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 531 毫秒
1.
Abstract

Attempts to prepare 1H-1,2,4-triazol-1-ylmethylphosphonates (4 and 5) by a Mannichtype reaction or by transesterification of 1-hydroxymethyl-1H-1,2,4-triazol 1 with tertiary phosphites failed. On the other hand 4 and 5 are obtained by a Michaelis-Becker reaction from 1-chloromethyl-1H-1,2,4-triazol 3 and sodium phosphites in high yield. The Michaelis-Arbuzov reaction is less suited for the preparation of 4 and 5. 3 is obtained in good yield as a water clear liquid, b.p. 52–54°C/0.2 torr, from the interaction of 1 with thionyl chloride followed by treatment with a base. On standing at 0° or 20°C it decomposes within hours and yields the unsymmetrical methylen-bis(triazol) 3a in addition to other products. However an acetonitrile solution of 3 is stable for months. Heating this solution with tertiary phosphines gives triazolylsubstituted phosphoniumsalts 6 to 8. The Wittig-Horner reaction with 4 to 6 gives the olefinically substituted triazols 9–12 as a Z/E mixture in high yield. Alkylation of 4 with methyl-and ethyl iodide gives the corresponding alkylated diethyl-1H-1,2,4-triazol-1-yl-ethyl-1-and-propyl-1-phosphonates 14 and 15 which on hydrolysis with HCI yield 1H-1,2,4-triazol-1-yl-ethyl-1-and propyl-1-phosphonic acids 17 and 18, respectively. Hydrolysis of 4 gives the unsubstituted 1H-1,2,4-triazol-1-ylmethyl-phosphonic acid, 16.  相似文献   

2.
Abstract

The effects of phenyl and methyl groups at 3,5-positions of tetrasubstituted 4H-thiopyrans 1a and 1b on photoisomerization are investigated from a kinetic point of view using 1H-NMR spectroscopy. On irradiation of 1a-1d in benzene-d6 solution the hexasubstituted 4H-thi-opyrans 1c and 1d, unlike those of the tetrasubstituted analogues 1a and 1b, give the isomeric 2H-thiopyrans 3c and 3d with no detectable signals for intermediates in 1H-NMR spectra. The photoisomerization of hexasubstituted 4H-thiopyrans 1c and 1d occur with relative rate constants lower than the corresponding tetrasubstituted model compounds 1a and 1b. Moreover, the kinetic comparison of 1a with 1b reveal that the presence of two phenyl groups at 4-position of tetrasubstituted 4H-thiopyran increases the relative rate constant of photoisomerization.  相似文献   

3.
Abstract

The 1-acetoxy-1-methylthiomethanephosphoryl compounds (4) are prepared from methylthiomethanephosphoryl derivatives (1) by electrolysis in 0.6 molar sodium acetate/acetic acid solution. The reaction of 1-acetoxy compounds (4) with halogen carriers gives the 1-halogen-1-methylthiomethane-phosphoryl derivatives 5, 7 and 8. The reaction of 4a under PTC-conditions results in cleavage of the P-C-bond.

Die 1-Acetoxy-1-methylthiomethanphosphoryl-Verbindungen (4) wurden aus Methylthiomethanphos-phorylderivaten (1) durch Elektrolyse in 0.6 molarer Na-acetat/Essigsäure-Lösung dargestellt. Die Reaktion der 1-Acetoxyverbindungen mit Halogenüberträgern führt zu den 1-Halogen-1-methyl-thiomethanphosphorylderivaten 5, 7 und 8. Bei Reaktion von 4a unter PTC-Bedingungen beobachtet man eine Spaltung der P-C-Bindung.  相似文献   

4.
Abstract

Disodium 1, 6-disulfido-1, 3, 5-cycloheptatriene 3, formed by reduction of 1 with sodium in liq. ammonia, reacts with hydrogen chloride and methyl iodide to give 1, 6-dimercapto-, 4, and 1, 6-bis(methylthio)-1, 3, 5-cycloheptatrienes 5 respectively; however, it is oxidized by bromine to afford cyclic disulfide 6. 1, 6-Diiodo-1, 3, 5-cycloheptatriene 2 is converted to 1, 6-bis(benzylthio)-1, 3, 5-cycloheptatriene 7 by reaction with sodium phenylmethanethiolate, whereas similar reactions with 1-(2-hydroxyethyl)-6-iodo-1, 3, 5-cycloheptatriene 9, obtained from 2 via 1-iodo-6-vinyl-1, 3, 5-cycloheptatriene 8, give 1-benzylthio-6-(2-hydroxyethyl)-1, 3, 5-cycloheptatriene 10. 1-Benzylthio-6-benzylthioethyl-1, 3, 5-cycloheptatriene 11 is synthesized by the reaction from 9 via 1-(2-bromoethyl)-6-iodo-1, 3, 5-cycloheptatriene 10. Attempts to synthesize thiols from 7, 11, and 12 are also described.  相似文献   

5.
Abstract

1-Cyano-isothiochromane (1a) can be alkylated in position 1, using the carbanion that is formed from sodium amide, sodium hydride, or n-butyllithium. With methyl iodide or ethyl iodide 1c and 1d are formed; with α-halogenated ether or thioether, 1e and 1f; with propargyl bromide, 1h; with bromo acetophenone, li; and with ethyl chloroacetate, 1k. Similarly, acylation with benzoyl chloride leads to 11, and with 2,4-dinitrofluorobenzene to 1m. The alkylation products of 1a can be oxidized with peracids to the sulfones 2 and with LiAlH4 reduced to 1-aminomethyl-isothiochroman (3): Acid hydrolysis of 1 gave isothiochromane-1-carboxylic acids, 4, whereas when 1 is treated with hydrogen peroxide in alkaline medium the S-dioxide and the S-oxide acid amides, 5 and 6 respectively, are formed.  相似文献   

6.
Chemical investigation of the resinous exudates of Commiphora myrrha has led to the isolation of four sesquiterpenes (1a/1b, 2, and 3), including one pair of new sesquiterpene enantiomers (1a/1b), one new racemic mixture 2, and two steroids (4 and 5). Their structures were elucidated by spectroscopic analysis, and the absolute configurations of 1a/1b were determined by CD analysis. The antimigratory potential of compounds 15 were evaluated and compound 3 was found to inhibit human hepatocellular liver carcinoma HepG2 cell migration in dose-dependent manner.  相似文献   

7.
Repetition of the work ofSugino andTamaka 1 showed that acrylonitrile and guanidine react inDMF to yield not only 3,4,6,7-tetrahydro-2H-pyrimido[1,2—a]pyrimidine-2,8(1H)-diimine (1), but a mixture (F) of1 (as a main product) and 2-amino-4-imino-1,4,5,6-tetrahydropyrimidine-1-propionitrile (7) besides one or two bases not identified so far.1 and7 were isolated as picrates. For the prove of their structures,1- and7-picrate were also prepared by an unequivocal synthesis starting from iminodipropionitrile hydrochloride8 · HCl: The latter on reaction with cyanamide gave9 which cyclized to afford a mixture of1,7 and 2-amino-4-oxo-1,4,5,6-tetrahydropyrimidine-1-propionitrile (10). The picrates of1 and7 were identical with those prepared from the acrylonitrileguanidine-condensateF. This result supports the prior proposed1 structures of pyrimidopyrimidine1 and of4,5 and6, obtained by hydrolysis of1. Nmr-, ir-and some of the mass spectra of1,4,7–10 (and their salts) are reported.
  相似文献   

8.
The polymeric precursor [RuCl2(CO)2]n reacts with the ligands, P∩P (a, b) and P∩O (c, d), in 1:1 M ratio to generate six-coordinate complexes [RuCl2(CO)2(?2-P∩P)] (1a, 1b) and [RuCl2(CO)2(?2-P∩O)] (1c, 1d), where P∩P: Ph2P(CH2)nPPh2, n = 2(a), 3(b); P∩O: Ph2P(CH2)nP(O)Ph2, n = 2(c), 3(d). The complexes are characterized by elemental analyses, mass spectrometry, thermal studies, IR, and NMR spectroscopy. 1a1d are active in catalyzed transfer hydrogenation of acetophenone and its derivatives to corresponding alcohols with turnover frequency (TOF) of 75–290 h?1. The complexes exhibit higher yield of hydrogenation products than catalyzed by RuCl3 itself. Among 1a1d, the Ru(II) complexes of bidentate phosphine (1a, 1b) show higher efficiency than their monoxide analogs (1c, 1d). However, the recycling experiments with the catalysts for hydrogenation of 4-nitroacetophenone exhibit a different trend in which the catalytic activities of 1a, 1b, and 1d decrease considerably, while 1c shows similar activity during the second run.  相似文献   

9.
Ab initio molecular orbital and density functional theory were used to investigate energetic and structural properties of the various conformations of hexa-tertbutylbenzene (1), hexakis(trimethylsilyl)benzene (2), hexakis (trimethylgermyl)benzene (3), and hexakis(trimethylstannyl)benzene (4). HF/3-21G//HF/3-21G and B3LYP/3-21G//HF/3-21G results revealed that the Twist-Boat (TB) conformer of compound 1 is more stable than the 1-Chair (C), 1-Boat (B), and 1-Planar (P) conformers. B3LYP/3-21G//HF/3-21G results show that the 1- TB conformer is more stable than 1- C, 1- B, and 1- P conformers of about 1.13, 4.34, and 99.94 kcal mol?1 , respectively. Contrary to the stability order of compound 1 conformers, the C conformer of compounds 2–4 is more stable than TB, B, and P conformations, as calculated by B3LYP/3-21G//HF/3-21G and HF/3-21G//HF/3-21G levels of theory. The energy gap between the C and P conformers in compounds 1–4 is decreased in the following order: ΔE(4: C, P) < ΔE (3: C, P) < ΔE(2: C, P) < ΔE (1: C, P). This fact can be explained in terms of the increase of C aromatic -M (M═C, Si, Ge, and Sn) bond lengths and the decrease of steric (van der Waals) repulsions in the previously discussed compounds. For compounds 1–3, the calculations were also performed at the B3LYP/ 6-31G*//HF/3-21G level of theory. However, the comparison showed that the results at B3LYP/3-21G//HF/3-21G methods correlated well with those obtained at the B3LYP/6-31G*// HF/6-31G method. Further, NBO analysis revealed that in compounds 1–4, the resonance energy associated with the σM-C1 to σ*C2-C3 delocalization is 5.20, 9.68, 11.15, and 12.27 kcal mol?1, respectively. These resonance energy values could explain the easiness of the ring flipping processes of C, B, and TB conformers of compounds 4 to 1. Also, the NBO results showed that by an increase of the σM-C1 → σ *C2-C3 resonance energies in compounds 1–4, the σM-C1 bonding orbital occupancies decrease. This fact could fairly explain the increase of the Caryl-M bond length from compound 1 to 4. The NBO results are also in good agreement with the calculated energy barriers for the ring flipping of the chair conformations in compounds 1–4, as calculated by B3LYP and HF methods.  相似文献   

10.
Compounds 3a–k were obtained from the reactions of compounds 1a–k with homopiperazine (2) in CH 2 Cl 2 . Compounds 1a–b, 1d–f, and 1h–l gave compounds 5a–b, 5d–f, and 5h–l with 2-methylpiperazine (4) in dichloromethane. Compounds 7c and 9c were obtained from the reactions of compound 1c with 4-ethoxycarbonyl piperazine (6) and 4-piperidinol (8) in CH 2 Cl 2 . Compounds 1a and 1f gave compounds 11a and 11f with 4-methylpiperazine (10), and compound 13f was obtained from the reactions of compound 1f with 4-methylpiperidine (12) in CH 2 Cl 2 .  相似文献   

11.
The simple triarylmethanol hosts, 2 and 4, and their silicon analogues, 1 and 3, have been studied for comparison of the formation of crystalline inclusion compounds. Clathrate formation experiments showed that replacement of the carbinol C atoms in 2 and 4 by Si atoms to give 1 and 3 resulted in a distinct increase of the capability to form inclusion compounds with organic guests, in particular with alcohols. Moreover, the naphthyl derivatives are much more efficient than the phenyl species, irrespective of the carbinol and silanol features. In order to investigate and compare the guest recognition modes and packing relations of hosts 1–4 in their crystalline inclusion compounds, 11 selected co-crystals, namely 1·DMSO (2:1), 3·EtOH (1:1), i-PrOH (1:1), acetone (1:1), DMSO (1:1), THF (1:1), piperidine (1:1), acetone (1:1), DMSO (1:1), 1,4-dioxane (1:1) and benzene (1:1), were studied by X-ray diffraction from single crystals. The survey contains additional 11 crystal structures from the literature and provides a detailed discussion of isostructurality relationships.Supplementary Data relevant to this publication have been deposited with the Cambridge Crystallographic Data Centre as supplementary publications nos. CCDC 274780–274790.  相似文献   

12.
Summary AM1 semi-empirical SCF MO calculations are reported for important conformations of oxocane (1) and 1,3-dioxocane (2). The boat-chair conformation of1 (BC-1) is found to be the most stable form, whereas the crown family conformation is calculated to be 3.7 kJ·mol–1 less stable. The boat-boat form of1 is 15.9 kJ·mol–1 less stable thanBC-1. The boat-chair conformation of2 (BC-1,3) is calculated to be the most stable form of 1,3-Dioxocane. The crown-family conformation and the boat-boat geometry of this compound are 4.2 and 8.3 kJ mol–1 less stable thanBC-1,3.
AM1-Rechnungen zu den Konformationen von Oxocane und 1,3-Dioxocan
Zusammenfassung Die wesentlichen Konformationen von Oxocan (1) und 1,3-Dioxocan (2) wurden mittels semiempirischer AM1-Rechnungen (SCF MO) untersucht. Die Wanne-Sessel-Konformation von1 (BC-1) ist am stabilsten; Konformationen aus der Klasse der Kronen sind um 3.7 kJ·mol–1 energiereicher. Die Wanne-Wanne-Konformation von1 ist um 15.9 kJ·mol–1 instabiler alsBC-1. Die Wanne-Sessel-Konformation von2 (BC-1,3) ist das stabilste Konformer von 1,3-Dioxocan. Kronenkonformationen und Wanne-Wanne-Geometrien sind um 4.2 und 8.3 kJ·mol–1 energiereicher alsBC-1,3.
  相似文献   

13.
During our continual searching programme for novel bioactive metabolites from Sarcophyton trocheliophorum, collected from Red Sea, we describe herein the isolation and structural elucidation of further two new pyrane-based cembranoid diterpenes: 9-hydroxy-7,8-dehydro-sarcotrocheliol (1) and 8,9-expoy-sarcotrocheliol acetate (2), along with the well-known sarcotrocheliol acetate (3), (+)-sarcophine (4), (+)-sarcophytoxide (5) and (-)-sarcophytoxide (6). The chemical structures of compounds 1 and 2 were determined on the basis of 1D and 2D NMR (1H, 13C, 1H–1H COSY, HMQC, HMBC and NOE), mass spectra (ESI and HR-ESIMS) and by comparison with related structures. The antimicrobial activities of the reported compounds 16 were investigated. According to the molecular docking study of compounds 16 using 3D structure of α,β tubulin in complex with taxol (PDB code 1JFF) and epothilone A (PDB code 1TVK), sarcophine (4) displayed the highest affinity towards both crystal structures, followed by 5 and 6, meanwhile pyrane-based cembranoid diterpenes (1–3) showed less affinity.  相似文献   

14.
ABSTRACT

Two acyclic CB[n]-type hosts (1 and 2) which possess four 2° or 3° amide arms are reported. Host 2 has four 3° amide arms that exist as a mixture of E- and Z-isomers. 1H NMR was used to qualitatively investigate the binding properties of 1 and 2 which indicates they retain the essential binding features of macrocyclic CB[n] hosts. We measured the Ka values of 1 and 2 toward guests 614 by ITC. Neutral hosts 1 and 2 bind less tightly than tetraanionic hosts M1, ACB1, and ACB2. We attribute the lower Ka values to the absence of secondary ion-ion electrostatic interactions for host?guest complexes of 1 and 2. The secondary amide functionality on 1 decreases affinity by the formation of intramolecular NH???O=C H-bonds. Tertiary amide host 2 binds even more weakly than 1 due to backfolding of the amide N-CH3-groups of 2 into its own cavity.  相似文献   

15.
Present paper reports on tensiometric studies of tetramethylsulfonatoresorcinarenes 1 and 2 with nonionogenic guests 3 and 4, pyrimidin derivative and O,O-dymethyl-1,1-dimethyl-3-oxobutylphosphonate, respectively. Association of resorcinarenes with these guests leads to dramatic change of adsorption characteristics of their solutions. CCMs1 of associates (1&3, 1&4, 2&3, and 2&4) are lower and the estimated surface activity, as well as the height of adsorption layers are higher than for individual substances. Aggregation of compounds 14 and association of 1 with 3 and 4 in solution were confirmed by 1H NMR spectra and studied by diffusion NMR with impulse magnetic field gradient.  相似文献   

16.
Reaction of 1,3-bis(imino)benzenes with a stoichiometric amount of LiAlH4 in THF yields iminoaminobenzenes L1 and L2. Further reaction of iminoaminobenzenes L1 and L2 with an equivalent of AlR3 in toluene affords macrocyclic binuclear aluminum complexes 1a, 1b, and 2a. These macrocyclic aluminum complexes were characterized by 1H NMR, 13C NMR, and IR spectroscopy. The molecular structures of 1a, 1b, and 2a were further confirmed by X-ray crystallography. X-ray diffraction analysis revealed that 1a, 1b, and 2a adopted a distorted tetrahedral geometry around aluminum. These complexes have efficient activities toward ring-opening polymerization of ε-caprolactone in the presence of benzyl alcohol.  相似文献   

17.
Two unsymmetric bis-aroyl-hydrazines, N′-(2-hydroxybenzoyl)isonicotinohydrazide (L1) and N′-(2-hydroxybenzoyl)nicotinohydrazide (L2), were synthesized through reactions of salicyl hydrazide with isonicotinoyl chloride and nicotinoyl chloride, respectively. Reactions of metal salts with L1 or L2 gave three new complexes, [Cd(L1)2(SCN)2] n (1), [Zn(L1)2Cl2]?·?H2O (2), and [Zn(L2)2Cl2] (3). Complex 1 features a 1-D double-chain structure built by SCN bridging six-coordinate CdII centers while 2 and 3 are mononuclear ZnII complexes. In 13, isomeric ligands L1 and L2 coordinate with metal ions in a terminal coordination mode. Ligands L1 and L2 through O–H···N and N–H···O hydrogen-bonding interactions in 13 are crucial for the structure extension into 3-D supramolecular structures of 1 and 2, or 2-D sheet of 3. Complexes 13 emit interesting blue-green luminescence. Thermal behaviors of 13 as well as the specific rotation of 2 were also investigated.  相似文献   

18.
Mono(thio)substituted 1a–c gave compounds 3a–c and 5a with o-toluidin (2) and m-toluidin (4) in ether. Compounds 9a–c and 11a, b were obtained from the reaction of compounds 1a–c with p-fluorophenylamine (8) and p-fluorobenzylamine (10). Compounds 7a and 15c were obtained from the reaction of 1a and 1c with p-phenylendiamine (6) and o-phenylendiamine (14). Compound 13c was synthesized from the reaction of compound 1c with benzidine (2).  相似文献   

19.
Four derivatives of acridine and acridinium compounds (L1, L2, L1H and L2H) comprised thiourea-binding sites were synthesised. The binding abilities of receptors L1, L2, L1H and L2H towards amino acids (l-Trp, l-Phe, l-Leu, l-Ala and l-Gly) were studied by 1H NMR spectroscopy, UV–vis and fluorescence spectrophotometry. Hydrogen bonding interactions between thiourea-binding site of the ligand and the carboxylate groups in zwitterionic amino acids were found to be the main interactions driving complexation to take place. The stoichiometry of 1:1 ligand to amino acid was observed in all cases. Neutral ligands L1 and L2 showed weak binding towards all studied amino acids. The cyclic ligand L1 showed better binding ability towards tryptophan (Trp) than the acyclic ligand L2 did (K for Trp is 307 and 266 M? 1 for L1 and L2, respectively). Interestingly, binding abilities of the protonated ligands, L1H and L2H, towards studied amino acids, especially Trp (K for Trp is 3157 and 2873 M? 1 for L1H and L2H, respectively), were increased due to R–COO…H…N+–acridinium interactions. Calculated structures of L1H·Trp and L2H·Trp showed that the polyglycol moiety in L1H provided a hydrophobic cavity for binding Trp resulting in a stronger binding affinity of L1H over L2H.  相似文献   

20.

Alkytris(2-pyridyl)phosphonium salts [(2-Py) 3 PR]X 1 [1a, R = Et, X = Br; 1b, R = Pr, X = Br; 1c, R = Bu, X = Br; 1d, R = CH2Ph, X = Br; 1e, R = CH 2 Ph, X = Cl] were synthesised from (2-Py) 3 P and an excess of RCl. 1c and 1e were found to rapidly decompose in hot acetone to 2,2′-bipyridinium(+1) bromide 2 and (2-Py)P(O)(CH 2 Ph)C(OH)Me 2 3, respectively. A reaction mechanism for both products is proposed. All compounds were fully characterized, including X-ray crystallography for 1a and 3 with 1a being the first representative of this class of compounds characterized by this technique.  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