3)2 ( 2 ) (triclinic, space group P1, a = 10.068(2), b = 11.001(2), c = 23.658(5) Å, α = 101.196(14), β = 93.977(15), γ = 101.484(13)°, Z = 2) have been performed. The bond lengths are Pt–N = 2.088 ( 1 ), 2.105 ( 2 ) and Pt–Cl = 2.318 Å ( 2 ). The approximate linear azido ligands with Nα–Nβ–Nγ‐angles = 173.5–174.6° are bonded with Pt–Nα–Nβ‐angles = 116.4–121.0°. In the vibrational spectra the PtCl stretching vibrations of (n‐Bu4N)2[Pt(N3)Cl5] are observed at 318–345, the PtN stretching modes of (n‐Bu4N)2[Pt(N3)6] at 401–428 and of (n‐Bu4N)2[Pt(N3)Cl5] at 408–413 cm–1. The mixtures (n‐Bu4N)2[Pt(15NN2)n(N215N)6–n], n = 0–6 and (n‐Bu4N)2[Pt(15NN2)Cl5]/(n‐Bu4N)2[Pt(N215N)Cl5] exhibit 15N‐isotopic shifts up to 20 cm–1. Based on the molecular parameters of the X‐ray determinations the vibrational spectra are assigned by normal coordinate analysis. The average valence force constants are fd(PtCl) = 1.93, fd(PtNα) = 2.38 and fd(NαNβ, NβNγ) = 12.39 mdyn/Å. In the 195Pt NMR spectrum of [Pt(N3)nCl6–n]2–, n = 0–6 downfield shifts with the increasing number of azido ligands are observed in the range 4766–5067 ppm. The 15N NMR spectrum of (n‐Bu4N)2[Pt(15NN2)n(N215N)6–n], n = 0–6 exhibits by 15N–195Pt coupling a pseudotriplett at –307.5 ppm. Due to the isotopomeres n = 0–5 for terminal 15N six well‐resolved signals with distances of 0.03 ppm are observed in the low field region at –201 to –199 ppm.  相似文献   

6.
The title compound  C6N7(NHNH2)3 ( 1 ) was obtained from melem C6N7(NH2)3 or melon [C6N7(NH2)NH]n and hydrazine by an autoclave synthesis. Upon treatment with a 10 % HCl solution it is transformed into the trihydrochloride  [C6N7(NHNH3)3]Cl3 ( 2 ). Compounds 1 and 2 were analysed with 13C NMR, 15N NMR, FTIR and Raman spectroscopy. Furthermore, the single‐crystal X‐ray structure of the pentahydrate of 2 is reported (P\bar{1} , a = 674.96(3), b = 1214.17(6), c = 1272.15(6) pm, α = 66.288(2)°, β = 75.153(2)°, γ = 80.420(2)°, V = 920.30(8)·106 pm3, Z = 2, T = 90(2) K). The thermal decomposition of 1 and 2 was investigated with TG/DTA. Reaction of 1 with NaNO2/HCl yields triazido‐s‐heptazine, C6N7(N3)3 ( 3 ). Tris(tri‐n‐butylphosphinimino)‐s‐heptazine ( 4 ) was synthesised from 3 and characterised by means of 13C, 31P, 1H NMR, FTIR and MALDI‐TOF spectroscopy. Similar to s‐heptazine derivative 3 , compounds 1 and 4 are precursors for graphitic carbon nitrides, which have attracted considerable attention recently, and to various potential applications, such as flame retardants and (photo) catalysis.  相似文献   

7.
Building upon previous studies on the synthesis of bis(sigma)borate and agostic complexes of ruthenium, the chemistry of nido‐[(Cp*Ru)2B3H9] ( 1 ) with other ligand systems was explored. In this regard, mild thermolysis of nido‐ 1 with 2‐mercaptobenzothiazole (2‐mbzt), 2‐mercaptobenzoxazole (2‐mbzo) and 2‐mercaptobenzimidazole (2‐mbzi) ligands were performed which led to the isolation of bis(sigma)borate complexes [Cp*RuBH3L] ( 2 a – c ) and β‐agostic complexes [Cp*RuBH2L2] ( 3 a – c ; 2 a , 3 a : L=C7H4NS2; 2 b , 3 b : L=C7H4NSO; 2 c , 3 c : L=C7H5N2S). Further, the chemistry of these novel complexes towards various diphosphine ligands was investigated. Room temperature treatment of 3 a with [PPh2(CH2)nPPh2] (n=1–3) yielded [Cp*Ru(PPh2(CH2)nPPh2)‐BH2(L2)] ( 4 a – c ; 4 a : n=1; 4 b : n=2; 4 c : n=3; L=C7H4NS2). Mild thermolysis of 2 a with [PPh2(CH2)nPPh2] (n=1–3) led to the isolation of [Cp*Ru(PPh2(CH2)nPPh2)(L)] (L=C7H4NS2 5 a – c ; 5 a : n=1; 5 b : n=2; 5 c : n=3). Treatment of 4 a with terminal alkynes causes a hydroboration reaction to generate vinylborane complexes [Cp*Ru(R?C?CH2)BH(L2)] ( 6 and 7 ; 6 : R=Ph; 7 : R=COOCH3; L=C7H4NS2). Complexes 6 and 7 can also be viewed as η‐alkene complexes of ruthenium that feature a dative bond to the ruthenium centre from the vinylinic double bond. In addition, DFT computations were performed to shed light on the bonding and electronic structures of the new compounds.  相似文献   

8.
Six mono/double‐layered 2D and three 3D coordination polymers were synthesized by a self‐assembly reaction of Zn (II) salts, organic dicarboxylic acids and L1/L2 ligands. These polymeric formulas are named as [Zn(L1)(C4H2O4)0.5 (H2O)]n·0.5n(C4H2O4)·2nH2O ( 1 ), [Zn2(L2)(C4H2O4)2]n·2nH2O ( 2 ), [Zn(L1)(m‐BDC)]n ( 3 ), [Zn2(L2)(m‐BDC)2]n·2nH2O ( 4 ), [Zn3(L1)2(p‐BDC)3(H2O)4]n·2nH2O ( 5 ), [Zn2(OH)(L2) (p‐BDC)1.5]n ( 6 ), [Zn2(L1)(p‐BDC)2]n·5nH2O ( 7 ), [Zn2(L2)(p‐BDC)2]n·3nH2O ( 8 ) and [Zn2(L1)(C4H4O4)1.5(H2O)]n·n(ClO4nH2O ( 9 ) [L1 = N,N′‐bis (pyridin‐4‐ylmethyl)propane‐1,2‐diamine, L2 = N,N′‐bis (pyridin‐3‐ylmethyl)propane‐1,2‐ diamine, m‐BDC2? = m‐benzene dicarboxylate, p‐BDC2? = p‐benzene dicarboxylate]. Meanwhile, these polymers have been characterized by elemental analysis, infrared, thermogravimetry (TG), photoluminescence, powder and single‐crystal X‐ray diffraction. Polymers 1–6 present mono‐ and double (4,4)‐layer motifs accomplished by L1/L2 ligands with diverse conformations and organic dicarboxylates, and the layer thickness locates in the range of 5.8–15.0 Å. In three 3D polymers, the L1 and L2 molecules adopt the same cis‐conformations and join adjacent Zn (II) cations together with p‐BDC2? or succinate, giving rise to different binodal (4,4)‐c nets with (4.52.83)(4.53.72) ( 7 ), pts ( 8 ) topology and twofold interpenetrated binodal (5,5)‐c nets with (32.44.52.62)(3.43.52.64) ( 9 ). Therefore, the diverse conformations of the two bis (pyridyl)‐propane‐1,2‐diamines and the feature of different organic dicarboxylate can effectively influence the architectures of these polymers. Powder X‐ray diffraction patterns demonstrate that these bulk solid polymers are pure phase. TG analyses indicate that these polymers have certain thermal stability. Luminescent investigation reveals that the emission maximum of these polymers varies from 402 to 449 nm in the solid state at room temperature. Moreover, 1 , 3 and 5–8 show average luminescence lifetimes from 8.81 to 16.30 ns.  相似文献   

9.
Imine complexes [IrCl(η5‐C5Me5){κ1‐NH=C(H)Ar}{P(OR)3}]BPh4 ( 1 , 2 ) (Ar = C6H5, 4‐CH3C6H4; R = Me, Et) were prepared by allowing chloro complexes [IrCl25‐C5Me5){P(OR)3}] to react with benzyl azides ArCH2N3. Bis(imine) complexes [Ir(η5‐C5Me5){κ1‐NH=C(H)Ar}2{P(OR)3}](BPh4)2 ( 3 , 4 ) were also prepared by reacting [IrCl25‐C5Me5){P(OR)3}] first with AgOTf and then with benzyl azide. Depending on the experimental conditions, treatment of the dinuclear complex [IrCl25‐C5Me5)]2 with benzyl azide yielded mono‐ [IrCl25‐C5Me5){κ1‐NH=C(H)Ar}] ( 5 ) and bis‐[IrCl(η5‐C5Me5){κ1‐NH=C(H)Ar}2]BPh4 ( 6 ) imine derivatives. In contrast, treatment of chloro complexes [IrCl25‐C5Me5){P(OR)3}] with phenyl azide C6H5N3 gave amine derivatives [IrCl(η5‐C5Me5)(C6H5NH2){P(OR)3}]BPh4 ( 7 , 8 ). The complexes were characterized spectroscopically (IR, NMR) and by X‐ray crystal structure determination of [IrCl(η5‐C5Me5){κ1‐NH=C(H)C6H4‐4‐CH3}{P(OEt)3}]BPh4 ( 2b ).  相似文献   

10.
Rubidium dibarium penta­azide, RbBa2(N3)5, was prepared from an aqueous solution of the binary azides at room temperature. It crystallizes in the monoclinic system (space group P2/n). Two central atoms of azide groups occupy the 2c () and 2b () positions, another azide group lies completely on a twofold axis (2f), while Rb atoms are situated in 2e (2) positions. The crystal structure of RbBa2(N3)5 can be regarded as a distorted AlB2‐type arrangement of the metal atoms, with the azide groups occupying the voids between the cations. This results in coordination numbers of 8 (Rb) and 10 (Ba). The N—N distances are in the range 1.169 (8)–1.190 (5) Å, typical for the azide group.  相似文献   

11.
The first charge‐neutral Lewis base adducts of tin(IV) tetraazide, [Sn(N3)4(bpy)], [Sn(N3)4(phen)] and [Sn(N3)4(py)2], and the salt bis{bis(triphenylphosphine)iminium} hexa(azido)stannate [(PPN)2Sn(N3)6] (bpy = 2,2′‐bipyridine; phen = 1,10‐phenanthroline; py = pyridine; PPN = N(PPh3)2) have been prepared using covalent or ionic azide‐transfer reagents and ligand‐exchange reactions. The azides were isolated on the 0.3 to 1 g scale and characterized by IR and NMR spectroscopies, microanalytical and thermal methods and their molecular structures determined by single‐crystal XRD. All complexes have a distorted octahedral Sn[N]6 coordination geometry and possess greater thermal stability than their Si and Ge homologues. The nitrogen content of the adducts of up to 44 % exceed any SnIV compound known hitherto.  相似文献   

12.
In poly[di‐μ‐chlorido‐μ‐(4,4′‐bipyridazine)‐κ2N1:N1′‐cadmium(II)], [CdCl2(C8H6N4)]n, (I), and its isomorphous bromide analogue, [CdBr2(C8H6N4)]n, (II), the halide atom lies on a mirror plane and the CdII ion resides at the intersection of two perpendicular mirror planes with m2m site symmetry. The pyridazine rings of the ligand lie in a mirror plane and are related to each other by a second mirror plane perpendicular to the first. The compounds adopt the characteristic structure of the [MIIX2(bipy)] type (bipy is bipyridine) based on crosslinking of [Cd(μ‐X)2]n chains [Cd—Cl = 2.5955 (9) and 2.6688 (9) Å; Cd—Br = 2.7089 (4) and 2.8041 (3) Å] by bitopic rod‐like organic ligands [Cd—N = 2.368 (3)–2.380 (3) Å]. This feature is discussed in terms of supramolecular stabilization, implying that the periodicity of the inorganic chain [Cd...Cd = 3.7802 (4) Å in (I) and 3.9432 (3) Å in (II)] is favourable for extensive parallel π–π stacking of monodentate pyridazine rings, with centroid–centroid distances of 3.7751 (4) Å in (I) and 3.9359 (4) Å in (II). This is not the case for the longer iodide bridges, which cannot stabilize such a pattern. In poly[tetra‐μ‐iodido‐μ4‐(4,4′‐bipyridazine)‐κ4N1:N2:N1′:N2′‐dicadmium(II)], [Cd2I4(C8H6N4)]n, (III), the ligands are situated across a centre of inversion; they are tetradentate [Cd—N = 2.488 (2) and 2.516 (2) Å] and link successive [Cd(μ‐I)2]n chains [Cd—I = 2.8816 (3)–3.0069 (4) Å] into corrugated layers.  相似文献   

13.
利用密度泛函理论在B3LYP/6-311G*水平上对碱土金属叠氮化合物(MgN6)n(n=1~5)团簇各种可能构型进行了几何优化,预测了各团簇的最稳定结构。并对最稳定结构的成键特性、电荷分布、振动特性及稳定性进行理论研究。结果表明,叠氮化合物中叠氮基以直线型存在,MgN6团簇最稳定结构为直线型;(MgN6)2团簇最稳定结构为Mg2N2四元环平面结构;(MgN6)n(n=3~5)团簇最稳定结构是由2个叠氮基与2个Mg原子首先构成近似菱形,再由近似菱形延伸形成的链状结构。叠氮基中间的N原子显示正电性,两端的N原子显示负电性,且与Mg直接作用的N原子负电性更强,金属Mg原子和N原子之间形成很强的离子键。(MgN6)n(n=1~5)团簇最稳定结构的IR光谱分为4个部分,其最强振动峰均位于2209~2313cm-1,振动模式为叠氮基中N-N键的反对称伸缩振动。稳定性分析显示,(MgN6)3和(MgN6)5团簇相对于其他团簇较为稳定。  相似文献   

14.
Reaction of the ferriochlorosilanes R5C5(CO)2FeSiR′3-nCln (1a–1f) with sodium azide in tetrahydrofuran yields the ferrio- (mono-, bis-, and tris-azido)silanes R5C5(CO)2FeSiR′3-n(N3)n (R = H, Me; R′ = Me, H; n = 1–3) (2a–2f). CCl4 converts Cp(CO)2FeSiMe(H)N3 (2a) into the ferrioazido(chloro)silane Cp(CO)2-FeSiMe(Cl)N3 (3). Treatment of 2d, 2f with Me3P results in the formation of the ferriosilyl-iminophosphoranes Cp(CO)2FeSi(N3)(R)NPMe3 (R = Me, N3), (4a, 4b) by N2 elimination.  相似文献   

15.
Cyclic ketene N,X‐acetals 1 are electron‐rich dipolarophiles that undergo 1,3‐dipolar cycloaddition reactions with organic azides 2 ranging from alkyl to strongly electron‐deficient azides, e.g., picryl azide ( 2L ; R1=2,4,6‐(NO2)3C6H2) and sulfonyl azides 2M – O (R1=XSO2; cf. Scheme 1). Reactions of the latter with the most‐nucleophilic ketene N,N‐acetals 1A provided the first examples for two‐step HOMO(dipolarophile)–LUMO(1,3‐dipole)‐controlled 1,3‐dipolar cycloadditions via intermediate zwitterions 3 . To set the stage for an exploration of the frontier between concerted and two‐step 1,3‐dipolar cycloadditions of this type, we first describe the scope and limitations of concerted cycloadditions of 2 to 1 and delineate a number of zwitterions 3 . Alkyl azides 2A – C add exclusively to ketene N,N‐acetals that are derived from 1H‐tetrazole (see 1A ) and 1H‐imidazole (see 1B , C ), while almost all aryl azides yield cycloadducts 4 with the ketene N,X‐acetals (X=NR, O, S) employed, except for the case of extreme steric hindrance of the 1,3‐dipole (see 2E ; R1=2,4,6‐(tBu)3C6H2). The most electron‐deficient paradigm, 2L , affords zwitterions 16D , E in the reactions with 1A , while ketene N,O‐ and N,S‐acetals furnish products of unstable intermediate cycloadducts. By tuning the electronic and steric demands of aryl azides to those of ketene N,N‐acetals 1A , we discovered new borderlines between concerted and two‐step 1,3‐dipolar cycloadditions that involve similar pairs of dipoles and dipolarophiles: 4‐Nitrophenyl azide ( 2G ) and the 2,2‐dimethylpropylidene dipolarophile 1A (R, R=H, tBu) gave a cycloadduct 13 H , while 2‐nitrophenyl azide ( 2 H ) and the same dipolarophile afforded a zwitterion 16A . Isopropylidene dipolarophile 1A (R=Me) reacted with both 2G and 2 H to afford cycloadducts 13G , J ) but furnished a zwitterion 16B with 2,4‐dinitrophenyl azide ( 2I) . Likewise, 1A (R=Me) reacted with the isomeric encumbered nitrophenyl azides 2J and 2K to yield a cycloadduct 13L and a zwitterion 16C , respectively. These examples suggest that, in principle, a host of such borderlines exist which can be crossed by means of small structural variations of the reactants. Eventually, we use 15N‐NMR spectroscopy for the first time to characterize spirocyclic cycloadducts 10 – 14 and 17 (Table 6), and zwitterions 16 (Table 7).  相似文献   

16.
Reactions of Cp*NbCl4 and Cp*TaCl4 with Trimethylsilyl‐azide, Me3Si‐N3. Molecular Structures of the Bis(azido)‐Oxo‐Bridged Complexes [Cp*NbCl(N3)(μ‐N3)]2(μ‐O) and [Cp*TaCl2(μ‐N3)]2(μ‐O) (Cp* = Pentamethylcyclopentadienyl) The chloro ligands in Cp*TaCl4 (1c) can be stepwise substituted for azido ligands by reactions with trimethylsilyl azide, Me3Si‐N3 (A) , to generate the complete series of the bis(azido)‐bridged dimers [Cp*TaCl3‐n(N3)n(μ‐N3)]2 ( n = 0 (2c) , n = 1 (3c) , n = 2 (4c) and n = 3 (5c) ). If the solvent CH2Cl2 contains traces of water, an additional oxo bridge is incorporated to give [Cp*‐TaCl2(μ‐N3)]2(μ‐O) (6c) or [Cp*TaCl(N3)(μ‐N3)]2(μ‐O) (7c) , respectively. Both 6c and 7c are also formed in stoichiometric reactions from [Cp*TaCl2(μ‐OH)]2(μ‐O) (8c) and A . Analogous reactions of Cp*NbCl4 (1b) with A were used to prepare the azide‐rich dinuclear products [Cp*NbCl3‐n(N3)n(μ‐N3)]2 (n = 2 (4b) , and n = 3 (5b) ), and [Cp*NbCl(N3)(μ‐N3)]2(μ‐O) (7b) . The mononuclear complex Cp*Ta(N3)Me3 (10c) is obtained from Cp*Ta(Cl)Me3 and A . All azido complexes were characterised by their IR as well as their 1H and 13C NMR spectra; X‐ray crystal structure analyses are available for 6c and 7b .  相似文献   

17.
We synthesized melemium hydrogensulfate H3C6N7(NH2)3(HSO4)3 by reaction of melem with 70 % sulfuric acid. The crystal structure was elucidated by single‐crystal XRD (P21/n (no. 14), Z = 4, a = 10.277(2), b = 14.921(3), c = 11.771(2) Å, β = 99.24(3)°, V = 1781.5(6) Å3). H3C6N7(NH2)3(HSO4)3 is the first compound displaying a triple protonation of melem., In this contribution an overview of accessible melemium sulfates depending on the concentration of sulfuric acid is given. Two additional melemium sulfates were identified this way.  相似文献   

18.
Positive singly charged ionic liquid aggregates [(Cnmim)m+1(BF4)m]+ (mim = 3‐methylimidazolium; n = 2, 4, 8 and 10) and [(C4mim)m+1(A)m]+ (A = Cl, BF4, PF6, CF3SO3 and (CF3SO2)2N) were investigated by electrospray ionisation mass spectrometry and energy‐variable collision induced dissociation. The electrospray ionisation mass spectra (ESI‐MS) showed the formation of an aggregate with extra stability for m = 4 for all the ionic liquids with the exception of [C4mim][CF3SO3]. ESI‐MS‐MS and breakdown curves of aggregate ions showed that their dissociation occurred by loss of neutral species ([Cnmim][A])a with a ≥ 1. Variable‐energy collision induced dissociation of each aggregate from m = 1 to m = 8 for all the ionic liquids studied enabled the determination of Ecm, 1/2 values, whose variation with m showed that the monomers were always kinetically much more stable than the larger aggregates, independently of the nature of cation and anion. The centre‐of‐mass energy values correlate well with literature data on ionic volumes and interaction and hydrogen bond energies. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

19.
High-pressure Synthesis of Cesium Amide Azide, Cs2(NH2)N3 from Cesium Metal and Ammonia The reaction of cesium and yttrium metal with ammonia at 5–6 kbar and 190–220°C led to a well crystallized cesium amide azide and to YN. The formation of the cesium compound is discussed by volume effects. X-ray investigations gave the atomic arrangement of the compound. The tetragonal unit cell with a = 8.194(3) and c = 4.450(1) Å contains two formula units. The structure determination was successfull in the space group P4/mbm. The azide ion has different coordination and bond length (1.255 Å) as compared with that in the alkali metal azides (1.17 Å). The amide ions carry out a strong libration.  相似文献   

20.
Connecting two discotic mesogens via a spacer not only stabilizes the columnar mesophase but also leads to the formation of glass columnar phase, and therefore improves the physical properties of discotic liquid crystals as organic semiconductor. Here, we report the synthesis of eight diacetylene-bridged triphenylene discotic liquid crystal dimers, [C18H6(OCnH2n+1)4(OMe)O2C-C8H16-C≡≡ C-]2, 3(n), (n = 4-8), [C18H6(OC6H13)5O2C-C8H16-C≡≡ C-]2, 6 and [C18H6(OC6H13)5O-(CH2)m-C≡≡ C-]2, 8(m), (m = 1, 3) by Eglinto...  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