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.
Chloro-dimethylamino-phenyl-p-tolylthio-phosphonium chloride 2 , dimethylamino-diphenyl-p-tolylthio-phosphonium chloride 3 , bis(diethylamino)-dimethylamino-p-tolylthiophosphonium chloride 4 and tert-butyl-dimethylamino-phenyl-p-tolylthio-phosphonium chloride 5 were prepared by the reaction of N,N-dimethylamino-p-tolylsulfenamide 1 with PhPCl2, Ph2PCl, (Et2N)2PCl and tBu(Ph)PCl, respectively. The reaction of N,N′-dimethyl-N,N′-bis(trimethylsilyl)urea 9 and N-methyl-N′-phenyl-N,N′-bis(trimethylsilyl)urea 10 with phenylsulfenyl chloride 6 or p-nitrophenylsulfenyl chloride 8 furnished the N-arylthio-N,N′-diorgano-N′-(trimethylsilyl)-ureas 11 – 14 . The reaction of 11 – 14 and of the previously known compounds 15 and 16 with MePCl2, ClCH2PCl2, tBuPCl2 and PhPCl2 resulted in the formation of the 2-arylthio-2-chloro-1,2,3-triorgano-1,3,2λ5-diazaphosphetidin-4-ones 17 – 26 . 1,3-Dimethyl-2-(1,1,1,3,3,3-hexafluoro-2-propoxy)-2-phenyl-2-phenylthio-1,3,2λ5-diazaphosphetidin-4-one 29 and the 2-arylthio-1,3-dimethyl-2-(p-nitrophenoxy)-2-organyl-1,3,2λ5-diazaphosphetidin-4-ones 30 – 32 were obtained in the reactions of compounds 17, 24 and 27 with 1,1,1,3,3,3-hexafluoro-2-propanol or p-nitrophenol in the presence of triethylamine. The reaction of compound 21 with thiophenol in the presence of triethylamine resulted in a mixture of products, from which 1,3,4,5,7-pentamethyl-1,3,5,7-tetraaza-4λ5-phosphaspiro[3.3]heptan-2,6-dione 33 was isolated. The identity and structure of all the new compounds were established by 1H-, 13C- and 31P-NMR spectroscopy and by elemental analysis. A possible mechanism of reaction of sulfenamides with compounds of trivalent phosphorus is discussed. For the compounds 5a, 32 and 33 X-ray structure analyses were conducted. The cation of compound 5a involves four-coordinate phosphorus (essentially tetrahedral geometry) and is a rare example of a P–S single bond in such a system (P–S 207.37(9) pm). In 32 the geometry at phosphorus is distorted trigonal bipyramidal, with axial positions occupied by oxygen and nitrogen atoms. In the spirophosphorane 33 the geometry at phosphorus is intermediate between trigonal bipyramidal and square pyramidal, with essentially planar four-membered rings.  相似文献   

2.
The reactions of urea with methylolphenols under acidic conditions were investigated using 2- and 4-hydroxybenzyl alcohol and crude 2,4,6-trimethylophenol as model compounds. The reaction products were analyzed with 13C-NMR spectroscopy and GPC. From the reaction of urea with 4-hydroxybenzyl alcohol, the formations of 4-hydroxybenzylurea, N,N′-bis (4-hydroxybenzyl) urea, and tris(4-hydroxybenzyl) urea were confirmed and the formations of N,N-bis(4-hydroxybenzyl) urea and tetrakis (4-hydroxybenzyl) urea were suggested. From the reaction of urea and 2-hydroxybenzyl alcohol, 2-hydroxybenzylurea and N,N′-bis(2-hydroxybenzyl) urea were identified. Further, the alternative copolymer of urea and phenol could be synthesized by the reaction of urea with 2,4,6-trimethylophenol. It was also found that the cocondensation between p-methylol group and urea prevails against the self-condensation of the methylolphenol even at the low pH below 3.0, and that p-methylol group has the stronger reactivity to urea than o-methylol group. © 1992 John Wiley & Sons, Inc.  相似文献   

3.
The synthesis of a new series of 4‐aryl‐3‐chloro‐2‐oxo‐N‐[3‐(10H‐phenothiazin‐10‐yl)propyl]azetidine‐1‐carboxamides, 4a – 4m , is described. Phenothiazine on reaction with Cl(CH2)3Br at room temperature gave 10‐(3‐chloropropyl)‐10H‐phenothiazine ( 1 ), and the latter reacted with urea to yield 1‐[3‐(10H‐phenothiazin‐10‐yl)propyl]urea ( 2 ). Further reaction of 2 with several substituted aromatic aldehydes led to N‐(arylmethylidene)‐N′‐[3‐(phenothiazin‐10‐yl)propyl]ureas 3a – 3m , which, on treatment with ClCH2COCl in the presence of Et3N, furnished the desired racemic trans‐2‐oxoazetidin‐1‐carboxamide derivatives 4a – 4m . The structures of all new compounds were confirmed by IR, and 1H‐ and 13C‐NMR spectroscopy, FAB mass spectrometry, and chemical methods.  相似文献   

4.
Contributions to the Chemistry of Transition Metal Alkyl Compounds. 64 [1] Preparation and Crystal Structure of Bis[N,N′ -bis(trimethylsilyl)-benzamidinato]dimethyltitanium Tetramethyltitanium reacts with excess N,N′-bis(trimethylsilyl)-benzamidine ( 1 ) to give bis[N,N′-bis(trimethylsilyl)benzamidinato]dimethyl-titanium ( 2 ) via protolysis. This compound was isolated and characterized by X-ray crystal structure analysis. In the crystal, 2 adopts an octahedral coordination with the methyl groups in cis-positions. The monomethyl complexes [PhC(NSiMe3)2]2M(Me)Cl ( 3 : M = Ti, 4 : M = Zr) have been prepared by treatment of the dichloro precursors with one equivalent of methyllithium.  相似文献   

5.
A dinucleating spacer 1,4-bis(salicylidene)phenylenediamine (SALPHEN) derived from 1,4-phenylenediamine and salicylaldehyde has been synthesized and characterized. The ruthenium(II) sulfoxide derivative of 2,2′-bipyridine or 1,10-phenanthroline on reaction with this ligand resulted in the formation of eight dinuclear complexes, which were characterized by elemental analyses, conductivity measurements, magnetic susceptibility, FT-IR, fast atom bombardment-mass spectra, electronic spectroscopy, 1H-NMR, 13C{1H}-NMR, and 2D-NMR spectra (HETCOR). The prepared complexes have two different formulations, [{trans-RuCl2(so)(N–N′)}2(μ-SALPHEN)] and [{cis-RuCl2(so)(N–N′)}2(μ-SALPHEN)], where so?=?dimethyl sulfoxide (DMSO)/tetramethylene sulfoxide (TMSO), N–N′?=?2,2′-bipyridine/1,10-phenanthroline, and SALPHEN?=?1,4-bis(salicylidene)phenylenediamine. Two moles of ruthenium sulfoxide bipyridine precursor were coordinated to the bidentate SALPHEN through nitrogen. All the complexes possess antibacterial activity against Escherichia coli in comparison to Chloramphenicol.  相似文献   

6.
Abstract

A series of novel acyclic and cyclic diaryl sulphides was synthesized from 2,2′-dithiobenzoic acid. The various diaryl sulphides were characterized by use of spectral (IR, 1H and 13C NMR, ESI/MS) and elemental analyses. The antimicrobial activities of the compounds were evaluated in terms of their minimum inhibition concentration (MIC) against a panel of clinical isolates bacteria and were found to possess only moderate antimicrobial activities. N,N′-Bis(2-hydroxyphenyl)-2,2′-thiodibenzamide (13), exhibiting a hydroxy group at the phenyl ring, was the most active antimicrobial agent within the series, with MIC values of 0.05 mg mL–1 and 10 mg mL–1 against Staphylococcus aureus and Bacillus cereus, respectively. The antioxidant efficiency of the diaryl sulphides was determined by 2,2-diphenyl-1-picrylhydrazyl (DPPH) scavenging activity with 13 being the most active compound. The interaction of 2,2′-thiodibenzanilide, N,N′-bis(2-methylphenyl)-2,2′-thiodibenzamide, and N,N′-bis(2-hydroxyphenyl)-2,2′-thiobenzamide with guanine, glutamic acid, and urea were studied quantitatively with binding constants ranging from 1 × 103 M?1 to 2.7 × 104 M?1.  相似文献   

7.
Monomeric Dialkyl Metal Complexes of the R2M(NR′)2XR Type with M = Al, Ga, In, Tl; X = S, C and R, R′ = Alkyl and Silyl N,N′-Bis(trimethylsilyl)sulfurdiimide reacts with the trimethyl derivatives of aluminium, gallium, and indium within insertion. Hereby monomeric sulfinic acid imidamidates Me2M(NSiMe3)2SMe (Me = CH3) are formed. The lithium amidinates Li(NR′)2CMe (R′ = i-C3H7 and SiMe3) are formed likewise by insertion reactions with LiMe and the corresponding carbodiimides R′N?C?NR′ and were used in reactions with R2MCl (M = Al to Tl) to synthesize dialkyl metal amidinates R2M(NR′)2CMe. The NMR (1H and 13C) and the vibrational spectra (IR and Raman) are discussed and applied to describe the structure of these chelat complexes.  相似文献   

8.
A number of mono- and dioligosilanylated silocanes were prepared. Compounds included silocanes with 1-methyl-1-tris(trimethylsilyl)silyl, 1,1-bis[tris(trimethylsilyl)silyl], and 1,1-bis[tris(trimethylsilyl)germyl] substitution pattern as well as two examples where the silocane silicon atom is part of a cyclosilane or oxacyclosilane ring. The mono-tris(trimethylsilyl)silylated compound could be converted to the respective silocanylbis(trimethylsilyl)silanides by reaction with KOtBu and in similar reactions the cyclosilanes were transformed to oligosilane-1,3-diides. However, the reaction of the 1,1-bis[tris(trimethylsilyl)silylated] silocane with two equivalents of KOtBu leads to the replacement of one tris(trimethylsilyl)silyl unit with a tert-butoxy substituent followed by silanide formation via KOtBu attack at one of the SiMe3 units of remaining tris(trimethylsilyl)silyl group. For none of the silylated silocanes, signs of hypercoordinative interaction between the nitrogen and silicon silocane atoms were detected either in the solid state. by single crystal XRD analysis, nor in solution by 29Si-NMR spectroscopy. This was further confirmed by cyclic voltammetry and a DFT study, which demonstrated that the N-Si distance in silocanes is not only dependent on the energy of a potential N-Si interaction, but also on steric factors and through-space interactions of the neighboring groups at Si and N, imposing the orientation of the pz(N) orbital relative to the N-Si-X axis.  相似文献   

9.
The synthesis of three novel pyrazole-containing complexing acids, N,N,N′,N′-{2, 6-bis[3-(aminomethyl)pyrazol-1-yl]-4-methoxypyridine}tetrakis(acetic acid)( 1 ), N,N,N′,N′-{2, 6-bis[3-(aminomethyl)pyrazol-1-yl]pyrazine}-tetrakis(acetic acid) ( 2 ), and N,N,N′,N′-{6, 6′-bis[3-(aminomethyl)pyrazol-1-yl]-2, 2′-bipyridine}tetrakis(acetic acid) ( 3 ) is described. Ligands 1–3 formed stable complexes with EuIII, TbIII, SmIII, and DyIII in H2O whose relative luminescence yields, triplet-state energies, and emission decay lifetimes were measured. The number of H2O molecules in the first coordination sphere of the lanthanide ion were also determined. Comparison of data from the EuIII and TbIII complexes of 1–3 and those of the parent trisheterocycle N,N,N′,N′-{2, 6-bis[3-(aminomethyl)pyrazol-l-yl]pyridine}tetrakis(acetic acid) showed that the modification of the pyridine ring for pyrazine or 2, 2′-bipyridine strongly modify the luminescence properties of the complexes. MeO Substitution at C(4) of 1 maintain the excellent properties described for the parent compound and give an additional functional group that will serve for attaching the label to biomolecules in bioaffinity applications.  相似文献   

10.
The complex formation reaction between N,N′-bis(2-pyridylmethylidene)-1,2-diiminoethane (BPIE) di-Schiff base ligand as an electron donor and iron(III) chloride as an electron acceptor have been studied spectrophometrically in methanol at 28°C. The values of equilibrium constants, K and molar absorptivities, ε were obtained from the Benesi–Hildebrand, Scott and Foster–Hammick–Wardley equations. The results indicate the formation of 1?:?1 charge transfer complex. The absorption band energy of the complex, E CT, the ionization potential of the BPIE Schiff base ligand, I D, and the Gibbs energy changes of the above reaction, ΔG 0, were calculated. Finally, the kinetics of the complex formation reaction were studied and was found to be second-order in each reactant. The values of the rate constants of the forward and reverse reactions k 1 and k ?1 were determined.  相似文献   

11.
The reaction of (chloromethyl)dichlorophosphine 1 with N,N′-dimethyl-N,N′-bis(trimethylsilyl)urea 2 furnished the σ5P-σ3P-diphosphorus compound 3 . The reaction of 3 with hexafluoroacetone proceeded in an unusual fashion, with the rupture of the P? P bond, resulting in 4,4-bis(trifluoromethyl)-3-chloro-2-hexfluoroisopropoxy-2-oxo-1,2-oxaphosphetane 7 and the spirophosphorane 4-chloromethyl-1,3,5,7-tetramethyl-1,3,5,7-tetraaza-4σ5-phosphaspiro-[3,3]heptan-2,6-dione 8 . The reaction of 2-chloro-1,2-dimethyl-3-phenyl-2-phenylseleno-1,3,2σ5-diazaphosphetidin-4-one 9 with bis(2-chloroethyl)amine hydrochloride/triethylamine 10 also proceeded in an unexpected fashion, leading to the spirophosphorane 11 as the only identified product. Single-crystal X-ray structure analyses of compounds 8 and 11 were conducted. The coordination geometry at phosphorus in both compounds shows a large deviation from idealized forms. This distortion arises mainly from the presence of the four-membered rings.  相似文献   

12.
In the reaction of TiCl4 in benzene as solvent with the imidoyl chloride p‐Tolyl(Cl)C=NPh ( 1 ) the abstraction of the chloride substituent is observed, leading to the nitrilium salt [p‐Tolyl–C≡N–Ph]+[Ti2Cl9] ( 2 ) in quantitative yield. The highly electrophilic salt 2 is characterized by IR‐ and NMR spectroscopy. The observed band for the C≡N stretching mode of 2 clearly indicates the formation of a nitrilium ion. Especially a characteristic line broadening of the 13C{1H}‐NMR signals related to carbon atoms next to the nitrogen is observed. By 15N,1H‐HMBC NMR experiments it is shown that the nitrogen signal of 2 is significantly shifted to high‐field in relation to nitriles and imines. The molecular structure of 2 was confirmed by single‐crystal X‐ray diffraction. The C≡N bond length and the linearity of the C–C≡N–C unit in 2 confirm the triple bond character of this bond.  相似文献   

13.
The hitherto unreported compounds of general structure 3,3′-(alkanediyl)bis-(2,2,2-triaryl-1-oxa-2-stiba-3-azabenzo[d]cyclohex-5-ene) have been synthesized in 48-56% yields by the cyclization of the tetrasodium salt of N,N′-bis(2-hydroxybenzyl)-1,2-diaminoethane(II) or of N,N′-bis(2-hydroxybenzyl)-1,3-diaminopropane(II*) with R3SbBr2 (R = phenyl, p-tolyl, or mesityl). The tetrasodium salts were prepared by the reactions of the corresponding amines with sodium hydride. The amines (II and II*), in turn, were obtained by the sodium borohydride reduction of N,N′-bis(salicylidene)-1,2-diaminoethane and N,N′-bis-(salicylidene)-1,3,-diaminopropane, respectively. The heterocyclic compounds are air stable and moisture insensitive. These compounds have been characterized by elemental analyses, molecular weight determinations, and by IR, far IR, 1H, and 13C NMR spectral studies. © 1996 John Wiley & Sons, Inc.  相似文献   

14.
The five benzene rings-containing (hereafter referred to as “five-ring”) dicarboxylic acids α,α′-bis[4-(4-carboxyphenoxy)phenyl]-1,4-diisopropylbenzene (p- III ) and α,α′-bis[4-(4-carboxyphenoxy)phenyl]-1,3-diisopropylbenzene (m- III ) were prepared by the fluoro-displacement of α,α′-bis(4-hydroxyphenyl)-1,4-diisopropylbenzene and α,α′-bis(4-hydroxyphenyl)-1,3-diisopropylbenzene with p-fluorobenzonitrile, and subsequent alkaline hydrolysis of the intermediate dinitriles. A number of high-molecular-weight polyamides based on these two “five-ring” dicarboxylic acids (p- III and m- III ) and various aromatic diamines were directly synthesized in N-methyl-2-pyrrolidone (NMP) containing lithium chloride (LiCl) or calcium chloride (CaCl2) using triphenyl phosphite and pyridine as condensing agents. These polyamides were obtained with inherent viscosities above 0.51 and up to 0.91 dL/g. The weight-average molecular weight were in the range of 51,000–211,000. Most of these polyamides were amorphous and readily soluble in polar solvents such as NMP, N,N-dimethylacetamide (DMAc), N,N-dimethylformamide (DMF), and dimethyl sulfoxide (DMSO), and afforded tough, flexible, and transparent films by solution-casting. The films had tensile strength of 50–83 MPa, elongation to break of 4–8%, and tensile modulus of 1.3–2.0 GPa. Most polyamides showed distinct glass transitions on the differential scanning calorimetry (DSC) curves ranging from 147 to 177°C. In nitrogen or air, all the polymers showed no significant weight loss up to 490°C, as indicated by thermogravimetric analysis (TG). © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Reactions of Fluorophosphoranes with the N,O-Bis(trimethylsilyl) Derivative of o-Aminophenol The reaction of the N,O-bis(trimethylsilyl) derivative of o-aminophenol, 5 , with the tetrafluorophosphoranes, RPF4, 2a–2d , (R = F, Me, Ph, and 1-adamantyl) in a 1:1 molar ratio led to monocyclic-1,3,2λ5-4,5-benzoxazaphospholes, C6H4(O)(NH)PF2R, 6a–6d . 19F n.m.r. spectroscopic studies suggest a trigonal-bipyramidal structure with the C6H4(O)(NH) grouping attached to one axial and one equatorial position at five-coordinate phosphorus for these compounds. The spirophosphoranes, [C6H4(O)(NH)]2PR, 8a – 8d (R = F, Me, Ph, 1-adamantyl) were obtained from the reaction of the appropriate tetrafluorophosphorane, RPF4, 2a – 2d with 5 in a 1:2 molar ratio. The compounds 8a – 8d also result from a spontaneous scrambling reaction of 6a – 6d , with the corresponding tetrafluorophosphoranes, RPF4 ( 2a – 2d ) as the other product. Reaction of the difluorophosphorane, Bu3nPF2 with 5 and with N,N′-dimethyl-N,N′-bis(trimethylsilyl)urea furnished the cyclic, fluorine-free phosphoranes, 9 and 10 , respectively. The phosphonium bromide, Bu3nPFBr, reacted with 5 in a 1:1 and a 2:1 molar ratio to produce the ionic compounds, [C6H4(OSiMe3)(NHPBu3n)]+Br?, 11 , and [C6H4(OPBu3n)HNPBu3n]2+ 2 Br?, 12 , respectively.  相似文献   

16.
Hexamethyldisilazane 1 reacts with borane in tetrahydrofuran (THF? BH3, 2 ) first by formation of an adduct (Me3Si)2NH? BH3 ( 3 ), and then either to the N,N-bis-(trimethylsilyl)-μ-aminodiborane 5 or to the mixture of 5 and N-trimethylsilyl-μ-aminodiborane(6) 6 , depending on the reaction conditions. The compounds 5 and 6 can be quantitatively converted to the N,N′,N″-tris(trimethylsilyl)borazine 4 . Three intermediates can be identified, namely N,N-bis(trimethylsilyl)borane 7 , N,N-bis(trimethylsilyl)amino(N′-trimethylsilylamino)borane 8 and N-trimethylsilylaminoborane-trimer. All products and intermediates were characterized by multinuclear NMR spectroscopy, and coupling constant 1J(29Si, 15N) were measured from 29Si NMR spectra by using the Hahn-echo-extended (HEED) INEPT pulse sequence.  相似文献   

17.
The novel (E,E)-dioxime 5,6:13,14-dibenzo-9,10-benzo(15-crown-5)-2,3-bis(hydroxyimino)-7,12-dioxo-1,4,8,11-tetraazacyclotetradecane (H2L) has been synthesized by the reaction of 4′,5′-diaminobenzo(15-crown-5) with N,N′-bis(2-carbomethoxyphenyl)diaminoglyoxime (1). Only mononuclear CoIII and RuII complexes with a metal/ligand ratio of 1:2 have been isolated. The cobalt(III) complex bridged with BF2+ is achieved with H-bonded cobalt(III) complex and borontrifluoride ethyl ether complex. The reaction of BF2 bridged cobalt(III) complex with bis(benzonitril)palladium(II) chloride gives a trinuclear complex. The structures of dioxime and its complexes are proposed according to elemental analyses, 1H and 13C-NMR, IR and mass spectral data.  相似文献   

18.
The new α‐aminophosphonic acids are synthesized, reacting (9H‐fluoren‐9‐yl)urea with formaldehyde and phosphorus trichloride. (9H‐Fluoren‐9‐yl)urea was prepared from spiro(fluoren‐9,4′‐imidazolidine)‐2′,5′‐dione by alkaline hydrolysis with Ba(OH)2. The structure of the title compounds was proved by means of IR, 1H, 13C, and 31P NMR spectroscopy. © 2008 Wiley Periodicals, Inc. Heteroatom Chem 19:719–722, 2008; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20500  相似文献   

19.
Mononuclear palladium‐hydroxo complexes of the type [Pd(N–N)(C6F5)(OH)][(N–N) = 2,2′‐bipyridine (bipy), 4,4′‐dimethyl‐2,2′‐bipyridine (Me2bipy), 1,10‐phenantroline (phen) or N,N,N′,N′‐tetramethylethylenediamine (tmeda) react with phenols ArOH in tetrahydrofuran giving the corresponding aryloxo complexes [Pd(N–N)(C6F5)(OAr)]. Elemental analyses and spectroscopic (IR, 1H and 19F) methods have been used to characterize the new complexes. The X‐ray crystal structure of [Pd(tmeda)(C6F5)(OC6H4NO2p)] has been determined. In the crystal packing the planes defined by two C6H4 rings show a parallel orientation. There are also intermolecular C–H···F and C‐H···O hydrogen bonds.  相似文献   

20.
Reaction of either 9,10‐phenanthrenedione (phenanthrenequinone) or diphenylethanedione (benzil) with two equivalents of Li[N(SiMe3)2], followed by quenching of the reaction with excess ClSiMe3, produces the corresponding N,N′‐bis(trimethylsilyl)‐α‐diimines in high yields (85–95%). Subsequent dehalosilylation/ring‐closure reactions with SbCl3 and BiCl3 produce, in 90–95% yields, the first examples of 1,3,2‐diazaheterole ring compounds containing antimony or bismuth. These 2‐chloro‐1,3,2‐diazaheteroles can be further functionalized at the pnictogen by reaction with, for example, Li[N(SiMe3)2], to produce the corresponding 2‐bis(trimethylsilyl)amido‐1,3,2‐diazaheteroles. All of these new main group element–containing heterocycles have been characterized through 1H and 13C NMR, elemental analysis, and two of the diazastiboles have been structurally characterized by single‐crystal X‐ray analysis, confirming the ring structures. Both of these diazastiboles exist as associated dimers in the solid state; half of the dimer represents the asymmetric unit. © 1999 John Wiley & Sons, Inc. Heteroatom Chem 10: 423–429, 1999  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