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.
1H, 13C, 14N and 15N NMR chemical shifts were used to prove the structures of the products of 2‐chloropyrazine oxidation. It was shown that oxidation by hydrogen peroxide in acetic acid or m‐chloroperbenzoic acid leads to the N4‐oxide, whereas potassium persulfate in sulfuric acid gives the N1‐oxide as the main product. Additionally, the results of NMR measurements of products from the nucleophilic substitution of the chlorine atom by azide anion, yielding the respective azides, and ethylation reactions of both 2‐chloropyrazine N‐oxides leading to the N‐ethyl salts confirm the structures of both isomeric N‐oxides. Protonation studies of the compounds obtained are also reported. The favoured protonation site is found to be the N atom that is not hindered by any substituents, and in some cases probably the oxygen atom of the N‐oxide function. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

2.
Quinoline bridged imidazolium precursors 5,8‐bis(NR‐imidazolylidenylmethylene)quinoline PF6 salts [H2L](PF6)2 [R = Me ( 1a ), R = naphthylmethyl ( 1b )] were prepared by quaternization of N‐methylimidazole and N‐naphthylmethylimidazole with 5,8‐bis(bromomethyl)quinoline, respectively. Reaction of the imidazolium ligands 1a and 1b with Hg(OAc)2 and Ag2O in acetonitrile gave the macrocyclic transition metal carbene complexes [Hg2L2](PF6)4 ( 2a and 2b ) and [Ag2L2](PF6)2 ( 3a and 3b ), respectively. All the N‐heterocyclic carbene complexes were characterized in detail by NMR, ESI‐MS, and elemental analysis. Structures of complexes 2a and 3a were determined by X‐ray diffraction studies. Structural studies revealed that the coordination arrangement of the central mercury atom in complex 2a displays a tricoordinate mode and the molecular conformation results in a“closed” form with the bridging quinoline functionality in the macrocycle, whereas the silver complex 3a does not show an coordiantion between the bridging quinoline and the AgI ion, which results in an “open” conformation of the macrocycle. The HgII and AgI NHC complexes showed similar UV absorption and luminescence in acetonitrile solutions.  相似文献   

3.
Through the use of a fully C/N‐functionalized imidazole‐based anion, it was possible to prepare nitrogen‐ and oxygen‐rich energetic salts. When N,N‐dinitramino imidazole was paired with nitrogen‐rich bases, versatile ionic derivatives were prepared and fully characterized by IR, and 1H, and 13C NMR spectroscopy and elemental analysis. Both experimental and theoretical evaluations show promising properties for these energetic compounds, such as high density, positive heats of formation, good oxygen balance, and acceptable stabilities. The energetic salts exhibit promising energetic performance comparable to the benchmark explosive RDX (1,3,5‐trinitrotriazacyclohexane).  相似文献   

4.
The reaction of 1‐methylimidazole and α,α‐dibromo‐p‐xylene was followed by a metathesis reaction with fluorinated anion sources, which yielded new fluorinated imidazolium salts [C6H4(CH2(C4H6N2)2]2+ 2[A] where A = BF4 ( 2 ), PF6 ( 3 ), CF3SO3 ( 4 ), and CF3COO ( 5 ). The compounds were characterized by 1H‐, 13C‐, 19F‐, 31P NMR, and IR spectroscopy. Single crystal X‐ray diffraction data of compounds 2 , 3 , and 4 were also reported, whereas compound 5 was found to be a liquid. The solid compounds crystallized in the monoclinic P21/c space group and have similar crystallographic parameters. The study revealed that the different fluorinated anions affected the spatial arrangement of atoms and the extent of cation–anion interactions, hence, influenced the stability and coordination properties of the imidazolium salts. A trend was observed which related the strength of cation–anion interaction to physical properties such as melting point.  相似文献   

5.
The structures of the proton‐transfer compounds of 4,5‐dichlorophthalic acid (DCPA) with the aliphatic Lewis bases triethylamine, diethylamine, n‐butylamine and piperidine, namely triethylaminium 2‐carboxy‐4,5‐dichlorobenzoate, C6H16N+·C8H3Cl2O4, (I), diethylaminium 2‐carboxy‐4,5‐dichlorobenzoate, C4H12N+·C8H3Cl2O4, (II), bis(butanaminium) 4,5‐dichlorobenzene‐1,2‐dicarboxylate monohydrate, 2C4H12N+·C8H2Cl2O42−·H2O, (III), and bis(piperidinium) 4,5‐dichlorobenzene‐1,2‐dicarboxylate monohydrate, 2C5H12N+·C8H2Cl2O42−·H2O, (IV), have been determined at 200 K. All compounds have hydrogen‐bonding associations, giving discrete cation–anion units in (I) and linear chains in (II), while (III) and (IV) both have two‐dimensional structures. In (I), a discrete cation–anion unit is formed through an asymmetric R12(4) N+—H...O2 hydrogen‐bonding association, whereas in (II), chains are formed through linear N—H...O associations involving both aminium H‐atom donors. In compounds (III) and (IV), the primary N—H...O‐linked cation–anion units are extended into a two‐dimensional sheet structure via amide–carboxyl N—H...O and amide–carbonyl N—H...O interactions. In the 1:1 salts (I) and (II), the hydrogen 4,5‐dichlorophthalate anions are essentially planar with short intramolecular carboxyl–carboxyl O—H...O hydrogen bonds [O...O = 2.4223 (14) and 2.388 (2) Å, respectively]. This work provides a further example of the uncommon zero‐dimensional hydrogen‐bonded DCPA–Lewis base salt and the one‐dimensional chain structure type, while even with the hydrate structures of the 1:2 salts with the primary and secondary amines, the low dimensionality generally associated with 1:1 DCPA salts is also found.  相似文献   

6.
Three energetic salts of cyclo‐N5? were synthesized via a metathesis reaction of barium pentazolate and sulfates which was driven by the precipitation of BaSO4. All the energetic cyclo‐N5? salts were characterized by single‐crystal X‐ray diffraction, infrared (IR), 1H and 13C multinuclear NMR spectroscopies, thermal analysis (TGA and DSC), and elemental analysis. The salts exhibit relatively good detonation performance with low sensitivities and good thermal stabilities. This new method opens the door to exploring more pentazolate anion‐containing high‐performance energetic materials.  相似文献   

7.
Protonation of the highly reactive 1:1 intermediates produced in the reaction between triphenylphosphine and acetylenic esters by tetrazole derivatives leads to the formation of vinyltriphenylphosphonium salts. The cation of these salts undergoes an addition reaction with the counter anion in CH2Cl2 at room temperature to yield the corresponding stabilized phosphorus ylides. Elimination of triphenylphosphine from the stabilized phosphorus ylides leads to the corresponding electron‐poor N‐vinyl tetrazoles in fairly high yields. Structures of N‐vinyl tetrazoles were determined by IR, 1H NMR, 13C NMR and single crystal X‐ray structure analyses. The reaction is fairly regioselective and stereoselective.  相似文献   

8.
According to previous reports, metal cations or water molecules are necessary for the stabilization of pentazolate anion (cyclo‐N5?) at ambient temperature and pressure. Seeking a new method to stabilize N5? is a big challenge. In this work, three anhydrous, metal‐free energetic salts based on cyclo‐N5? 3,9‐diamino‐6,7‐dihydro‐5 H‐bis([1,2,4]triazolo)[4,3‐e:3′,4′‐g][1,2,4,5] tetrazepine‐2,10‐diium, N‐carbamoylguanidinium, and oxalohydrazinium (oxahy+) pentazolate were synthesized and isolated. All salts were characterized by elemental analysis, IR spectroscopy, 1H, 13C, and (in some cases) 15N NMR spectroscopy, thermal analysis (TGA and DSC), and single‐crystal XRD analysis. Computational studies associated with heats of formation and detonation performance were performed by using Gaussian 09 and Explo5 programs, respectively. The sensitivity of the salts towards impact and friction was determined, and overall the real N5 explosives showed promising energetic properties.  相似文献   

9.
The 2D porous copper(Ⅰ) complex with 1,3-dicyanobenzene (DCB), [Cu(DCB)2](PF6)(Me2CO) 1, exhibits channels along axis c, in which one molecule acetone and one anion PF6 per formula unit are included respectively. The reversible incorporation of guest acetone and acetonitrile, as well as the anion exchange from PF6^- to BF4^- or CF3SO3^-, was investigated by thermogravimetric (TG) analysis, ^1H NMR spectra and/or infrared absorption spectroscopy. Additionally, the incorporation of benzene and toluene into complex 1 was also discussed. Complex 1 exhibited size selectivity for guest inclusion or anion exchange.  相似文献   

10.
The two new title complexes, [Cu(N3)(dpyam)2]PF6 (dpyam is di‐2‐pyridylamine, C10H11N3), (I), and [Cu(N3)(dpyam)2]Cl·4H2O, (II), respectively, have been characterized by single‐crystal X‐ray diffraction. Both complexes display a distorted square‐pyramidal geometry. Each Cu atom is coordinated in the basal plane by three dpyam N atoms and one azide N atom in equatorial positions, and by another N atom from the dpyam group in the apical position. In complex (I), the one‐dimensional supra­molecular architecture is assembled via hydrogen‐bonding inter­actions between the amine N atom and terminal azide N atoms and the F atoms of the PF6 anion. For complex (II), hydrogen‐bonding inter­actions between the amine N atom, the Cl anion and water O atoms result in a two‐dimensional lattice.  相似文献   

11.
Dibenzo[18]crown‐6 derivatives 1 with two lateral tetraalkyloxy o‐terphenyl units were prepared and converted to the corresponding complexes KX ?1 (X=halide, BF4, PF6, SCN) and NH4PF6 ?1 . Complexation was probed by MALDI‐TOF spectrometry and NMR spectroscopy. Downfield shifts of 1H NMR signals for complexes with soft anions Br, I, SCN, and PF6 indicated the presence of tight ion pairs, whereas complexes with hard anions F, Cl, or BF4 showed no or little shifts. In 13C NMR spectra, upfield shifts were detected for soft anions. The character of the anion also influenced the mesomorphic properties of complexes MX ?1 (M=K, NH4), which were investigated by differential scanning calorimetry (DSC), polarizing optical microscopy (POM), and XRD in comparison to neat 1 . Hard anions slightly stabilize or even destabilize the mesophase. Soft anions, however, improve the mesomorphic properties yielding mesophases with up to 70 °C phase widths in the case of KI ?1 , KPF6 ?1 , and NH4PF6 ?1 . For complexes KSCN ?1 with a soft and bridging anion, the balance between mesophase stabilization and high order is shifted in favor of the plastic crystal phase.  相似文献   

12.
In the asymmetric unit of the title compound, C10H15N4O2+·H2PO4, there are two protonated amino­guanidinium cations and two dihydrogenphosphate anions. The positive charge on the protonated amidine group is delocalized over the three C—N bonds in a manner similar to that found in guanidinium salts. The amino­guanidinium cations are found to be the E‐isomer structures. Intra­molecular inter­actions of the N—H⋯N type are observed, leading to the formation of five‐membered rings. Extensive networks of O—H⋯O, N—H⋯O and C—H⋯O hydrogen bonds stabilize the three‐dimensional network. In the crystal structure, π–π inter­actions between the benzene rings, with a distance of 3.778 (2) Å between the ring centroids, also affect the packing of the mol­ecules.  相似文献   

13.
The intramolecular hydrogen‐bonding interactions and properties of a series of nitroamino[1,3,5]triazine‐based guanidinium salts were studied by using the dispersion‐corrected density functional theory method (DFT‐D). Results show that there are evident LP(N or O; LP=lone pair)→σ*(N? H) orbital interactions related to O???H? N or N???H? N hydrogen bonds. Quantum theory of atoms in molecules (QTAIM) was applied to characterize the intramolecular hydrogen bonds. For the guanidinium salts studied, the intramolecular hydrogen bonds are associated with a seven‐ or eight‐membered pseudo‐ring. The guanylurea cation is more helpful for improving the thermal stabilities of the ionic salts than other guanidinium cations. The contributions of different substituents on the triazine ring to the thermal stability increase in the order of ? NO223 (? ONO2)2. Energy decomposition analysis shows that the salts are stable owing to electrostatic and orbital interactions between the ions, whereas the dispersion energy has very small contributions. Moreover, the salts exhibit relatively high densities in the range of 1.62–1.89 g cm?3. The detonation velocities and pressures lie in the range of 6.49–8.85 km s?1 and 17.79–35.59 GPa, respectively, which makes most of them promising explosives.  相似文献   

14.
《中国化学会会志》2017,64(7):843-850
The organic salts 1‐(2‐pyridylmethyl)‐3‐alkylbenzimidazolium halide (pm‐RbH +X) and 1‐(2‐pyridylmethyl)‐3‐alkylimidazolium halide (pm‐R′iH +X′) were prepared (where R = 4‐, 3‐, 2‐fluorobenzyl ( 4f , 3f , and 2f , respectively), 4‐, 3‐, 2‐chlorobenzyl ( 4c , 3c , and 2c , respectively); 4‐methoxybenzyl (4mo); 2,3,4,5,6‐pentafluorobenzyl (f5); benzyl (b); and methyl (m)); X = Cl and Br; R′ = benzyl (b) and methyl (m); and X′ = Cl and I. From these salts, heteroleptic Ir(III ) complexes containing one N ‐heterocyclic carbene (NHC ) ligand [Ir(κ2‐ppy)22‐(pm‐Rb))]PF6 (R = 4f, 1 (PF6 ); 3f, 2 (PF6 ); 2f, 3 (PF6 ); f5b, 4 (PF6 ); 4c, 5 (PF6 ); 3c, 6 (PF6 ); 2c, 7 (PF6 ); 4mo, 8 (PF6 ); b, 9 (PF6 ); m, 10 (PF6 )) and [Ir(κ2‐ppy)22‐(pm‐R′i))]PF6 (R = b, 11 (PF6 ); m, 12 (PF6 )), were synthesized, and the crystal structures of 1 (PF6 ), 2 (PF6 ), 3 (PF6 ), 5 (PF6 ), 6 (PF6 ), 7 (PF6 ), 9 (PF6 ), 10 (PF6 ), and 12 (PF6 ) were determined by X‐ray diffraction. The neutral NHC ligands 1‐(2‐pyridylmethyl)‐3‐alkylbenzimidazolin‐2‐ylidene (pm‐Rb) and 1‐(2‐pyridylmethyl)‐3‐alkylimidazolin‐2‐ylidene (pm‐R′i) of all cations were found to be involved in the intermolecular π−π stacking interactions with the surrounding cations in the solid state, thereby probably influencing the photophysical behavior in the solid state and in solution. The absorption and emission properties of all the complexes show only small variations.  相似文献   

15.
Guanidine is the functional group on the side chain of arginine, one of the fundamental building blocks of life. In recent years, a number of compounds based on the aminoguanidine (AG) moiety have been described as presenting high anticancer activities. The product of condensation between two molecules of AG and one molecule of formaldehyde was isolated in the protonated form as the dinitrate salt (systematic name: 2,8‐diamino‐1,3,4,6,7,9‐hexaazanona‐1,8‐diene‐1,9‐diium dinitrate), C3H14N82+·2NO3, (I). The cation lacks crystallographically imposed symmetry and comprises two terminal planar guanidinium groups, which share an N—C—N unit. Each cation in (I) builds 14 N—H…O hydrogen bonds and is separated from adjacent cations by seven nitrate anions. The AG self‐condensation reaction in the presence of copper(II) chloride and chloride anions led to the formation of the organic–inorganic hybrid 1,2‐bis(diaminomethylidene)hydrazine‐1,2‐diium tetrachloridocuprate(II), (C2H10N6)[CuCl4], (II). Its asymmetric unit is composed of half a diprotonated 1,2‐bis(diaminomethylidene)hydrazine‐1,2‐diium dication and half a tetrachloridocuprate(II) dianion, with the CuII atom situated on a twofold rotation axis. The planar guanidinium fragments in (II) have their planes twisted by approximately 77.64 (5)° with respect to each other. The tetrahedral [CuCl4]2− anion is severely distorted and its pronounced `planarity' must originate from its involvement in multiple N—H…Cl hydrogen bonds. It was reported that [CuCl4]2− anions, with a trans‐Cl—Cu—Cl angle (Θ) of ∼140°, are yellow–green at room temperature, with the colour shifting to a deeper green as Θ increases and toward orange as Θ decreases. Brown salt (II), with a Θ value of 142.059 (8)°, does not fit the trend, which emphasizes the need to take other structural factors into consideration. In the crystal of salt (II), layers of cations and anions alternate along the b axis, with the minimum Cu…Cu distance being 7.5408 (3) Å inside a layer. The structures of salts (I) and (II) were substantiated via spectroscopic data. The endothermic reaction involved in the thermal decomposition of (I) requires additional oxygen. The title salts may be useful for the screening of new substances with biological activity.  相似文献   

16.
The title compounds, namely 2,6‐bis[(1,3‐dimethylimidazolin‐2‐ylidene)amino]pyridinium perchlorate, C15H24N7+·ClO4, (I), and bis{2,6‐bis[(1,3‐dimethylimidazolin‐2‐ylidene)amino]pyridinium} μ‐oxido‐bis[trichloridoiron(III)], (C15H24N7)2[Fe2Cl6O], (II), are structurally unusual examples of the organization of molecular units via base pairing. The cations in salts (I) and (II) are derived from the bisguanidine N2,N6‐bis(1,3‐dimethylimidazolin‐2‐ylidene)pyridine‐2,6‐diamine, which associates in centrosymmetric pairs via two N—H...N hydrogen‐bond interactions. N—H...N bridges are formed between the protonated pyridine N atom and one of the nonprotonated guanidine N atoms, with N...H distances of 2.01 (1)–2.10 (1) Å. Compound (I) contains two crystallographically independent cations and anions per asymmetric unit. One of the perchlorate anions is disordered, while the [Fe2Cl6O]2− anion lies on an inversion centre.  相似文献   

17.
Pulsed gradient spin‐echo (PGSE) diffusion characteristics for a) the new [brucinium][X] salts 6 a – f [ a : X=BF4?; b : X=PF6?; c : X=MeSO3?, d : X=CF3SO3?; e : X=BArF?; f : X=PtCl3(C2H4)?], b) 4‐tert‐butyl‐N‐benzyl analogue, 7 and c) the aryl carbocations (p‐R‐C6H4)2CH 9 a (R=CH3O) and 9 b (R=(CH3)2N), (p‐CH3O‐C6H4)xCPh3?x+ 10 a – c (x=1–3, respectively) and (p‐R‐C6H4)3C+ 11 (R=(CH3)2N) and 12 (R=H) all in several different solvents, are reported. The solvent dependence suggests strong ion pairing in CDCl3, intermediate ion pairing in CD2Cl2 and little ion pairing in [D6]acetone. 1H, 19F HOESY NMR spectra (HOESY: heteronuclear Overhauser effect spectroscopy) for 6 and 7 reveal a specific approach of the anion with respect to the brucinium cation plus subtle changes, which are related to the anion itself. Further, for carbocations 9 – 12 , (all as BF4? salts) based on the NOE results, one finds marked changes in the relative positions of the BF4? anion. In these aryl cationic species the anion can be located either a) very close to the carbonium ion carbon b) in an intermediate position or c) proximate to the N or O atom of the p‐substituent and remote from the formally positive C atom. This represents the first example of such a positional dependence of an anion on the structure of the carbocation. DFT calculations support the experimental HOESY results. The solid‐state structures for 6 c and the novel Zeise's salt derivative, [brucinium][PtCl3(C2H4)], 6 f , are reported. Analysis of 195Pt NMR and other NMR measurements suggest that the η2‐C2H4 bonding to the platinum centre in 6 f is very similar to that found in K[PtCl3(C2H4)]. Field dependent T1 measurements on [brucinium][PtCl3(C2H4)] and K[PtCl3(C2H4)], are reported and suggested to be useful in recognizing aggregation effects.  相似文献   

18.
The interaction of the antimigraine pharmaceutical agent frovatriptan with acetic acid and succinic acid yields the salts (±)‐6‐carbamoyl‐N‐methyl‐2,3,4,9‐tetrahydro‐1H‐carbazol‐3‐aminium acetate, C14H18N3O+·C2H3O2, (I), (R)‐(+)‐6‐carbamoyl‐N‐methyl‐2,3,4,9‐tetrahydro‐1H‐carbazol‐3‐aminium 3‐carboxypropanoate monohydrate, C14H18N3O+·C4H5O4·H2O, (II), and bis[(R)‐(+)‐6‐carbamoyl‐N‐methyl‐2,3,4,9‐tetrahydro‐1H‐carbazol‐3‐aminium] succinate trihydrate, 2C14H18N3O+·C4H4O42−·3H2O, (III). The methylazaniumyl substitutent is oriented differently in all three structures. Additionally, the amide group in (I) is in a different orientation. All the salts form three‐dimensional hydrogen‐bonded structures. In (I), the cations form head‐to‐head hydrogen‐bonded amide–amide catemers through N—H...O interactions, while in (II) and (III) the cations form head‐to‐head amide–amide dimers. The cation catemers in (I) are extended into a three‐dimensional network through further interactions with acetate anion acceptors. The presence of succinate anions and water molecules in (II) and (III) primarily governs the three‐dimensional network through water‐bridged cation–anion associations via O—H...O and N—H...O hydrogen bonds. The structures reported here shed some light on the possible mode of noncovalent interactions in the aggregation and interaction patterns of drug molecule adducts.  相似文献   

19.
The photochemical behavior of quaternary ammonium salts (QA salts) with N,N‐dimethyldithiocarbamate as photobase generators and the photoinitiated thermal crosslinking of poly(glycidyl methacrylate) (PGMA) with the QA salts were investigated. The formation of basic compounds in the photolysis of 1‐phenacyl‐(1‐azonia‐4‐azabicyclo[2,2,2]octane) N,N‐dimethyldithiocarbamate was ascertained by the color change of phenol red as an acid–base indicator. 1H NMR spectra of photoproducts in CDCl3 under N2 showed that the photolysis of 1‐naphthoylmethyl‐(1‐azonia‐4‐azabicyclo[2,2,2]octane) N,N‐dimethyldithiocarbamate resulted in the quantitative formation of triethylenediamine and a dithiocarbamate derivative. The presence of oxygen in the photolysis decreased the photolysis rate. The amine was also detected in its photolysis in polystyrene films. The effects of ammonio groups and counteranions of QA salts on the photoinitiated thermal crosslinking of PGMA films were also investigated. Quaternary ammonium dithiocarbamates acted as excellent photobase generators and effective photoinitiated thermal crosslinkers for PGMA. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 1329–1341, 2001  相似文献   

20.
The title compounds, diaquadinitramidatolithium(I), [Li(N3O4)(H2O)2], (I), and pyridinium dinitramidate, C5H6N+·N3O4, (II), differ significantly in their cation–anion contacts. The Li+ atom of (I) is coordinated by two O atoms of the dinitramide anion in a chelate and by four additional water molecules, with the Li and central N atom of the anion on a twofold rotation axis. The pyridinium cation of (II) exhibits a contact with the dinitramide anion via an intermolecular N—H...N hydrogen bridge. These interactions are compared with those found in reported anhydrous lithium dinitramide and ammonium dinitramide salts.  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