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 title compound [systematic name: 7‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐7H‐imidazo[1,2‐c]pyrrolo[2,3‐d]pyrimidine hemihydrate], 2C13H14N4O3·H2O or (I)·0.5H2O, shows two similar conformations in the asymmetric unit. These two conformers are connected through one water molecule by hydrogen bonds. The N‐glycosylic bonds of both conformers show an almost identical anti conformation with χ = −107.7 (2)° for conformer (I‐1) and −107.0 (2)° for conformer (I‐2). The sugar moiety adopts an unusual N‐type (C3′‐endo) sugar pucker for 2′‐deoxyribonucleosides, with P = 36.8 (2)° and τm = 40.6 (1)° for conformer (I‐1), and P = 34.5 (2)° and τm = 41.4 (1)° for conformer (I‐2). Both conformers and the solvent molecule participate in the formation of a three‐dimensional pattern with a `chain'‐like arrangement of the conformers. The structure is stabilized by intermolecular O—H...O and O—H...N hydrogen bonds, together with weak C—H...O contacts.  相似文献   

2.
The title compound {systematic name: 4‐amino‐1‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐5‐[6‐(1‐benzyl‐1H‐1,2,3‐triazol‐4‐yl)hex‐1‐ynyl]pyrimidin‐2(1H)‐one}, C24H28N6O4, shows two conformations in the crystalline state, viz. (I‐1) and (I‐2). The pyrimidine groups and side chains of the two conformers are almost superimposable, while the greatest differences between them are observed for the sugar groups. The N‐glycosylic bonds of both conformers adopt similar anti conformations, with χ = −168.02 (12)° for conformer (I‐1) and χ = −159.08 (12)° for conformer (I‐2). The sugar residue of (I‐1) shows an N‐type (C3′‐endo) conformation, with P = 33.1 (2)° and τm = 29.5 (1)°, while the conformation of the 2′‐deoxyribofuranosyl group of (I‐2) is S‐type (C3′‐exo), with P = 204.5 (2)° and τm = 33.8 (1)°. Both conformers participate in hydrogen‐bond formation and exhibit identical patterns resulting in three‐dimensional networks. Intermolecular hydrogen bonds are formed with neighbouring molecules of different and identical conformations (N—H...N, N—H... O, O—H...N and O—H...O).  相似文献   

3.
The title compound [systematic name: 4‐amino‐1‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐5‐ethynylpyrimidin‐2(1H)‐one], C11H13N3O4, shows two conformations in the crystalline state. The N‐glycosylic bonds of both conformers adopt similar conformations, with χ = −149.2 (1)° for conformer (I‐1) and −151.4 (1)° for conformer (I‐2), both in the anti range. The sugar residue of (I‐1) shows a C2′‐endo envelope conformation (2E, S‐type), with P = 164.7 (1)° and τm = 36.9 (1)°, while (I‐2) shows a major C3′‐exo sugar pucker (C3′‐exo‐C2′‐endo, 3T2, S‐type), with P = 189.2 (1)° and τm = 33.3 (1)°. Both conformers participate in the formation of a layered three‐dimensional crystal structure with a chain‐like arrangement of the conformers. The ethynyl groups do not participate in hydrogen bonding, but are arranged in proximal positions.  相似文献   

4.
Multiple bonds between boron and transition metals are known in many borylene (:BR) complexes via metal dπ→BR back-donation, despite the electron deficiency of boron. An electron-precise metal–boron triple bond was first observed in BiB2O [Bi≡B−B≡O] in which both boron atoms can be viewed as sp-hybridized and the [B−BO] fragment is isoelectronic to a carbyne (CR). To search for the first electron-precise transition-metal-boron triple-bond species, we have produced IrB2O and ReB2O and investigated them by photoelectron spectroscopy and quantum-chemical calculations. The results allow to elucidate the structures and bonding in the two clusters. We find IrB2O has a closed-shell bent structure (Cs, 1A′) with BO coordinated to an Ir≡B unit, (OB)Ir≡B, whereas ReB2O is linear (C∞v, 3Σ) with an electron-precise Re≡B triple bond, [Re≡B−B≡O]. The results suggest the intriguing possibility of synthesizing compounds with electron-precise M≡B triple bonds analogous to classical carbyne systems.  相似文献   

5.
Benzene oxide and the potential 8π-electron system oxepin exist in valence-tautomeric equilibrium with each other, to which both components contribute to approximately the same extent. NMR spectroscopic measurements show that the equilibrium is rapidly established (activation energies of the forward and reverse reactions 9.1 and 7.2 kcal mole?1, respectively). The present knowledge of the properties of oxepin justifies its classification as a “heterotropilidene”. Benzene oxide-oxepin represents a system having fluctuating bonds, the equilibrium of which can be displaced from one extreme to the other by means of suitable substituents. The oxide component determines the reactions of the system with most agents. With 1,6-oxido[10]annulene, which is formally a 2,7-bridged oxepin, the oxepin character is completely suppressed by the formation of a delocalized 10π-electron system extending over the C10 perimeter. The existence and aromatic character of 1,6-oxido[10]-annulene give rise to the conception of a homologous series of oxygen bridged annulenes (1,6; 8,13-bisoxido[14]annulene, 1,6; 8,17; 10,15-trisoxido[18]annulene etc.), which, like the parent acenes, possess a (4n + 2)π-electron system. Molecular models demonstrate that a considerable flattening of the C4n+2 perimeter is achievable in the case of a syn or all-syn arrangement of the oxygen bridges, and that the requirement for aromaticity is thus satisfied. This is confirmed in a striking manner by the synthesis and properties of syn-1,6; 8,13-bisoxido[14]annulene.  相似文献   

6.
The reactions of the phosphaethynolate anion ([PCO]) with a range of boranes were explored. BPh3 and [PCO] form a dimeric anion featuring P−B bonds and is prone to dissociation at room temperature. The more Lewis acidic borane B(C6F5)3 yields a less symmetric dimer of [PCO] with P−B and P−O bonds. Less sterically demanding HB(C6F5)2 and H2B(C6F5) boranes form a third isomer with [PCO] featuring both boranes bound to the same phosphorus atom. Despite the unexpected thermodynamic preference for P‐coordination, computational data illustrate that electronic and steric features impact the binding modes of the resulting dianionic dimers.  相似文献   

7.
Density functional theory, B3LYP/6‐31G** and B3LYP/6‐311+G(2d,p), and ab initio MP2/6‐31G** calculations have been carried out to investigate the conformers, transition states, and energy barriers of the conformational processes of oxalic acid and its anions. QCISD/6‐31G** geometrical optimization is also performed in the stable forms. Its calculated energy differences between the two most stable conformers are very near to the related observed value at 7.0 kJ/mol. It is found that the structures and relative energies of oxalic acid conformers predicted by these methods show similar results, and that the conformer L1 (C2h) with the double‐interfunctional‐groups hydrogen bonds is the most stable conformer. The magnitude of hydrogen bond energies depends on the energy differences of various optimized structures. The hydrogen bond energies will be about 32 kJ/mol for interfunctional groups, 17 kJ/mol for weak interfunctional groups, 24 kJ/mol for intra‐COOH in (COOH)2, and 60 kJ/mol for interfunctional groups in (COOH)COO−1 ion if calculated using the B3LYP/6‐311+G(2d,p) method. © 2000 John Wiley & Sons, Inc. Int J Quant Chem 76: 541–551, 2000  相似文献   

8.
A new class of propel‐ ler‐shaped compound ( 4 ), which consisted of dehydrobenzo[14]annulene ([14]DBA) blades, as well as its naphtho homologues ( 5 and 6 ), have been prepared. Although NMR studies of compound 4 did not provide useful information regarding its conformation in solution, DFT calculations with different functionals and the 6‐31G* basis set all indicated that the D3‐symmetric structure was energetically more favorable than the C2 conformer. From X‐ray crystallographic analysis, it appeared that compound 4 adopted a propeller‐shaped‐, approximately D3‐symmetric structure in the solid state, in which the [14]DBA blades were twisted substantially owing to steric repulsion between the neighboring benzene rings. On the contrary, in the case of compound 6 , although the DFT calculations with the B3LYP functional predicted that the D3‐symmetric conformation was more stable, calculations with the M05 and M05‐2X functionals indicated that the C2 conformer was more favorable because of π–π interactions between the naphthalene units of a pair of neighboring blades. Indeed, X‐ray analysis of compound 6 showed that it adopted an approximately C2‐symmetric conformation. Moreover, on the basis of variable‐temperature 1H NMR measurements, we found that compound 6 adopted a C2 conformation and the barrier for interconversion between the C2C2 conformers was estimated to be 16.2 kcal mol?1; however, no indication of the presence of the D3 isomer was obtained. The relatively small energy barriers to interconversion, despite the large overlapping of neighboring blades, was ascribed to the flexibility of the acetylene linkages, which could be deformed substantially in the transition state of the ring‐flip.  相似文献   

9.
The rod‐like molecule of the title hydro­carbon, C24H18, is centrosymmetric, with the centroid of the central benzene ring residing on an inversion center. The molecules display a planar conformation of the benzene rings and aggregate into stacks along the [010] direction via Csp3—H⋯π(arene) interactions, thus forming a stair‐like pseudo‐two‐dimensional network. Each molecule acts as both a C—H hydrogen donor and a π‐arene acceptor, forming four hydrogen bonds per molecule.  相似文献   

10.
11.
Diepoxy[18]annulenes(10.0): ( Z , E , Z , E , Z )‐Diepoxy[18]annulene(10.0) – a Highly Dynamic Annulene The McMurry reaction of (all‐E)‐5,5′‐([2,2′‐bifuran]‐5,5′‐diyl)bis[penta‐2,4‐dienal] ( 13 ) only occurs intramolecularly to give a mixture of the diepoxy[18]annulenes(10.0) 6 and 7 . Tetraepoxy[36]annulene(10.0.10.0) resulting from an intermolecular McMurry reaction is not formed. According to spectroscopic data, 6 is (Z,E,Z,E,Z)‐ and 7 (Z,E,E,Z,E)‐configured. The 1H‐NMR data confirm that in 6 the (E)‐ethene‐1,2‐diyl bonds (C(11)=C(12) and C(15)=C(16)) rotate around the adjacent σ‐bonds. Beginning at −70°, this rotation freezes, and 6 is becoming a diatropic aromatic ring system. Beside [18]annulene itself, (Z,E,Z,E,Z)‐diepoxy[18]annulene(10.0) 6 is the only hitherto known [18]annulene derivative with dynamic properties.  相似文献   

12.
Multiple bonds between boron and transition metals are known in many borylene (:BR) complexes via metal dπ→BR back‐donation, despite the electron deficiency of boron. An electron‐precise metal–boron triple bond was first observed in BiB2O? [Bi≡B?B≡O]? in which both boron atoms can be viewed as sp‐hybridized and the [B?BO]? fragment is isoelectronic to a carbyne (CR). To search for the first electron‐precise transition‐metal‐boron triple‐bond species, we have produced IrB2O? and ReB2O? and investigated them by photoelectron spectroscopy and quantum‐chemical calculations. The results allow to elucidate the structures and bonding in the two clusters. We find IrB2O? has a closed‐shell bent structure (Cs, 1A′) with BO? coordinated to an Ir≡B unit, (?OB)Ir≡B, whereas ReB2O? is linear (C∞v, 3Σ?) with an electron‐precise Re≡B triple bond, [Re≡B?B≡O]?. The results suggest the intriguing possibility of synthesizing compounds with electron‐precise M≡B triple bonds analogous to classical carbyne systems.  相似文献   

13.
In the title salt, (C6H8N4)[Mn(C14H8O4)2(C6H6N4)2]·6H2O, the MnII atom lies on an inversion centre and is coordinated by four N atoms from two 2,2′‐biimidazole (biim) ligands and two O atoms from two biphenyl‐2,4′‐dicarboxylate (bpdc) anions to give a slightly distorted octahedral coordination, while the cation lies about another inversion centre. Adjacent [Mn(bpdc)2(biim)2]2− anions are linked via two pairs of N—H...O hydrogen bonds, leading to an infinite chain along the [100] direction. The protonated [H2biim]2+ moiety acts as a charge‐compensating cation and space‐filling structural subunit. It bridges two [Mn(bpdc)2(biim)2]2− anions through two pairs of N—H...O hydrogen bonds, constructing two R22(9) rings, leading to a zigzag chain in the [2] direction, which gives rise to a ruffled set of [H2biim]2+[Mn(bpdc)2(biim)2]2− moieties in the [01] plane. The water molecules give rise to a chain structure in which O—H...O hydrogen bonds generate a chain of alternating four‐ and six‐membered water–oxygen R42(8) and R66(12) rings, each lying about independent inversion centres giving rise to a chain along the [100] direction. Within the water chain, the (H2O)6 water rings are hydrogen bonded to two O atoms from two [Mn(bpdc)2(biim)2]2− anions, giving rise to a three‐dimensional framework.  相似文献   

14.
The title compound, (S)‐(+)‐4‐[5‐(2‐oxo‐4,5‐di­hydro­imidazol‐1‐yl­sulfonyl)­indolin‐1‐yl­carbonyl]­anilinium chloride (S)‐(+)‐1‐[1‐(4‐amino­benzoyl)­indoline‐5‐sulfonyl]‐4‐phenyl‐4,5‐di­hydro­imidazol‐2‐one, C24H23N4O4S+·Cl?·C24H22N4O4S, crystallizes in space group C2 from a CH3OH/CH2Cl2 solution. In the crystal structure, there are two different conformers with their terminal C6 aromatic rings mutually oriented at angles of 67.69 (14) and 61.16 (15)°. The distances of the terminal N atoms (of the two conformers) from the chloride ion are 3.110 (4) and 3.502 (4) Å. There are eight distinct hydrogen bonds, i.e. four N—H?Cl, three N—H?O and one N—H?N, with one N—H group involved in a bifurcated hydrogen bond with two acceptors sharing the H atom. C—H?O contacts assist in the overall hydrogen‐bonding process.  相似文献   

15.
The reaction of [Cp*2RuBr]+Br with bromine in CH2Cl2 (CD2Cl2) in an inert atmosphere at room temperature produces the complexes [Cp*Ru(Br)C5Me4CH2Br]+Br3 (syn conformer), [Cp*Ru(Br)C5Me3(CH2Br)2]+ (syn and anti conformers), and [Ru(Br)(C5Me4CH2Br)2]+ (syn conformer). All complexes were characterized by 1H and 13C NMR spectroscopy; the former complex, by elemental analysis. These complexes were also prepared by the reaction of [Cp*RuC5Me4CH2]+BF4 with bromine in CH2Cl2. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2712–2718, December, 2005.  相似文献   

16.
The far infrared spectrum [350 to 25 cm–1] of gaseous chloroacetaldehyde, ClCH2CHO, has been recorded at a resolution of 0.10 cm–1. The first excited-state transition of the asymmetric torsion of the more stable near s-cis [chlorine atom s-cis to the aldehyde hydrogen atom] conformer has been observed at 26.9 cm–1, with seven additional upper state transitions falling to higher frequency. Additionally, the fundamental torsional transition of the s-trans conformer has been observed at 58.9 cm–1 with two excited states also falling to higher frequency. From these data, the asymmetric torsional potential coefficients have been determined to be:V 1=414±11;V 2 = 191±3;V 3=–203±5;V 4=44±1 andV 6=–26±1 cm–1. The s-cis to s-trans barrier is 500±5 cm–1 (1.43±0.01 kcal mol–1) with the s-cis conformer being more stable by 267±19 cm–1 (0.76±0.05 kcal mol–1) than the s-trans form. The Raman [4000 to 100 cm–1] and infrared (4000 to 400 cm–1] spectra of the gas have been recorded. Additionally, the Raman spectrum of the liquid has been recorded and qualitative depolarization values obtained. Complete vibrational assignments are proposed for both conformers based on band contours, depolarization values, and group frequencies. The assignments are supported by ab initio Hartree-Fock gradient calculations employing the 3–21G* basis set to obtain the frequencies and the potential energy distributions for the normal vibrations for both rotamers. Additional ab initio calculations at the MP4/6-31G* level have been carried out to determine the structural parameters for both conformers. The results are discussed and compared with the corresponding quantities obtained for some similar molecules.This contribution taken in part from the thesis of C. L. Tolley which will be submitted to the Department of Chemistry in partial fulfillment of the Ph.D. degree.  相似文献   

17.
The title compound, (C10H10N2)[CdBr4], was synthesized via a hydro­thermal reaction. Its structure features discrete 4,4′‐bipyridinium cations and tetra­hedral [CdBr4]2− anions linked into ion pairs by single N—H⋯Br hydrogen bonds. Photoluminescent investigation reveals that the title compound displays a strong emission in the blue region, which may originate from π→π* charge‐transfer inter­actions of the 4,4′‐bipyridinium cations.  相似文献   

18.
ESR. spectra of the radical anion (I?) produced from dimethyl-phenyl-phosphine (I) both by electrolysis and reaction with alkali metals have been studied upon variation of temperature. The coupling constant assigned to the 31P nucleus depends strongly on temperature, whereas the coupling constants attributed to protons do not exhibit such a dependence. The π-spin populations at the benzene ring of I? give evidence - in accordance with other experimental data [1] [2] – that the dimethylphosphino substituent is electron-attracting. This effect is thought to be due mainly to P ← Cπ delocalization, which is analogous to the Si? Cπ interaction in trimethylsilyl-substituted π-systems [3]. The ESR. spectrum previously [4] ascribed to I? is shown to arise from a secondary radical. The formation and structure of this radical are briefly discussed.  相似文献   

19.
[NMe4]2[TCNE]2 (TCNE=tetracyanoethenide) formed from the reaction of TCNE and (NMe4)CN in MeCN has νCN IR absorptions at 2195, 2191, 2172, and 2156 cm?1 and a νCC absorption at 1383 cm?1 that are characteristic of reduced TCNE. The TCNEs have an average central C?C distance of 1.423 Å that is also characteristic of reduced TCNE. The reduced TCNE forms a previously unknown non‐eclipsed, centrosymmetric π‐[TCNE]22? dimer with nominal C2 symmetry, 12 sub van der Waals interatomic contacts <3.3 Å, a central intradimer separation of 3.039(3) Å, and comparable intradimer C???N distances of 3.050(3) and 2.984(3) Å. The two pairs of central C???C atoms form a ?C?C???C?C of 112.6° that is substantially greater than the 0° observed for the eclipsed D2h π‐[TCNE]22? dimer possessing a two‐electron, four‐center (2e?/4c) bond with two C???C components from a molecular orbital (MO) analysis. A MO study combining CAS(2,2)/MRMP2/cc‐pVTZ and atoms‐in‐molecules (AIM) calculations indicates that the non‐eclipsed, C2 π‐[TCNE]22? dimer exhibits a new type of a long, intradimer bond involving one strong C???C and two weak C???N components, that is, a 2e?/6c bond. The C2 π‐[TCNE]22? conformer has a singlet, diamagnetic ground state with a thermally populated triplet excited state with J/kB=1000 K (700 cm?1; 86.8 meV; 2.00 kcal mol?1; H=?2 JSa?Sb); at the CAS(2,2)/MBMP2 level the triplet is computed to be 9.0 kcal mol?1 higher in energy than the closed‐shell singlet ground state. The results from CAS(2,2)/NEVPT2/cc‐pVTZ calculations indicate that the C2 and D2h conformers have two different local metastable minima with the C2 conformer being 1.3 kcal mol?1 less stable. The different natures of the C2 and D2h conformers are also noted from the results of valence bond (VB) qualitative diagram that shows a 10e?/6c bond with one C???C and two C???N bonding components for the C2 conformer as compared to the 6e?/4c bond for the D2h conformer with two C???C bonding components.  相似文献   

20.
(2R,4S)‐2‐(3‐Methylthiophen‐2‐yl)‐2,3,4,5‐tetrahydro‐1,4‐epoxynaphtho[1,2‐b]azepine, C19H17NOS, (I), crystallizes with a single enantiomer in each crystal, whereas its geometrical isomer (2RS,4SR)‐2‐(5‐methylthiophen‐2‐yl)‐2,3,4,5‐tetrahydro‐1,4‐epoxy‐naphtho[1,2‐b]azepine, (II), and (2RS,4SR)‐2‐(5‐bromothiophen‐2‐yl)‐2,3,4,5‐tetrahydro‐1,4‐epoxynaphtho[1,2‐b]azepine, C18H14BrNOS, (III), both crystallize as racemic mixtures. A combination of one C—H...O hydrogen bond and two C—H...π(arene) hydrogen bonds links the molecules of (I) into a three‐dimensional framework; the molecules of (II) are linked into a C(4)C(4)[R22(7)] chain of rings by a combination of C—H...N and C—H...O hydrogen bonds; and in (III), where Z′ = 2, a combination of four C—H...π(arene) hydrogen bonds and two C—H...π(thienyl) hydrogen bonds links the molecules into complex sheets. Comparisons are made with the assembly patterns in some aryl‐substituted 1,4‐epoxynaphtho[1,2‐b]azepines.  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