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 15 毫秒
1.
The methylation of dihydropyrimidinylidenecyanoacetic esters by various methylating reagents leads to better yields when anions of ylidene derivatives of pyrimidine are used. The structure of the products of N- and C-methylation formed and their ratio were established with the aid of the IR and UV spectra and by 1H and 13C NMR spectroscopy. The product of C-methylation contains a methyl group at the C(7) atom of the side chain, while the product of N-methylation contains a methyl group at the nitrogen atom of the pyrimidine ring. For 6-ylidene derivatives, N-methylation occurs at the heterocyclic atom distant from the tautomerizing chain. For 2-and 6-ylidene derivatives of pyrimidine, the ratio of the products of N- and C-methylation can be changed by varying the alkylating reagent.For communication 1, see [1].Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 11, pp. 1537–1543, November, 1984.  相似文献   

2.
The existence of a pyrimidinyl-pyrimidinylidene tautomeric equilibrium in solutions of 2-pyrimidinylcyanoacetic acid esters in CDCl3 was observed. Unsymmetrically substituted compounds form two types of ylidene tautomers that differ with respect to the position of the NH proton, the ratio between which is controlled by the substituents in the 4 (6) position. The introduction of both donor and acceptor substituents into the 5 position of the pyrimidine ring increases the amount of the pyrimidine form. The same thing occurs when the polarity of the solvent is decreased. The addition of DMSO or DMF to CDCl3 leads to conversion of the intrachelate ylidene tautomers to unchelated tautomers. Protonation (CF3COOH) shifts the equilibrium to favor the ylidene tautomer that has higher basicity.See [1] for communication 5.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 6, pp. 827–831, June, 1984.  相似文献   

3.
Conclusions A preparative method for the synthesis of binuclear hexacarbonyldiiron complexes from aromatic azines and Fe3(CO)12in cyclohexane solutions in the presence of carbonyls of group VIb as catalysts has been proposed.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 2, pp. 418–422, February, 1982.  相似文献   

4.
Novel palladium‐1,3‐dialkylperhydrobenzimidazolin‐2‐ylidene (2a–c) and palladium‐1,3‐dialkylimidazolin‐2‐ylidene complexes (4a,b) have been prepared and characterized by C, H, N analysis, 1H‐NMR and 13C‐NMR. Styrene or phenylboronic acid reacts with aryl halide derivatives in the presence of catalytic amounts of the new palladium‐carbene complexes, PdCl2(1,3‐dialkylperhydrobenzimidazolin‐2‐ylidene) or PdCl2(1,3‐dialkylimidazolin‐2‐ylidene) to give the corresponding C? C coupling products in good yields. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

5.
A new series of potentially biological active derivatives, namely alkyl‐2‐((4‐oxo‐2‐(phenylimino)‐3‐(β‐d ‐pyranosyl‐2‐ylamino)thiazolidine‐5‐ylidene)acetate ( 5a–f ), 4‐(4‐bromophenyl)thiazol‐2(3H)‐ylidene)hydrazinyl)‐β‐d ‐pyranosyl ( 4a–c ), and 5‐(4‐bromophenyl)‐2‐(phenylimino)‐3‐(β‐d ‐pyranosyl‐2‐ylamino)thiazolidine‐4‐one ( 6 ) were synthesized via a reaction of the sugar thiosemicarbazone derivatives with 2,4′‐dibromoacetophenone, dialkylacetylenedicarboxylate, and ethylbromoacetate, respectively. The structures of the synthesized compounds were established by spectroscopic methods (FT‐IR, 1H NMR, 13C NMR, and 2D NMR) and elemental analyses. Furthermore, the effect of various solvents at reflux and also ambient temperature on the reactions of the sugar thiosemicarbazone with 2,4′dibromoacetophenone, diethyl acetylenedicarboxylate, and dimethyl acetylenedicarboxylate was investigated. © 2013 Wiley Periodicals, Inc. Heteroatom Chem 24:200–207, 2013; View this article online at wileyonlinelibrary.com . DOI 10.1002/hc.21083  相似文献   

6.
Solvation and association interactions in solutions of LiBF4/DMCC (DMCC for N,N-dimethylcarbamoyl chloride) and LiBF4/DMCC–DME (DME for 1,2-dimethoxyethane) have been studied as a function of concentration of lithium tetrafluoroborate by infrared and Raman spectroscopy. Strong interactions between Li+ and solvent molecules or BF4 anions are observed. The apparent solvation numbers of Li+ in LiBF4/DMCC solutions were deduced. Band-fitting to the B–F stretching band of BF4 anion permits detailed assess of the ion pairing. Based on the calculations of density function theory, optimal structures of Li+(DMCC)n (n = 1–3) were suggested. It is found that the lithium ion was preferentially solvated by DME in DMCC–DME binary solvents. This finding is supported by quantum chemistry calculations.  相似文献   

7.
The relative sound speed of mixtures of aqueous solutions of NaCl–MgSO4 and MgCl2–Na2SO4 at I=0.1 and 0.5m have been determined at 5, 15, and 25°C and pressures to 1000 bars. The resulting sound speeds, adiabatic and apparent molal compressibilities have been compared to results estimated from binary solutions using an additivity principle — Young's rule. The estimated sound speeds agree with the measured values for the NaCl–MgSO4 system to ±0.15 m-sec–1 and for the Na2SO4–MgCL2 system to ±0.20 m-sec–1. The deviations increase with increasing ionic strength (±0.08 m-sec–1 at I=0.1 and ±0.25 m-sec–1 at I=0.5 m).The sound speed of seawater have also been estimated from 0 to 40°C, 0.1 to 0.7 ionic strength and 0 to 1000 bars. The estimates were found to be in good agreement (±0.4 m-sec–1) with the measured values.These results indicate that reasonable estimates of the adiabatic PVT properties of dilute mixtures of electrolyte solutions can be made using the additivity principle, without excess mixing terms.  相似文献   

8.
The structural units in diphenylsilanediol/titanium-isopropoxide solutions with molar ratio Si:Ti between 1:0.1 and 1:5 were examined by means of 29Si and 17O NMR. The main component in solutions with molar ratio Si/Ti=1:0.1 is the chain-like octaphenyltetrasiloxanediol. With increasing Ti-isopropoxide content (1:0.25–1:05) Si–O–Ti units of the spirocyclic titanosiloxane Ti[O5Si4(C6H5)8]2 prevail in the solutions accompanied by the chain-like tetrasiloxanediol. The 29Si NMR spectra of 1:1 solutions indicate a lot of different Si containing building units with chemical shifts mainly between-40 and-46 ppm. The signals with a chemical shift between-40 and-46 ppm are probably caused by Si atoms which are connected via oxygen bridges directly (Si–O–Ti) or indirectly (Si–O–Si–O–Ti) with titanium. Contrary to the 1:1 solutions only one or two different species with Si–O–Ti units are present in high Ti-alkoxide containing solutions (1:5). 29Si and 17O NMR results reveal a quick hydrolysis of the Ti–O–Si bonds to titanium-oxo-hydroxo-polymers and phenylsiloxanediols or their isopropyl esters after the addition of water to the solutions. This separation into species only containing either Ti–O–Ti or Si–O–Si bonds can entail a decreased homogeneity of the reaction products on a molecular level.  相似文献   

9.
The densities of dilute aqueous solutions of [CoL3]X3 [L=1,2-diaminoethane(en), 1,2-diaminopropane(pn), 1,3-diaminopropane(tn) X=Cl, Br and (ClO4)] have been measured at 25°C from 0 to 5×10–2m. The apparent molar volumes were calculated and extrapolated to infinite dilution. Ion-solvent interactions were detected from the change of the ionic partial molar volumes with concentration. These interactions depend both on the properties of the ion (polarization charge density at the surface, hydrophobic groups, etc.) and the characteristics and structure of the solvent.  相似文献   

10.
Potentiometric titrations of uranyl(VI) solutions were conducted using a standard glasslcalomel electrode combination over the pH range 3 to 12 at 0.1 molkg–1 ionic strength with tetramethylammonium trifluoromethanesulfonate as the supporting electrolyte. The electrodes were calibrated directly on the hydrogen ion concentration scale during the initial stage of each titration. The species, UO 2 2+ , (UO2)2(OH) 2 2+ , (UO2)3(OH) 5 + , (UO2)3(OH) 7 , (UO2)3(OH) 8 2– , and (UO2)3(OH) 10 4– identified in an earlier Raman study were compatible with the analysis of the titration data. Based on this analysis and application of the extended Debye-Hückel treatment, the polynuclear species indicated above were assigned overall formation constants at 25°C and at infinite dilution of –5.51±0.04, –15.3±0.1, –27.77±0.09, –37.65±0.14, and –62.4±0.3, respectively. The results are discussed in reference to hydrolysis quotients reported in the literature for the first three species. Formation quotients for the last two species have not been reported previously.  相似文献   

11.
To design innovative and novel optical materials with high mobility, two kinds of disubstituted derivatives for meridianal isomer of tris(8-hydroxyquinolinato)aluminum (mer-Alq3) with push–pull (X–Y) substituents have been designed. The structures of tris(4-X-6-Y-8-hydroxyquinolinato)aluminum (type 1) and tris(4-Y-6-X-8-hydroxyquinolinato)aluminum (type 2) (where X = –CH3/–NH2 and Y = –CN/–Cl) in the ground (S0) and first excited (S1) states have been optimized at the B3LYP/6-31G* and CIS/6-31G* level of theory, respectively. All the designed derivatives of type 1 show blue shift while most of the type 2 derivatives show red shift as compared to the mer-Alq3. The emitting color could be tuned from blue to red. We have explained the distribution of HOMOs and LUMOs on different individual ligands. The reorganization energies of tris(4-methyl-6-chloro-8-hydroxyquinolinato)aluminum (1), tris(4-methyl-6-cyano-8-hydroxyquinolinato) aluminum (2), tris(4-chloro-6-methyl-8-hydroxyquinolinato)aluminum (5) and tris(4-cyano-6-methyl-8-hydroxyquinolinato)aluminum (6) are comparable with mer-Alq3. Thus these derivatives might be good candidates for emitting materials possessing comparable charge carrier mobility as mer-Alq3.  相似文献   

12.
The reviews and monographs on magnetochemistry of boron compounds are discussed. The structural peculiarities of borane derivatives relevant to magnetochemical calculations of diamagnetic contributions are are considered. Experimental measurements of diamagnetic susceptibility for deltahedral B10H 10 2– and B12H 12 2– cluster closo-anions were used to derive the diamagnetic atomic increments of the B atoms (B) with coordination numbers 5 and 6. The atomic increments B thus obtained were used to calculate molecular diamagnetic susceptibility M of borane derivatives. Diamagnetic susceptibility M was measured and calculated for the series of borane derivatives BnH n 2– and B10H12L2 (L is a Lewis base)and cobalt(III) derivatives of ortho-carborane(12) [(B9C2H11)2Con(B8C2H10) n–1] n. Diamagnetic increments were obtained for |B10H12| fragments and (B9C2H11)2– and (B8C2H10)4– ligands. The increments can be employed for calculating M for novel compounds. The calculated values of M coincide with the experimental values within 2%–6%.Original Russian Text Copyright © 2004 by V. V. Volkov and V. N. IkorskiiTranslated from Zhurnal Strukturnoi Khimii, Vol. 45, No. 4, pp. 729–740, July–August 2004.This revised version was published online in April 2005 with a corrected cover date.  相似文献   

13.
Tetracethoxysilane (TEOS) and methyltriethoxysilane (MTEOS) have been co-hydrolyzed in methanolic solutions containing tetramethylammonium ions that only affect polymerization of silicate species (hydrolysis products of TEOS) to form the Si8O 20 8– cubic octameric silicate species. The effects of water content and TEOS-to-MTEOS molar ratio on the distribution of species formed in the solutions have been investigated with the trimethylsilylation technique and 29Si n.m.r. spectroscopy. Formation of Si8O 20 8– and the cubic octameric species consisting of both Si(O)4 and CH3Si(O)3 units, (CH3)nSi8O 20–n (8–n)– (n=1–5), is found in the solutions. The increase of water content in the solutions solely results in increasing yield of Si8O 20 8– in spite of the presence of hydrolysis products of MTEOS together with those of TEOS, suggesting that water in the solutions plays an important role in the formation of Si8O 20 8– with the aid of tetramethylammonium ions. The TEOS-to-MTEOS molar ratio varies the distribution that is kept under control by the water content, increasing yields of (CH3)nSi8O 20–n (8–n)– (n=1, 2). It is found that the water content and TEOS-to-MTEOS molar ratio determine the reaction conditions effective for the formation of CH3Si(O)3 unit-containing cubic octameric species.  相似文献   

14.
Two novel 1,3‐dithiole‐2‐ylidene derivatives with a push–pull structures, 3‐(4,5‐dicarbomethoxy‐1,3‐dithiol‐2‐ylidene)naphthopyranone 1 and 3‐(4,5‐dimethylthio‐1,3‐dithiol‐2‐ylidene)naphthopyranone 2 , have been synthesized and characterized by 1H NMR, IR, MS. The UV–vis spectra of 1 , 2 in CH2Cl2, the lowest‐energy absorption bands, are centered at 280, 316, and 430 nm for 1 and 284, 317, and 450 nm for 2 , respectively, which are caused by the HOMO → LUMO single electron promotion. Furthermore, the steady‐state fluorescence originating states of 1 , 2 from the excited charge‐transfer were observed. To estimate the position and energies of frontier orbitals for 1 , 2 , DFT calculations were performed using the Gaussian 03 program at the B3LYP/6‐31 G* level. The calculated vertical excitation energies are in good agreement with the experimental data. The high HOMO–LUMO gaps of 1 (3.08 eV) and 2 (3.00 eV) indicate high kinetic stability of the title compounds.  相似文献   

15.
The conformational preferences for 2,3-O-isopropylidene-α- -sorbopyranose derivatives 3–6 were determined by using 1H NMR data and empirical force field calculations. Proton NMR studies of 3–6 indicate that a twist-boat (or skew) conformation (3S0) prevails over possible chair forms for each compound. Force-field calculations (MM2, MNDO, AM1) on a model 2,3-O-isopropylidene-α- -sorbopyranose system (18) indicate that the 3S0 conformation is among the low-energy structures. X-Ray crystallographic analysis of α- -sorbopyranose sulfamate 3, a compound with potent anticonvulsant activity, demonstrates that the 3S0 skew conformation is manifested in the solid state, as well.  相似文献   

16.
Some new ligand exchange reactions of [Co(diph·H)2Cl(H2O)] and [Co(diph·H)2(SO3)(H2O)] complexes with N3 , S2O3 2– and with aromatic and heterocyclic amines were carried out. A series of derivatives of the types [Co(diph·H)2(SO3)X] n– (X=N3 , S2O3 2– oramine) and [Co(diph·H)2(S2O3)2]3– were described and characterized. Some structural problems are resolved and discussed on the basis of UV and IR spectral data.  相似文献   

17.
Raman and infrared line parameters of Zn(NO3)2-H2O systems ranging from dilute solutions (25°C) to ionic liquids of low water content (75°C) are reported. At 25°C the solutions contain a very low concentration of inner sphere [Zn(ONO2)(H2O)5]+, outer sphere [Zn(H2O)6]2+[NO3], Zn(H2O) 6 2+ , and NO 3 (aq.). In the ionic liquids the ion triplet also exists. Manifestations of a change from the octahedral coordination of zinc to tetrahedral coordination when the water content is very low include the appearance of a 285 cm–1 band from the zinc nitrate bond and a shift to higher frequencies of the band from zinc-water.  相似文献   

18.
Raman OD stretching spectra of alcoholic LiX (X = Cl, Br, I, ClO4 , NO3 , and CH3COO) solutions (alcohols = methanol and ethanol) were measured in the liquid state at room temperature and in the glassy state at liquid nitrogen temperature. The effects of the anions on the Raman OD stretching spectra in these alcoholic solutions are investigated and the structural changes of the solutions are discussed. It is shown that the structure-breaking effects of anions on the intrinsic alcoholic structure increase in the order: Cl < Br < I < ClO4 . From spectral changes, it seems that CH3COOLi exerts little effect on the liquid structure of the alcohol.  相似文献   

19.
A series of new piano‐stool iron(II) complexes comprising N‐heterocyclic carbene ligands [Fe(Cp)(CO)2(NHC)]I (NHC = 1,3‐disubstituted imidazolidin‐2‐ylidene) have been synthesized and analyzed by 1H NMR, 13C NMR, IR, elemental analysis and mass spectrometric techniques. These compounds were easily prepared from the reaction of disubstituted imidazolidin‐2‐ylidene with [FeI(Cp)(CO)2] in toluene at room temperature. These complexes were tested in the catalytic hydrosilylation reaction of aldehydes and ketones with phenylsilane in solvent‐free conditions. After a basic hydrolysis step, the corresponding alcohols were obtained in good yields. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
A series of binuclear copper(ii) complexes with acyldihydrazones of aliphatic dicarboxylic acids (from malonic to adipic) and fluorinated -diketones (trifluoro- and hexafluoroacetylacetone) of composition Cu2L·2Py (H4L is acyldihydrazone) were studied by ESR spectroscopy. The ESR spectra of solutions of complexes with trifluoroacetylacetone acyldihydrazones have an isotropic signal with a seven-line hyperfine structure from two equivalent copper nuclei (g = 2.112, a Cu = (39—40)·10–4 cm–1), which is indicative of weak exchange interactions between the paramagnetic centers due to spin density delocalization through a chain of the -bonds of the polymethylene bridge. On going to hexafluoroacetylacetone derivatives, the coupling is suppressed and the ESR spectra of solutions of such complexes show a signal with a four-line hyperfine structure (g = 2.121—2.131, a Cu = (55—63)·10–4 cm–1) typical of mononuclear copper complexes.  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