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 general synthetic method of pyrazines1 has been established. The method for synthesizing symmetrically substituted pyrazine derivatives, however, has not been well-studied.2 Especially for synthesis of the title pyrazines, it is only described in a few words by Adkins3 and co-workers that ethyl 2-aminoacetoacetate employed in process to preparing of the threonine synthesis is spontaneously autoxidized to give 2,5-dimethyl-3,6-dicarbetoxy pyrazine as a byproduct. On the basis of this finding of Adkins, we have established a convenient method for the synthesis of 2,5-di-methylpyrazine-3,6-dicarboxylic acid derivatives (D1–5) using the corresponding acetoacetic acid derivatives (A1–4) as the starting materials (Scheme).  相似文献   

2.
This article presents a combined experimental and computational study of Ru(II) complex containing 2,5-pyridinedicarboxylic acid ligand. The novel complex [Ru(py-2,5-COOH)2(PPh3)2]·3H2O has been obtained in the reaction of [RuCl2(PPh3)3] with 2,5-pyridinedicarboxylic acid in methanol and has been studied by IR, 1H, 31P NMR, UV–Vis spectroscopy, and X-ray crystallography. The electronic structure of [Ru(py-2,5-COOH)2(PPh3)2] has been calculated with the density functional theory (DFT) method. The spin-allowed electronic transitions of the complex have been calculated with the time-dependent DFT method, and the UV–Vis spectrum has been discussed on this basis and rationalized by determination of ligand field splitting (10Dq) and Racah’s parameters from the experimental spectrum. The luminescence property of the complex has been examined.  相似文献   

3.
5-Hydroxymethylfurfural (5-HMF) has been described as one of the 12 key platform molecules derived from biomass by the US Department of Energy, and its hydrogenation reaction produces versatile liquid biofuels such as 2,5-dimethylfuran (2,5-DMF). Catalytic hydrogenation from 5-HMF to 2,5-DMF was thoroughly studied on the metal nickel catalysts supported on Al2O3-TiO2-ZrO2 (Ni/ATZ) mixed oxides using isopropanol and formic acid (FA) as hydrogen donors to find the best conditions of the reaction and hydrogen donor. The influence of metal content (wt%), Ni particle size (nm), Nickel Ni0, Ni0/NiO and NiO species, metal active sites and acid-based sites on the catalyst surface, and the effect of the hydrogen donor (isopropanol and formic acid) were systematically studied. The structural characteristics of the materials were studied using different physicochemical methods, including N2 physisorption, XRD, Raman, DRS UV-Vis, FT-IR, SEM, FT-IR Pyad, H2-TPD, CO2-TPD, H2-TPR, TEM and XPS. Second-generation 2,5-DMF biofuel and 5-HMF conversion by-products were analyzed and elucidated using 1H NMR. It was found that the Ni0NiO/ATZ3WI catalyst synthesized by the impregnation method (WI) generated a good synergistic effect between the species, showing the best catalytic hydrogenation of 5-HMF to 2,5-DMF using formic acid as a hydrogen donor for 24 h of reaction and temperature of 210 °C with 20 bar pressure of Argon (Ar).  相似文献   

4.
The mass spectra of 2,5-diaryl-1-(arylimino)-1λ4,2,5-thiadiazolidin-3,4-diones have been examined. Some fragmentations are explained by rearrangement of the title compounds to 1,3-(diarylimino)-5-aryl-1λ4,2,5-thioxazolidin-4-ones and 1-oxo-2,5-diaryl-3-(arylimino)-1λ4,2,5-thiadiazolidin-4-ones.  相似文献   

5.
Four 2,5-biphenylene based difunctional condensation monomers, such as 2,5-diphenyldicar-boxylic acid or phenylterephthalic acid (PTA), 2,5-bis(carbonylimino-4-benzoic acid)biphenyl (2,5-BCIBABP), 2,5-diaminobiphenyl hydrochloride (2,5-DABP.HCl) and 2,5-bis(iminocarbonyl-4-benzoic acid)biphenyl (2,5-BICBABP), have been synthesized and characterized. These monomers were polymerized in combination with terephthalic acid (TA) and p-phenylenediamine (PPD) via the phosphorylation reaction to prepare a series of phenyl-substituted random copolyamides having all amide groups attached to para-positions of the benzene rings. All the copolyamides have been characterized by solubility, solution viscosity, and by differential scanning calorimetry (DSC). Some of these copolyamides have unusual solubility in organic solvents such as N,N-dimethylacetamide (DMAc) and N-methyl-2-pyrrolidinone (NMP) containing dissolved lithium chloride. A few copolyamides were tested for lyotropic behavior and found to form anisotropic solutions at critical concentrations in organic solvents. A randomly distributed unsymmetrical phenyl substituent on the benzene ring of para-oriented wholly aromatic polyamides dramatically changes the solubility and melting point. The phenyl substituent on a terephthalic acid unit is more efficient in decreasing melting point and increasing solubility in organic solvents of aromatic polyamides than the one on a p-phenylenediamine unit. However, the former also introduces a more flexible link in the extended polyamide chain.  相似文献   

6.
Six new organotin carboxylates based on 1,3-benzenedicarboxylic acid and 1,4-benzenedicarboxylic acid derivatives, namely (Ph3Sn)2(2,5-L1)(C2H5OH)2 (1) (2,5-H2L1 = 2,5-dibenzoylterephthalic acid), (Ph3Sn)2(2,5-L2)(C2H5OH)2 (2) (2,5-H2L2 = 2,5-bis(4-methylbenzoyl)terephthalic acid), (Ph3Sn)2(2,5-L3)(C2H5OH)2 (3) (2,5-H2L3 = 2,5-bis(4-ethylbenzoyl)terephthalic acid), [(n-Bu2Sn)4(4,6-L1)O2(OH)(OC2H5)]2·2(C2H5OH) (4) (4,6- H2L1 = 4,6-dibenzoylisophthalic acid), [(n-Bu2Sn)4(4,6-L1)O2(OH)(OC4H9)]2·2(C4H9OH) (5) and [(n-Bu2Sn)4(4,6-L2)O2(OH)(OC2H5)]2·2(C2H5OH) (6) (4,6-H2L2 = 4,6-bis(4-methylbenzoyl)isophthalic acid), have been synthesized. All the organotin carboxylates have been characterized by elemental analysis, IR, 1H and 13C NMR spectroscopy and X-ray crystallography diffraction analyses. The structural analysis reveals that complexes 1-3 show similar structures, containing binuclear triorganotin skeletons. The significant intermolecular O-H?O hydrogen bonds linked the complexes 1-3 to form a novel 2D network polymer with 38-member macrocycles. In complexes 4-6, two Sn4O4 ladders are connected by two 1,3-benzenedicarboxylic acid derivatives to yield ladder-like octanuclear architectures and form macrocycle with 24 atoms. In addition, the antitumor activities of complexes 1-6 have been studied.  相似文献   

7.
P,P′-(2,5-Dimethoxy-3,6-dimethyl-2,5-dioxo-2λ5,5λ5-[1,4,2,5]dioxadiphosphinane-2,5-diyl)-bis-phosphonic acid tetramethylester The title compound 1 is formed by reaction of the corresponding phosphonic acid 2 and orthoformicacidmethylester as a mixture of four stereoisomeres. The RRSS isomer was separated. It crystallizes in the triclinic space group P ?1 with a = 649.2 pm, b = 976.1 pm, c = 1 571.7 pm, α = 80.9°, β = 88.1°, γ = 78.6° and Z = 2. The 31P and 13C NMR spectra (4 and 5 spin systems) are discussed.  相似文献   

8.
We report the preparation of phosphoric acid doped poly(2,5-benzimidazole) (ABPBI) membranes for PEMFC by simultaneously doping and casting from a poly(2,5-benzimidazole)/phosphoric acid/methanesulfonic acid (MSA) solution. The evaporation of MSA yields a very homogeneous membrane having a better controlled composition, avoiding the use of solvent-intensive procedures. Membranes have been prepared with contents of up to 3.0H3PO4 molecules per ABPBI repeating unit. These membranes achieve a maximum conductivity of 1.5 × 10−2 S cm−1 at temperatures as high as 180 °C in dry conditions. These ABPBI membranes are more conveniently prepared than those conventionally formed and doped in separate steps while featuring comparable conductivities (ABPBI × 2.7H3PO4 prepared by the soaking method showed a conductivity of 2.5 × 10−2 S cm−1 at 180 °C in dry conditions).  相似文献   

9.
The synthesis of three novel 5,5-dimethylhydantoin derivatives 2-amino-N-(4,4-dimethyl-2,5-dioxoimidazolidin-1-yl)acetamide, 2-amino-N-(4,4-dimethyl-2,5-dioxoimidazol idin-1-yl)-3-phenylpropanamide, and 2-amino-4-methyl-N-(4,4-dimethyl-2,5-dioxoimidazol idin-1-yl) pentanamide, is reported. The newly synthesized compounds have been characterized by infrared (IR), MS, and NMR (1H and 13C) spectra.  相似文献   

10.
Three routes have been explored in both a high-pressure chemical ionization (CI) source and a low-pressure Fourier transform ion cyclotron resonance (FT-ICR) cell to generate the spiro[2,5]octadienyl anion in the gas phase: (i) proton abstraction from spiro[2,5]octa-4,6-diene; (ii) expulsion of trimethysilyl fluoride by phenyl ring participation following fluoride anion attack upon the silicon centre of 2-phenylethyl trimethylsilane; and (iii) collisionally induced dissociation (CID) of the carboxylate anion of 3-phenylpropanoic acid via carbon dioxide loss. From comparison of the CID spectra of various reference [C8H9]? ions with those of the [C8H9]? ions which could be generated via the routes (i) and (iii) in the CI source it can be concluded that only the third route yields a [C8H9]?ion whose CID spectrum is not inconsistent with the one expected for the spiro[2,5]octadienyl anion. In the FT-ICR cell [C8H9]? ions are generated along all three routes; their structures have been identified by specific ion-molecule reactions and appear to be different. Route (i) yields an α-methyl benzyl anion, probably due to isomerization within the ion-molecule complex formed. An ortho-ethylphenyl anion is formed along route (ii), presumably due to an intramolecular ortho proton abstraction in the generated trimethylsilyl fluoride solvated 2-phenylethyl primary carbanion. The [C8H9]? ion formed along route (iii) shows reactions similar to those of the 1,1-dimethylcyclohexadienyl anion which is structurally related to the spiro[2,5]octadienyl anion. Furthermore, the [C8H9]? ion generated via route (iii) reacts with hexafluorobenzene under expulsion of only one hydrogen fluoride molecule which contains exclusively one of the original phenyl ring hydrogen atoms. On the basis of all these observations it is therefore quite likely that the spiro[2,5]octadienyl anion is formed by collisionally induced decarboxylation of the 3-phenylpropanoic acid carboxylate anion and can be a long-lived and stable species in the gas phase.  相似文献   

11.
Three lanthanide-organic frameworks formulated as [Yb2(1,3-pda)3(H2O)]n·nH2O (1) [La2(2,5-pydc)3(H2O)2]n (2) and [La(2,5-pydc)(2,5-Hpydc)(H2O)2]n·nH2O (3) (H21,3-pda=1,3-phenylenediacetic acid, H22,5-pydc=pyridine-2,5-dicarboxylic acid) have been synthesized under hydrothermal conditions and characterized by single-crystal X-ray diffraction. Both the frameworks of compounds 1 and 2 exhibit intricate 3D architectures which can be simplified as nets with mixed nodes. Compound 1 presents a very complicated net with five types of nodes comprising intersecting (3,4)-connected and CdSO4 nets. Compound 2 possesses a (4,4,6)-connected net with (4284)(4462)2(4966)2 circuit symbol. While compound 3 is a 2D layer based upon carboxyl-bridged LaIII chains.  相似文献   

12.
Poly(benzobisoxazoles) (PBOs), poly(benzobisthiazoles) (PBTs) and copolymers thereof containing the 2,5-dihydroxybicyclo[2.2.2]octane moiety have been prepared and studied. The homopolymers were synthesized by the polycondensation of 2,5-dihydroxybicyclo[2.2.2]octane-1,4-dicarboxylic acid with 4,6-diamino-1,3-benzenediol dihydrochloride or 2,5-diamino-1,4-benzenedithiol dihydrochloride in poly(phosphoric acid). Random and block copolymers (PBO–PBT) were also prepared. The polymers were characterized by solubility, X-ray diffraction, spectroscopy (infrared and solid-state 13C nuclear magnetic resonance), and thermal analysis such as differential scanning calorimetry and thermogravimetric analysis. Thermogravimetric analysis showed thermal stability of the polymers above 375°C in air and under argon atmosphere. The polymers exhibited high resistance to organic and inorganic solvents. The polymers were converted to the more stable aromatic polymers via dehydration and retro Diels–Alder reactions of the 2,5-dihydroxybicyclo[2.2.2]octyl moiety by pyrolysis. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 277–281, 1998  相似文献   

13.
The photochemical 1,3-adduct of benzene to cis-3,4-dichlorocyclobutene has been transformed in four steps to 2,5-endo-10-oxotricyclo[4.3.1.02,5]deca-3,8-diene 1. The key reaction is the acid catalyzed addition of water to the photoadduct 2 which gives the 3,4-dichloroalcohol 3 corresponding to 1. Similarly the photochemical 1,3-adduct of benzene to cyclobutene has also been cleaved with acid. A more convenient synthesis of 1 is provided by the photochemical 1,3-addition of anisole to cis-3,4-dichlorocyclobutene and the treatment of the adduct with acid. The 3,4-dichloro derivative of 1 is thereby obtained directly. Probable mechanisms for these transformations are discussed.  相似文献   

14.
Due to the severe damage caused by free hydroxyl radicals (OH·) to cells and tissues, there is much interest in finding and studying effective and non-toxic OH· scavengers, including traditional Chinese herbs. In this paper, the simple and highly-sensitive technique of capillary zone electrophoresis with amperometric detection (CZE-AD) was used to study the OH· scavenging activities of aqueous extracts from some traditional Chinese herbs. Salicylic acid (SAL) was used as an OH· trap, and the content of OH· could be determined by assaying their products, 2,3-dihydroxybenzoic acid (2,3-DHBA) and 2,5-dihydroxybenzoic acid (2,5-DHBA). The optimum conditions for CZE-AD for the determination of 2,3-DHBA and 2,5-DHBA were explored. The linearity ranges of 2,3-DHBA and 2,5-DHBA were 1.0 ×10–7~1.0 ×10–4 mol L–1, and their detection limits were as low as 2×10–8 mol L–1, which were much better than the CE-UV method often used. The traditional Chinese herbs studied included Radix angelicae sinensis, Rhizoma coptidis, Ligustrum lucidum, Ligusticum wallichii, Radices glycyrrhizae and Semen plantaginis. The experiments showed that the aqueous extracts from all of the above traditional Chinese herds had free OH· scavenging activities, although to different degrees.  相似文献   

15.
New macrocyclic polyether compounds containing a 2,5‐bis(2‐hydroxyphenyl)‐1,3,4‐thiadiazole moiety have been prepared by a nucleophilic substitution reaction involving ethylene or polyethylene glycol ditosylate and a bisphenol, the 2,5‐bis(2‐hydroxyphenyl)‐1,3,4‐thiadiazole, with solid anhydrous carbonate as a base. The structures of the macrocycles obtained were firmly established by 1H and 13C nmr spectroscopy and their mass spectra.  相似文献   

16.
Abstract

The peracid oxidation of substituted thiophenes to thiophene 1, 1-dioxides has been extensively investigated by Melles and Backer1,2. Most oxidations, especially those of 3,4-disubstituted thiophenes, afford the sulfones in fair to good yields. Oxidation of 2,5-dimehtylthiophene, however, only yields 14% of the corresponding dioxide. As we needed considerable amounts of this dioxide and related compounds for oxidation studies3 and the preparation of cyclo-heptatrienes4, we had to improve the method of its synthesis. We found that using the commercially available m-chloroperbenzoic acid as the oxidant, freezing out (?55 °C) most of the m-chlorobenzoic acid formed and subsequently removing the residual acid with the aid of tert-amino-substituted macroreticular resins, the yield could be increased considerably (Table I). An additional advantage of using the above-mentioned ion-exchange resins is that they decompose any excess of the peracid. This procedure was also successfully applied to the synthesis of other 2,5- or 2,4-disubstituted thiophene 1,1-dioxides (see Table I).  相似文献   

17.
Reactions of 3-acetyl-2,5-dimethylthiophene with thiosemicarbazide and semicarbazide hydrochloride resulted in the formation of new heterocyclic ketimines, 3-acetyl-2,5-dimethylthiophene thiosemicarbazone (C9H13N3OS2 or L1H) and 3-acetyl-2,5- dimethylthiophene semicarbazone (C9H13N3OS or L2H), respectively. The Pd(II) and Pt(II) complexes have been synthesized by mixing metal salts in 1:2 molar ratios with these ligands by using microwave as well as conventional heating method for comparison purposes. The authenticity of these ligands and their complexes has been established on the basis of elemental analysis, melting point determinations, molecular weight determinations, IR, 1H NMR and UV spectral studies. These studies showed that the ligands coordinate to the metal atom in a monobasic bidentate manner and square planar environment around the metal atoms has been proposed to the complexes. Both the ligands and their complexes have been screened for their antimicrobial activities. The antiamoebic activity of both the ligands and their palladium compounds against the protozoan parasite Entamoeba histolytica has been tested.  相似文献   

18.
A unique supramolecular framework, [Zn4(H2O)2(2,5-tdc)4(3,3'-bpe)3]n ( 1 ), was prepared by the self-assembly of Zn(NO3)2 · 6H2O, 2,5-thiophenedicarboxylic acid (2,5-H2tdc), and 1,2-bis(3-pyridyl)-ethene (3,3'-bpe) under hydrothermal conditions. The coordination network of 1 can be simplified as a (3,4)-connected dmc framework with a point symbol (4 · 82)(4 · 85). The void space of a single network of 1 is filled by mutual interpenetration of four crystallographically equivalent nets, generating a fivefold interpenetrating architecture. Interestingly, the strong hydrogen bonds between the adjoining coordination networks further connect the interpenetrating architecture into a three-dimensional supramolecular framework. Take the hydrogen bonds into consideration, the supramolecular structure of the title compound can be further regarded as an unprecedented 5-nodal framework. The thermal stability, photoluminescent and photocatalytic properties of the title compound have also been investigated.  相似文献   

19.
3,4-Diphenylfuran-2,5-dicarboxylic acid bis-N-phenylamide 1 and 3,4-biphenyl-furan-2,5-dicarboxylic acid bis-N-butylamide 2 have been synthesised and shown to act as fluoride selective anion receptors in DMSO-d6/0.5% water solution.  相似文献   

20.
In pH 5.0, 0.1 mol l−1 NaAc-HAc buffer solution, 2,4-dinitrophenol and 2,5-dinitrophenol exhibited sensitive and distinguishable voltammetric responses at the glassy carbon electrode modified with poly-aspartic acid. By measuring the reduction peak currents of nitro groups in different positions, dinitrophenol isomers have been determined simultaneously and quantitatively. The linear calibration ranges were 1.1 × 10−6–6.0 × 10−4 mol l−1 for 2,4-dinitrophenol and 7.0 × 10−7–6.0 × 10−4 mol l−1 for 2,5-dinitrophenol, with detection limits of 2.7 × 10−7 and 1.1 × 10−7 mol l-1 respectively. This method has been applied to the detection of dinitrophenols in simulation water sample, and the recovery was from 96.7 to 102.5%.  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