共查询到20条相似文献,搜索用时 15 毫秒
1.
Gerald R. Willey Miles T. Lakin Nathaniel W. Alcock Christopher J. Samuel 《Journal of inclusion phenomena and macrocyclic chemistry》1994,15(3):293-304
The X-ray crystal structure of 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane has been determined. The neutral ligand crystallises in the triclinic space groupP
with unit cell parametersa=8.467(4),b=10.057(5),c=10.637(5) Å, =67.68(3), =75.38(4), =76.34(4)°,V=800.75 Å3 andD
c=1.063 g cm–3 forZ=2.R=0.047 for 2140 unique observed (I/(I)2.0) reflections (R
w=0.052). Two crystallographically independent molecules are observed. Both are centrosymmetric and adopt a mesodentate conformation. Molecule (1) is identified as the (R, S, S, R)-[trans-IV] isomer and molecule (2) as the (R, R, S, S)-[trans-III] isomer. Molecular mechanics calculations provide a basis for understanding the structural and energetic relationships between the neutral and protonated forms of this tetraazamacrocyclic ligand.
Supplementary Data relating to this article have been deposited with the British Library at Boston Spa, Wetherby, West Yorkshire, U.K. as Supplementary Publication No. 82149 (13 pages). 相似文献
2.
Dohyun Moon Yong Pyo Hong Jong-Ha Choi 《Acta Crystallographica. Section C, Structural Chemistry》2016,72(9):701-704
Constrained cyclam derivatives have been found to exhibit anti‐HIV effects. The strength of binding to the CXCR4 receptor correlates with anti‐HIV activity. The conformation of the macrocyclic compound is very important for co‐receptor recognition. Therefore, knowledge of the conformation and crystal packing of macrocycles has become important in developing new highly effective anti‐HIV drugs. Structural modifications of N‐functionalized polyaza macrocyclic compounds have been achieved using various methods. A new synthesis affording single crystals of the title tetraazapentacyclo[16.4.0.12,17.16,13.07,12]tetracosane macrocycle, C22H40N4, is reported. Formaldehyde reacts readily at room temperature with the tetraazatricyclo[16.4.0.02,17]docosane precursor to yield a macropolycycle containing two five‐membered rings. Characterization by elemental, spectroscopic and single‐crystal X‐ray diffraction analyses shows that the asymmetric unit contains half of a centrosymmetric molecule. The molecular structure shows a trans conformation for the two methylene bridges owing to molecular symmetry. The crystal structure is stabilized by intramolecular C—H…N hydrogen bonds. NMR and IR spectroscopic properties support the methylene‐bridged macrocyclic structure. 相似文献
3.
Giulia Festa Claudia Scatigno Francesco Armetta Maria Luisa Saladino Veronica Ciaramitaro Viviana Mollica Nardo Rosina Celeste Ponterio 《Molecules (Basel, Switzerland)》2022,27(1)
Spectral preprocessing data and chemometric tools are analytical methods widely applied in several scientific contexts i.e., in archaeometric applications. A systematic classification of natural powdered pigments of organic and inorganic nature through Principal Component Analysis with a multi-instruments spectroscopic study is presented here. The methodology allows the access to elementary and molecular unique benchmarks to guide and speed up the identification of an unknown pigment and its recipe. This study is conducted on a set of 48 powdered pigments and tested on a real-case sample from the wall painting in S. Maria Delle Palate di Tusa (Messina, Italy). Four spectroscopic techniques (X-ray Fluorescence, Raman, Attenuated Total Reflectance and Total Reflectance Infrared Spectroscopies) and six different spectrometers are tested to evaluate the impact of different setups. The novelty of the work is to use a systematic approach on this initial dataset using the entire spectroscopic energy range without any windows selection to solve problems linked with the manipulation of large analytes/materials to find an indistinct property of one or more spectral bands opening new frontiers in the dataset spectroscopic analyses. 相似文献
4.
Photolysis of hexacarbonylchromium(0) in the presence of 2,6-diaminopyridine in toluene solution at 10 °C yields pentacarbonyl(2,6-diaminopyridine)chromium(0), which could be isolated from solution as plate-like crystals and fully characterized by using the single crystal X-ray diffractometry and MS, IR, 1H and 13C NMR spectroscopy. The complex was found to have the 2,6-diaminopyridine ligand bonded to the chromium atom through one of the NH2 groups. A single crystal X-ray structure of the complex reveals that the coordination sphere around the chromium atom is a slightly distorted octahedron, involving five carbonyls and one 2,6-diaminopyridine ligand. Because of the steric requirement of 2,6-diaminopyridine the four equatorial carbonyl groups are bended away from the N-donor ligand. The pyridine plane makes an angle of 112.9(3)° with the OC-Cr-N bond axis. The Cr-C distances have values between 1.833(7) and 1.935(7) Å. The Cr-N distance is 2.236(5) Å. 相似文献
5.
《Physics and Chemistry of Liquids》2012,50(5):487-501
Melting of aluminum and iron trichloride is accompanied by a structural transition from sixfold to fourfold coordination of the trivalent metal ions, and a widely accepted interpretation of the structure of their melts near freezing is that they mainly consist of strongly correlated dimers formed from two edge-sharing tetrahedra. We carry out classical molecular dynamics simulations to examine how a polarizable-ion force law, determined on isolated molecular monomers and dimers in the gaseous phase of these compounds, fares in accounting for the pair structure of their liquid phase and for mean square displacements and diffusion coefficients of the two species in each melt. The model reproduces the main features of the neutron diffraction structure factor, showing peaks due to intermediate range order and to charge and density short-range order, and accounts for the experimental data at a good semi-quantitative level. We find agreement with the neutron and X-ray diffraction data on metal–halogen and Cl–Cl bond lengths in the melt, and demonstrate the high sensitivity of the results for the width of the first-neighbor shell to truncation in obtaining it by Fourier transform of the neutron-weighted structure factor in momentum space. We also report comparisons with a recent first-principles study of the structure of the AlCl3 melt by the Car–Parrinello method. Finally, we demonstrate break-up of dimers into monomers upon raising the liquid temperature in the case of AlCl3. 相似文献
6.
Kochikpa Y. A. Okio Libasse Diop Aliou H. Barry Tahar Jouini Simn H. Ortega 《应用有机金属化学》2003,17(5):317-318
The structure of NMe4SnPh3(NCS)2 is molecular, without any significant intermolecular contacts in the lattice. The trigonal plane around the tin atom is defined by three carbon atoms from the phenyl groups and the axial positions occupied by the NCS groups. Copyright © 2003 John Wiley & Sons, Ltd. 相似文献
7.
Lina M. Acosta Quintero Alirio Palma Justo Cobo Christopher Glidewell 《Acta Crystallographica. Section C, Structural Chemistry》2016,72(4):346-357
A versatile synthetic method has been developed for the formation of variously substituted polycyclic pyrimidoazepine derivatives, formed by nucleophilic substitution reactions on the corresponding chloro‐substituted compounds; the reactions can be promoted either by conventional heating in basic solutions or by microwave heating in solvent‐free systems. Thus, (6RS)‐6,11‐dimethyl‐3,5,6,11‐tetrahydro‐4H‐benzo[b]pyrimido[5,4‐f]azepin‐4‐one, C14H15N3O, (I), was isolated from a solution containing (6RS)‐4‐chloro‐8‐hydroxy‐6,11‐dimethyl‐6,11‐dihydro‐5H‐benzo[b]pyrimido[5,4‐f]azepine and benzene‐1,2‐diamine; (6RS)‐4‐butoxy‐6,11‐dimethyl‐6,11‐dihydro‐5H‐benzo[b]pyrimido[5,4‐f]azepin‐8‐ol, C18H23N3O2, (II), was formed by reaction of the corresponding 6‐chloro compound with butanol, and (RS)‐4‐dimethylamino‐6,11‐dimethyl‐6,11‐dihydro‐5H‐benzo[b]pyrimido[5,4‐f]azepin‐8‐ol, C16H20N4O, (III), was formed by reaction of the chloro analogue with alkaline dimethylformamide. (6RS)‐N‐Benzyl‐8‐methoxy‐6,11‐dimethyl‐6,11‐dihydro‐5H‐benzo[b]pyrimido[5,4‐f]azepin‐4‐amine, C22H24N4O, (IV), (6RS)‐N‐benzyl‐6‐methyl‐1,2,6,7‐tetrahydropyrimido[5′,4′:6,7]azepino[3,2,1‐hi]indol‐8‐amine, C22H22N4, (V), and (7RS)‐N‐benzyl‐7‐methyl‐2,3,7,8‐tetrahydro‐1H‐pyrimido[5′,4′:6,7]azepino[3,2,1‐ij]quinolin‐9‐amine, C23H24N4, (VI), were all formed by reaction of the corresponding chloro compounds with benzylamine under microwave irradiation. In each of compounds (I)–(IV) and (VI), the azepine ring adopts a conformation close to the boat form, with the C‐methyl group in a quasi‐equatorial site, whereas the corresponding ring in (V) adopts a conformation intermediate between the twist‐boat and twist‐chair forms, with the C‐methyl group in a quasi‐axial site. No two of the structures of (I)–(VI) exhibit the same range of intermolecular hydrogen bonds: different types of sheet are formed in each of (I), (II), (V) and (VI), and different types of chain in each of (III) and (IV). 相似文献
8.
Makio Iwahashi Yasutoshi Kasahara 《Current Opinion in Colloid & Interface Science》2011,16(5):359-366
This paper discusses the molecular conformations and the liquid structures of triacylglycerols (TGs) and fatty acids in their melts. Three models for liquid state ordering have been proposed for TG melts to date: the smectic liquid crystal model, the nematic liquid crystal model, and the discotic model. To completely resolve the liquid structure of TGs, further research is required. However, some information on the molecular level has been obtained for fatty acids that are relatively simple compounds. The combination of various spectroscopic and thermodynamic measurements revealed that the hydrogen-bonded dimers of fatty acids are units of intermolecular and intramolecular movements in the liquids and in non-polar solvents. The dimers that construct the clusters resemble the smectic liquid crystal and determine the physicochemical properties of the liquid of the fatty acid. Cholesterol stabilizes the clusters, while ethanol destroys them. Self-diffusion and neutron diffraction measurements revealed that two kinds of fatty acids exist in their binary liquid mixture exist as the homodimers composed of same species. 相似文献
9.
A very brief review of gas-phase electron diffraction and one of its offshoots is given. Parallels are drawn between experimental studies of molecules, including conformational changes, and studies of clusters, including phase changes, calling particular attention to the use of computers as the preferred experimental apparatus. A sketch is presented of what has been learned about matter in transition by the application of computer simulations. 相似文献
10.
Sergio A. Guerrero Juan E. Ramírez Alirio Palma Justo Cobo Christopher Glidewell 《Acta Crystallographica. Section C, Structural Chemistry》2019,75(2):168-177
A concise, efficient and versatile route from simple starting materials to tricyclic tetrahydro‐1‐benzazepines carrying [a]‐fused heterocyclic units is reported. Thus, the easily accessible methyl 2‐[(2‐allyl‐4‐chlorophenyl)amino]acetate, (I), was converted, via (2RS,4SR)‐7‐chloro‐2,3,4,5‐tetrahydro‐1,4‐epoxy‐1‐benzo[b]azepine‐2‐carboxylate, (II), to the key intermediate methyl (2RS,4SR)‐7‐chloro‐4‐hydroxy‐2,3,4,5‐tetrahydro‐1H‐benzo[b]azepine‐2‐carboxylate, (III). Chloroacetylation of (III) provided the two regioisomers methyl (2RS,4SR)‐7‐chloro‐1‐(2‐chloroacetyl)‐4‐hydroxy‐2,3,4,5‐tetrahydro‐1H‐benzo[b]azepine‐2‐carboxylate, (IVa), and methyl (2RS,4SR)‐7‐chloro‐4‐(2‐chloroacetoxy)‐2,3,4,5‐tetrahydro‐1H‐benzo[b]azepine‐2‐carboxylate, C14H15Cl2NO4, (IVb), as the major and minor products, respectively, and further reaction of (IVa) with aminoethanol gave the tricyclic target compound (4aRS,6SR)‐9‐chloro‐6‐hydroxy‐3‐(2‐hydroxyethyl)‐2,3,4a,5,6,7‐hexahydrobenzo[f]pyrazino[1,2‐a]azepine‐1,4‐dione, C15H17ClN2O4, (V). Reaction of ester (III) with hydrazine hydrate gave the corresponding carbohydrazide (VI), which, with trimethoxymethane, gave a second tricyclic target product, (4aRS,6SR)‐9‐chloro‐6‐hydroxy‐4a,5,6,7‐tetrahydrobenzo[f][1,2,4]triazino[4,5‐a]azepin‐4(3H)‐one, C12H12ClN3O2, (VII). Full spectroscopic characterization (IR, 1H and 13C NMR, and mass spectrometry) is reported for each of compounds (I)–(III), (IVa), (IVb) and (V)–(VII), along with the molecular and supramolecular structures of (IVb), (V) and (VII). In each of (IVb), (V) and (VII), the azepine ring adopts a chair conformation and the six‐membered heterocyclic rings in (V) and (VII) adopt approximate boat forms. The molecules in (IVb), (V) and (VII) are linked, in each case, into complex hydrogen‐bonded sheets, but these sheets all contain a different range of hydrogen‐bond types: N—H…O, C—H…O, C—H…N and C—H…π(arene) in (IVb), multiple C—H…O hydrogen bonds in (V), and N—H…N, O—H…O, C—H…N, C—H…O and C—H…π(arene) in (VII). 相似文献
11.
Geoffrey D. Hietpas David L. Allara 《Journal of Polymer Science.Polymer Physics》1998,36(7):1247-1260
The molecular structure of poly[biphenyl dianhydride-p-phenylenediamine] (BPDA–PDA) polyimide in ultrathin (3–300 nm) films on silicon has been characterized by polarized infrared spectroscopy in conjunction with ellipsometry and X-ray reflectivity measurements. In spite of the high degree of crystalline packing of the polymer chains, the results show that an unexpected and significant content of imide rings exhibit local structural perturbations, including out-of-plane twisting. Further, the fraction of perturbed rings increases with increasing film thickness while, in contrast, the high degree of in-plane uniaxial film symmetry and planar stacking of the chains remain constant with thickness. These results reveal a new structural aspect of localized ring disorder that arises within the otherwise well-ordered, chain-stacked structure of BPDA–PDA polyimide films. © 1998 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 36: 1247–1260, 1998 相似文献
12.
The phospholipid bilayer plays a central role in the lifecycle of the endogenous cannabinoid, N-arachidonoylethanolamine (anandamide, AEA). Therefore, the orientation and location of AEA in the phospholipid bilayer with respect to key membrane associated proteins, is a central issue in understanding the mechanism of endocannabinoid signaling. In this paper, we report a test of the hypothesis that a βXXβ motif (formed by beta branching amino acids, V6.43 and I6.46) on the lipid face of the cannabinoid CB1 receptor in its inactive state may serve as an initial CB1 interaction site for AEA. Eight 6 ns NAMD2 molecular dynamics simulations of AEA were conducted in a model system composed of CB1 transmembrane helix 6 (TMH6) in a 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) bilayer. In addition, eight 6 ns NAMD2 molecular dynamics simulations of a low CB1 affinity (20:2, n−6) analog of AEA were conducted in the same model system. AEA was found to exhibit a higher incidence of V6.43/I6.46 groove insertion than did the (20:2, n−6) analog. In certain cases, AEA established a high energy of interaction with TMH6 by first associating with the V6.43/I6.46 groove and then molding itself to the lipid face of TMH6 to establish a hydrogen bonding interaction with the exposed backbone carbonyl of P6.50. Based upon these results, we propose that the formation of this hydrogen bonded AEA/TMH6 complex may be the initial step in CB1 recognition of AEA in the lipid bilayer. 相似文献
13.
Prakash S. Nayak Jerry P. Jasinski James A. Golen Badiadka Narayana Manpreet Kaur Hemmige S. Yathirajan Christopher Glidewell 《Acta Crystallographica. Section C, Structural Chemistry》2014,70(9):889-894
The structures of four halogenated N,2‐diarylacetamides are reported and compared with a range of analogues. N‐(4‐Chloro‐3‐methylphenyl)‐2‐phenylacetamide, C15H14ClNO, (I), and N‐(4‐bromo‐3‐methylphenyl)‐2‐phenylacetamide, C15H14BrNO, (II), are isostructural in the space group P. The molecules of (I) and (II) are linked into chains of rings by a combination of N—H...O and C—H...π(arene) hydrogen bonds. The molecules of N‐(4‐chloro‐3‐methylphenyl)‐2‐(2,4‐dichlorophenyl)acetamide, C15H12Cl3NO, (III), and N‐(4‐bromo‐3‐methylphenyl)‐2‐(2‐chlorophenyl)acetamide, C15H13BrClNO, (IV), are linked into simple C(4) chains by N—H...O hydrogen bonds, but significant C—H...π(arene) interactions are absent. The N‐aryl groups in compounds (III) and (IV) adopt a different orientation, by ca 180°, from that of the corresponding groups in compounds (I) and (II), but otherwise the conformations of (I)–(IV) are very similar. Comparisons are drawn between compounds (I) and (IV) and a range of analogues of the type R1CH2CONHR2, where R2 represents a halogenated aryl ring and R1 represents either another halogenated aryl ring or a naphthalen‐1‐yl unit. 相似文献
14.
The crystal and molecular structure of 1-chloromethyl-1-oxo-1,3-dihydro-2,1-benzoxaphosphole was determined by x-ray diffraction structural analysis. This compound has an almost planar bicyclic unit with extruding phosphorus atoms and atrans Cl-C-P=O fragment.Institute of Physiologically Active Compounds, Russian Academy of Sciences, 142432 Chernogolovka. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 7, pp. 1670–1673, July, 1992. 相似文献
15.
Badiadka Narayana Hemmige S. Yathirajan Ravindranath S. Rathore Christopher Glidewell 《Acta Crystallographica. Section C, Structural Chemistry》2016,72(9):664-669
4‐Antipyrine [4‐amino‐1,5‐dimethyl‐2‐phenyl‐1H‐pyrazol‐3(2H)‐one] and its derivatives exhibit a range of biological activities, including analgesic, antibacterial and anti‐inflammatory, and new examples are always of potential interest and value. 2‐(4‐Chlorophenyl)‐N‐(1,5‐dimethyl‐3‐oxo‐2‐phenyl‐2,3‐dihydro‐1H‐pyrazol‐4‐yl)acetamide, C19H18ClN3O2, (I), crystallizes with Z′ = 2 in the space group P, whereas its positional isomer 2‐(2‐chlorophenyl)‐N‐(1,5‐dimethyl‐3‐oxo‐2‐phenyl‐2,3‐dihydro‐1H‐pyrazol‐4‐yl)acetamide, (II), crystallizes with Z′ = 1 in the space group C2/c; the molecules of (II) are disordered over two sets of atomic sites having occupancies of 0.6020 (18) and 0.3980 (18). The two independent molecules of (I) adopt different molecular conformations, as do the two disorder components in (II), where the 2‐chlorophenyl substituents adopt different orientations. The molecules of (I) are linked by a combination of N—H…O and C—H…O hydrogen bonds to form centrosymmetric four‐molecule aggregates, while those of (II) are linked by the same types of hydrogen bonds forming sheets. The related compound N‐(1,5‐dimethyl‐3‐oxo‐2‐phenyl‐2,3‐dihydro‐1H‐pyrazol‐4‐yl)‐2‐(3‐methoxyphenyl)acetamide, C20H21N3O3, (III), is isomorphous with (I) but not strictly isostructural; again the two independent molecules adopt different molecular conformations, and the molecules are linked by N—H…O and C—H…O hydrogen bonds to form ribbons. Comparisons are made with some related structures, indicating that a hydrogen‐bonded R22(10) ring is the common structural motif. 相似文献
16.
The motivation for this study was the interesting, and contrasting, reported crystallization modes of triphenylmethane and triphenylsilane, the former said to crystallize as a conglomerate [Riche and Pascard-Billy,Acta Crystallog. B
1974,30, 1874], whereas the latter was reported to crystallize as a racemate [Allemand and Gerdi,Cryst. Struct. Commun.
1979,8, 927]. As we show in what follows, we had planned to justify these observations via molecular mechanics and semiempirical (AM1) calculations. However, in the process of doing some geometrical calculations unavailable in the original, we found that it was impossible for the reported structure [Riche and Pascard-Billy,op. cit.] to be correct. Therefore, we redetermined the structure and found that triphenylmethane also crystallizes as a racemate. The compound was recrystallized from toluene and its structure was solved in space groupPna21 (No. 33). The cell constants area=25.593(14),b=14.804(19),c=7.521(15) å;V=2849.5(6) å3,d(MW=244.32 g mol–1;z=8)=1.139 g cm–3. A total of 2183 independent reflections were collected over the range 4 2 40. The structure was solved by direct methods and refined using rigid-body constraints for the three phenyl rings as a result of the large, and unfavorable, ratio of variables/observed data had the individual atoms been refined. Molecular mechanics (HyperChem) calculations were carried out to ascertain the degree of agreement between the observed structural parameters and those obtained from an energy-minimized structure. The agreement is remarkably good, especially given the quality of the X-ray data. Similar comments can be applied to the results obtained by semiempirical (AM1) calculations (in HyperChem). Thus, this is a satisfying agreement lending weight to the suggestion that the X-ray data are very reasonable, even if flawed. Calculations were also carried out to determine a conformational potential surface for the molecule as a function of rotation of the phenyl rings around their HC-C(phenyl) vector. These calculations indicate a shallow minimum around the conformation with C3 symmetry with torsional angles [H-C-C(head)-C(ortho) of 30, 30, 30], as well as possible low-energy pathways between molecules of opposite axial chirality.See ref. 1 for previous papers of this series.on leave at the University of Houston, 1995–1996. 相似文献
17.
Ab initio and density functional calculations were used to analyze the interaction between a molecule of cyanic acid (HOCN) and up to 4 molecules of water at the B3LYP/6-311++G(d,p) and MP2/6-311++G(d,p) computational levels. The cooperative effect (CE) is increased with the increasing size of the studied clusters. Red shifts of the H–O stretching frequency for complexes involving HOCN as an H-donor were predicted. The strength of the hydrogen bonds in terms of molecular structures could be deduced from a comparison of HOCN–H2O with HCNO–H2O, HONC–H2O and HNCO–H2O HB clusters. The atom in molecules (AIM) method was used to analyze the cooperative effects on topological parameters. 相似文献
18.
Giuliano Freddi Patrizia Monti Masanobu Nagura Yohko Gotoh Masuhiro Tsukada 《Journal of Polymer Science.Polymer Physics》1997,35(5):841-847
Structural changes of tussah (Antheraea pernyi) silk fibroin films induced by heat treatment were studied as a function of the treatment temperature in the range 200–250°C. The DSC curve of tussah films with α-helix molecular conformation displayed characteristic endo and exo peaks at 216 and 226°C, respectively. These peaks first weakened and then completely disappeared after heating at 230°C. Accordingly, the TMA thermal shrinkage at 206°C disappeared when the films were heated at 230°C. The onset of weight loss was monitored at 210°C by means of TG measurements. X-ray diffraction profiles gradually changed from α-helix to β-sheet crystalline structure as the treatment temperature increased from 200 to 250°C. On raising the heating temperature above 200°C, the intensity of IR and Raman bands characteristic of β-sheet conformation increased in the whole ranges of amide and skeletal modes. The sample treated at 200°C showed a spectral pattern intermediate between α-helix and β-sheet molecular conformation. The IR marker band for random coil structure, still detectable at 200°C, disappeared at higher treatment temperatures. Spectral changes attributable to the onset of thermal degradation appeared at 230°C. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 841–847, 1997 相似文献
19.
Z. G. Aliev A. N. Maslivets O. P. Krasnykh Yu. S. Andreichikov L. O. Atovmyan 《Russian Chemical Bulletin》1993,42(9):1569-1572
Thermal decarboxylation of 3-p-toluoyl-1,2-dihydro-4H-pyrrolo[5,1-c][1,4]benzoxazine-1, 2,4-trione resulted in 2-oxo-2H-1,4-benzoxazin-3-yl(p-toluoyl)ketene; the latter under-goes [4+2]-cycloaddition withp-bromobenzaldehyde to form 1-p-bromophenyl-4-p-toluoyl-1, 3-dihydro-5H-[1,3]oxazino[4,3-c][1,4] benzoxazine-3,5-dione. The crystal and molecular structure of the latter has been studied by X-ray analysis.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1633–1636, September, 1993. 相似文献
20.
Thin films of semiconducting molecular materials can be grown with a seeded supersonic molecular beam epitaxy (SuMBE), which provides unprecedented control over structural, morphological, and, therefore, functional properties. This novel technique of deposition takes full advantage of its ability to regulate the initial state of the molecular precursors in the beams and, in particular, the kinetic energy, to control the morphology, structure, and functional properties of growing films. This article reviews the state of the art of SuMBE, discussing the basic aspects of the technique and the major achievements so far. The major results obtained with respect to growth on dielectrics and metal substrates of films of oligothiophenes and pentacene and with respect to the codeposition of phthalocyanines and fullerenes are discussed and compared with the state of the art of more conventional organic molecular beam deposition. The potential impact of SuMBE in the field of π‐conjugated materials and devices is also examined. © 2003 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 41: 2501–2521, 2003 相似文献