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.
In the crystal structure of (E)‐8‐(3‐chloro­styr­yl)‐1,3,7‐trimethylxanthine (CSC) [systematic name: (E)‐8‐(3‐chloro­styr­yl)‐1,3,7‐trimethyl‐3,7‐dihydro‐1H‐purine‐2,6‐dione], C16H15ClN4O2, the xanthine ring and the lateral styryl chain are coplanar. The crystal packing involves mainly parallel stacking of these planar mol­ecules. The electrostatic potential calculated on the crystal structure conformation confirms the pharmacophore elements associated with MAO‐B inhibition.  相似文献   

2.
1,3,7,8‐Tetraphenyl‐4,8‐dihydro‐1H‐imidazolo[4,5g][1,2,4]benzotriazin‐4‐yl ( 5 ), 8‐(4‐bromophenyl)‐1,3,7‐triphenyl‐4,8‐dihydro‐1H‐imidazolo[4,5g][1,2,4]benzotriazin‐4‐yl ( 6 ), and 8‐(4‐methoxyphenyl)‐1,3,7‐triphenyl‐4,8‐dihydro‐1H‐imidazolo[4,5g][1,2,4]benzotriazin‐4‐yl ( 7 ) were characterized by using X‐ray diffraction crystallography, variable‐temperature magnetic susceptibility studies, and DFT calculations. Radicals 5 – 7 pack in 1 D π stacks made of radical pairs with alternate short and long interplanar distances. The magnetic susceptibility (χ vs. T) of radicals 5 and 6 exhibit broad maxima at (50±2) and (50±4) K, respectively, and are interpreted in terms of an alternating antiferromagnetic Heisenberg linear chain model with average exchange‐interaction values of J=?31.3 and ?35.4 cm?1 (gsolid=2.0030 and 2.0028) and an alternation parameter a=0.15 and 0.38 for 5 and 6 , respectively. However, radical 7 forms 1 D columns of radical pairs with alternating distances; one of the interplanar distances is significantly longer than the other, which decreases the magnetic dimensionality and leads to discrete dimers with a ferromagnetic exchange interaction between the radicals (2J=23.6 cm?1, 2zJ′=?2.8 cm?1, gsolid=2.0028). Magnetic exchange‐coupling interactions in 1,2,4‐benzotriazinyl radicals are sensitive to the degree of slippage and inter‐radical separation, and such subtle changes in structure alter the fine balance between ferro‐ and antiferromagnetic interactions.  相似文献   

3.
《化学:亚洲杂志》2018,13(19):2868-2880
The reaction of 3,7‐diacetyl‐1,3,7‐triaza‐5‐phosphabicyclo[3.3.1]nonane (DAPTA) with metal salts of CuII or NaI/NiII under mild conditions led to the oxidized phosphane derivative 3,7‐diacetyl‐1,3,7‐triaza‐5‐phosphabicyclo[3.3.1]nonane‐5‐oxide (DAPTA=O) and to the first examples of metal complexes based on the DAPTA=O ligand, that is, [CuII(μ‐CH3COO)2O‐DAPTA=O)]2 ( 1 ) and [Na(1κOO′;2κO‐DAPTA=O)(MeOH)]2(BPh4)2 ( 2 ). The catalytic activity of 1 was tested in the Henry reaction and for the aerobic 2,2,6,6‐tetramethylpiperidin‐1‐oxyl (TEMPO)‐mediated oxidation of benzyl alcohol. Compound 1 was also evaluated as a model system for the catechol oxidase enzyme by using 3,5‐di‐tert‐butylcatechol as the substrate. The kinetic data fitted the Michaelis–Menten equation and enabled the obtainment of a rate constant for the catalytic reaction; this rate constant is among the highest obtained for this substrate with the use of dinuclear CuII complexes. DFT calculations discarded a bridging mode binding type of the substrate and suggested a mixed‐valence CuII/CuI complex intermediate, in which the spin electron density is mostly concentrated at one of the Cu atoms and at the organic ligand.  相似文献   

4.
In this research study, the formation and characterization of new ruthenium(II) and (III) complexes encompassing multidentate ligands derived from 6-acetyl-1,3,7-trimethyllumazine (almz) are reported. The 1:1 molar coordination reactions of trans-[RuCl2(PPh3)3] with N-1-[1,3,7-trimethyllumazine]benzohydride (bzlmz) and 6-(N-methyloxime)-1,3,7-trimethyllumazine (ohlmz) formed a diamagnetic ruthenium(II) complex, cis-[RuCl2(bzlmz)(PPh3)] (1), and paramagnetic complex, cis-[RuIIICl2(olmz)(PPh3)] (2) [Holmz = 6-(N-hydroxy-N′-methylamino)-1,3,7-trimethyllumazine], respectively. These ruthenium complexes were characterized via physico-chemical and spectroscopic methods. Structural elucidations of the metal complexes were confirmed using single crystal X-ray analysis. The redox properties of the metal complexes were investigated via cyclic voltammetry. Electron spin resonance spectroscopy confirmed the presence of a paramagnetic metal centre in 2. The radical scavenging activities of the metal complexes were explored towards the DPPH and NO radicals. Quantum calculations at the density functional theory level provided insight into the interpretation of the IR and UV–Vis experimental spectra of 1.  相似文献   

5.
The title compound {systematic name: tetra­kis(μ‐3,5‐dinitro­benzoato‐κ2O:O′)bis­[(3,7‐dihydro‐1,3,7‐trimethyl‐1H‐purine‐2,6‐dione‐κO2)copper(II)]}, [Cu2(C7H3N2O6)4(C8H10N4O2)2], consists of paddle‐wheel dimeric tetra­kis(μ‐3,5‐dinitro­benzoato‐κ2O:O′)dicopper(II) units with O‐coordinated caffeine mol­ecules in both apical positions. The entire dimeric mol­ecule lies on a tetra­gonal inversion axis, and thus one nitro­benzoate anion with one Cu atom in a special position belong to the independent part of the mol­ecule. The caffeine ligand bonded to the Cu atom is disordered on a local twofold non‐crystallographic axis coincident with the axis. A π–π stacking inter­action is observed between the caffeine rings and adjacent symmetry‐related benzene rings of the 3,5‐dinitro­benzoate anions.  相似文献   

6.
The ability of urea anions to react as nucleophiles with alkoxy derivatives of 1,3,7‐triazapyrenes has been investigated. It was found that against all expectations, the products of the substitution of an alkoxy groups (SNipso ) by amino group were isolated in good yields. The reactions proceed in anhydrous dimethyl sulfoxide solution at room temperature. But when anions of the mono‐substituted ureas containing bulky substituents were used, the first products of the earlier unknown SNAr reactions of alkyl carbamoyl amination were obtained.  相似文献   

7.
Lesinurad (systematic name: 2‐{[5‐bromo‐4‐(4‐cyclopropylnaphthalen‐1‐yl)‐4H‐1,2,4‐triazol‐3‐yl]sulfanyl}acetic acid, C17H14BrN3O2S) is a selective uric acid reabsorption inhibitor related to gout, which exhibits poor aqueous solubility. High‐throughput solid‐form screening was performed to screen for new solid forms with improved pharmaceutically relevant properties. During polymorph screening, we obtained two solvates with methanol (CH3OH) and ethanol (C2H5OH). Binary systems with caffeine (systematic name: 3,7‐dihydro‐1,3,7‐trimethyl‐1H‐purine‐2,6‐dione, C8H10N4O2) and nicotinamide (C6H6N2O), polymorphs with urea (CH4N2O) and eutectics with similar drugs, like allopurinol and febuxostat, were prepared using the crystal engineering approach. All these novel solid forms were confirmed by XRD, DSC and FT–IR. The crystal structures were solved by single‐crystal and powder X‐ray diffraction. The crystal structures indicate that the lesinurad molecule is highly flexible and the triazole moiety, along with the rotatable thioacetic acid (side chain) and cyclopropane ring, is almost perpendicular to the planar naphthalene moiety. The carboxylic acid–triazole heterosynthon in the drug is interrupted by the presence of methanol and ethanol molecules in their crystal structures and forms intermolecular macrocyclic rings. The caffeine cocrystal maintains the consistency of the acid–triazole heterosynthons as in the drug and, in addition, they are bound by several auxiliary interactions. In the binary system of nicotinamide and urea, the acid–triazole heterosynthon is replaced by an acid–amide synthon. Among the urea cocrystal polymorphs, Form I (P, 1:1) consists of an acid–amide (urea) heterodimer, whereas in Form II (P21/c, 2:2), both acid–amide heterosynthons and urea–urea dimers co‐exist. Density functional theory (DFT) calculations further support the experimentally observed synthon hierarchies in the cocrystals. Aqueous solubility experiments of lesinurad and its binary solids in pH 5 acetate buffer medium indicate the apparent solubility order lesinurad–urea Form I (43‐fold) > lesinurad–caffeine (20‐fold) > lesinurad–allopurinol (12‐fold) ? lesinurad–nicotinamide (11‐fold) > lesinurad, and this order is correlated with the crystal structures.  相似文献   

8.
A series of 6′‐chloro‐1′,1′‐dioxo‐2′H‐spiro[benzo[d][1,3,7]oxadiazocine‐4,3′‐(1,4,2‐benzodithiazine)]‐2,6(1H,5H)‐dione derivatives 2a , 2b and 3a , 3b have been synthesized starting from 3‐aminobenzodithiazines 1a , 1b and isatoic anhydride. Subsequent reactions of 2a with 3‐chlorophenyl isocyanate gave condensation products 4 and 5 . Compound 2a was also converted into 3‐(2‐aminobenzamido)‐6‐chloro‐7‐methyl‐1,1‐dioxo‐1,4,2‐benzodithiazine derivatives 6 , 7 , 8 , 9 , 10 . The mechanisms of the reactions are discussed.  相似文献   

9.
Acylations of 1,3-dimethyl- ( 1 ) and 1,3,7-trimethylpyrrolo[2,3-d]pyrimidine-2,4-dione ( 2 ) with anhydrides in the presence of trifluoroacetic acid proceed well to give in good yields the corresponding 7-acyl derivatives 3–11 . The 6-trichloroacetyl derivatives 5 and 6 are sensitive towards nucleophiles, which displace the trichloromethyl group easily by formation of the corresponding 6-carboxylic acid derivatives 12–23. The newly synthesized compounds have been characterized by elemental analysis, uv and 1H nmr spectra and pKa, determinations.  相似文献   

10.
1,3‐Dipolar cycloadditions of C‐(5‐nitro‐2‐furyl)‐N‐methyl nitrilimine (2a) , C‐(5‐nitro‐2‐furyl)‐N‐phenyl nitrilimine (2b) , C‐4‐nitrophenyl‐N‐methyl nitrilimine (2c) and C,N‐diphenyl nitrilimine (2d) with 1‐R‐substituted 3,3‐methylene‐5,5‐dimethylpyrrolidin‐2‐ones (1a‐d) where R is H, acetyl, 1,1‐dimethylethoxycarbonyl and 1‐methylethenyl proceed with complete regioselectivity in good yields to afford 1,3,7‐trisubstituted‐6‐oxo‐8,8‐dimethyl‐1,2,7‐triazaspiro[4,4]non‐2‐enes (5a‐g) exclusively. Cycloaddition of C‐(5‐nitro‐2‐furyl)‐N‐phenylnitrone (3b) to the exocyclic double bond of the dipolarophile 1a proceeds to 2‐phenyl‐3‐(5‐nitro‐2‐furyl)‐6‐oxo‐8,8‐dimethyl‐1‐oxa‐2,7‐diazaspiro[4,4]nonane (7) with complete regio‐ and stereoselectivity.  相似文献   

11.
Conclusions The systems Pd(acac)2-Ph3P and Pd(OAc)2-Ph3P cause the dimerization and telomerization of isoprene in methanol to give linear isoprene dimers and methoxydimethyloctadienes. In isopropanol these systems dimerize isoprene to 2,7-dimethyl-1,3,7-octatriene.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 9, pp. 2099–2100, September, 1976.  相似文献   

12.
《中国化学会会志》2017,64(7):843-850
The organic salts 1‐(2‐pyridylmethyl)‐3‐alkylbenzimidazolium halide (pm‐RbH +X) and 1‐(2‐pyridylmethyl)‐3‐alkylimidazolium halide (pm‐R′iH +X′) were prepared (where R = 4‐, 3‐, 2‐fluorobenzyl ( 4f , 3f , and 2f , respectively), 4‐, 3‐, 2‐chlorobenzyl ( 4c , 3c , and 2c , respectively); 4‐methoxybenzyl (4mo); 2,3,4,5,6‐pentafluorobenzyl (f5); benzyl (b); and methyl (m)); X = Cl and Br; R′ = benzyl (b) and methyl (m); and X′ = Cl and I. From these salts, heteroleptic Ir(III ) complexes containing one N ‐heterocyclic carbene (NHC ) ligand [Ir(κ2‐ppy)22‐(pm‐Rb))]PF6 (R = 4f, 1 (PF6 ); 3f, 2 (PF6 ); 2f, 3 (PF6 ); f5b, 4 (PF6 ); 4c, 5 (PF6 ); 3c, 6 (PF6 ); 2c, 7 (PF6 ); 4mo, 8 (PF6 ); b, 9 (PF6 ); m, 10 (PF6 )) and [Ir(κ2‐ppy)22‐(pm‐R′i))]PF6 (R = b, 11 (PF6 ); m, 12 (PF6 )), were synthesized, and the crystal structures of 1 (PF6 ), 2 (PF6 ), 3 (PF6 ), 5 (PF6 ), 6 (PF6 ), 7 (PF6 ), 9 (PF6 ), 10 (PF6 ), and 12 (PF6 ) were determined by X‐ray diffraction. The neutral NHC ligands 1‐(2‐pyridylmethyl)‐3‐alkylbenzimidazolin‐2‐ylidene (pm‐Rb) and 1‐(2‐pyridylmethyl)‐3‐alkylimidazolin‐2‐ylidene (pm‐R′i) of all cations were found to be involved in the intermolecular π−π stacking interactions with the surrounding cations in the solid state, thereby probably influencing the photophysical behavior in the solid state and in solution. The absorption and emission properties of all the complexes show only small variations.  相似文献   

13.
Complexes of the Alkali Metal Tetraphenylborates with Macrocyclic Crown Ethers Alkali metal tetraphenylborates, MB(C6H5)4 (M = Li to Cs), react in tetrahydrofuran with macrocyclic crown ethers to give complexes of the general formula MB(C6H5)4(crown)m(THF)n. Suitable single crystals for X‐ray structure analysis were grown from a solvent mixture of tetrahydrofuran and n‐hexane. The salt like complexes [Li(12‐crown‐4)(thf)][B(C6H5)4] ( 1 ), [Na(15‐crown‐5)(thf)][B(C6H5)4] ( 2 ), and [Cs(18‐crown‐6)2][B(C6H5)4] · THF ( 6 ), the mononuclear molecular complexes [KB(C6H5)4(18‐crown‐6)(thf)] ( 3 ), [RbB(C6H5)4(18‐crown‐6)] ( 4 ), and [CsB(C6H5)4(18‐crown‐6)] · THF ( 5 ), and the compound [CsB(C6H5)4(18‐crown‐6)]2[Cs(18‐crown‐6)2][B(C6H5)4] ( 7 ), which contains a binuclear molecule ([CsB(C6H5)4(18‐crown‐6)]2) beside a [Cs(18‐crown‐6)2]+ cation and a [B(C6H5)4]? anion, are described. All compounds are charactarized by infrared spectra, elemental analysis, NMR‐spectroscopy, and X‐ray single crystal structure analysis.  相似文献   

14.
Reaction fo (bipyridine)(1,5-cyclooctadiene)nickel with carbon dioxide and 1,3,7-octatriene yields products which indicate that the CO2 attacks only at the diene part of octatriene. In the same reaction 1,3,5-hexatriene couples with CO2 to yield 1/1, 1/2 and 2/2 adducts.  相似文献   

15.
Introducing substituents in the 6‐position of the 2‐pyridyl rings of tris(pyridyl)aluminate anions, of the type [EtAl(2‐py′)3]? (py′=a substituted 2‐pyridyl group), has a large impact on their metal coordination characteristics. This is seen most remarkably in the desolvation of the THF solvate [EtAl(6‐Me‐2‐py)3Li?THF] to give the monomer [EtAl(6‐Me‐2‐py)3Li] ( 1 ), containing a pyramidal, three‐coordinate Li+ cation. Similar monomeric complexes are observed for [EtAl(6‐CF3‐2‐py)3Li] ( 2 ) and [EtAl(6‐Br‐2‐py)3Li] ( 3 ), which contain CF3 and Br substituents (R). This steric influence can be exploited in the synthesis of a new class of terminal Al?OH complexes, as is seen in the controlled hydrolysis of 2 and 3 to give [EtAl(OH)(6‐R‐2‐py)2]? anions, as in the dimer [EtAl(OH)(6‐Br‐2‐py)2Li]2 ( 5 ). Attempts to deprotonate the Al?OH group of 5 using Et2Zn led only to the formation of the zincate complex [LiZn(6‐Br‐py)3]2 ( 6 ), while reactions of the 6‐Br substituted 3 and the unsubstituted complex [EtAl(2‐py)3Li] with MeOH give [EtAl(OMe)(6‐Br‐2‐py)2Li]2 ( 7 ) and [EtAl(OMe)(2‐py)2Li]2 ( 8 ), respectively, having similar dimeric arrangements to 5 . The combined studies presented provide key synthetic methods for the functionalization and elaboration of tris(pyridyl)aluminate ligands.  相似文献   

16.
Taking advantage of an improved synthesis of [Ti(η6‐C6H6)2], we report here the first examples of ansa‐bridged bis(benzene) titanium complexes. Deprotonation of [Ti(η6‐C6H6)2] with nBuLi in the presence of N,N,N′,N′′,N′′‐pentamethyldiethylenetriamine (pmdta) leads to the corresponding 1,1′‐dilithio salt [Ti(η6‐C6H5Li)2] ? pmdta that enables the preparation of the first one‐ and two‐atom‐bridged complexes by simple salt metathesis. The ansa complexes were fully characterized (NMR spectroscopy, UV/Vis spectroscopy, elemental analysis, and X‐ray crystallography) and further studied electrochemically and computationally. Moreover, [Ti(η6‐C6H6)2] is found to react with the Lewis base 1,3‐dimethylimidazole‐2‐ylidene (IMe) to give the bent sandwich complex [Ti(η6‐C6H6)2(IMe)].  相似文献   

17.
A novel free radical reaction combined with liquid chromatography electrospray ionization tandem mass spectrometry (FRR-LC–PDA-ESI/APCI-MS/MS) screening method was developed for the detection and identification of radical-scavenging natural antioxidants. Functionalized graphene was prepared by chemical method for loading free radicals (superoxide radical, peroxyl radical and PAHs free radical). Separation was performed with and without a preliminary exposure of the sample to specific free radicals on the functionalized graphene, which can facilitate reaction kinetics (charge transfers) between free radicals and potential antioxidants. The difference in chromatographic peak areas is used to identify potential antioxidants. The structure of the antioxidants in one sample (Swertia chirayita) is identified using MS/MS and comparison with standards. Thirteen compounds were found to possess potential antioxidant activity, and their free radical-scavenging capacities were investigated. The thirteen compounds were identified as 1,3,5-trihydroxyxanthone-8-O-β-d-glucopyranoside (PD1), norswertianin (PD2), 1,3,5,8-tetrahydroxyxanthone (PD3), 3, 3′, 4′, 5, 8-penta hydroxyflavone-6-β-d-glucopyranosiduronic acid-6′-pentopyranose-7-O-glucopyranoside (PD4), 1,5,8-trihydroxy-3-methoxyxanthone (PD5), swertiamarin (PS1), 2-C-β-d-glucopyranosyl-1,3,7-trihydroxylxanthone (PS2), 1,3,7-trihydroxylxanthone-8-O-β-d-glucopyranoside (PL1), 1,3,8-trihydroxyl xanthone-5-O-β-d-glucopyranoside (PL2), 1,3,7-trihydroxy-8-methoxyxanthone (PL3), 1,2,3-trihydroxy-7,8-dimethoxyxanthone (PL4), 1,8-dihydroxy-2,6-dimethoxy xanthone (PL5) and 1,3,5,8-tetramethoxydecussatin (PL6). The reactivity and SC50 values of those compounds were investigated, respectively. PD4 showed the strongest capability for scavenging PAHs free radical; PL4 showed prominent scavenging capacities in the lipid peroxidation processes; it was found that all components in S. chirayita exhibited weak reactivity in the superoxide radical scavenging capacity. The use of the free radical reaction screening method based on LC–PDA-ESI/APCI-MS/MS would provide a new approach for rapid detection and identification of radical-scavenging natural antioxidants from complex matrices.  相似文献   

18.
Syntheses are described of a number of 2,6‐difunctionalized dimethylsilylbenzenes, namely, 1‐(HMe2Si)‐2,6‐Cl2C6H3 ( 13 ), 1‐(HMe2Si)‐2,6‐Br2C6H3 ( 14 ), 1,2,3‐(HMe2Si)3C6H3 ( 15 ), 1,2‐(HMe2Si)2‐6‐ClC6H3 ( 16 ), 1,2‐(HMe2Si)2‐6‐BrC6H3 ( 17 ), 1‐(HMe2Si)‐2‐(Ph2P)‐6‐BrC6H3 ( 18 ), diphenyl(1,1,3,3‐tetramethyl‐1,3‐dihydrobenzo[c][1,2,5]oxadisilol‐4‐yl)phosphine oxide ( 19 ) and 8‐Brom‐1,1,3,3‐tetramethyl‐2,2,2,2,‐tetracarbonyl‐1,3‐dihydro‐benzo[d][2,1,3]ferra disilol ( 20 ). Compounds 13 – 20 were characterized by multinuclear NMR spectroscopy and in case of 18 – 20 also by single crystal X‐ray diffraction.  相似文献   

19.
A series of palladium complexes ( 2a–2g ) ( 2a : [6‐tBu‐2‐PPh2‐C6H3O]PdMe(Py); 2b : [6‐C6F5–2‐PPh2‐C6H3O]PdMe(Py); 2c : [6‐tBu‐2‐PPhtBu‐C6H3O]PdMe(Py); 2d : [2‐PPhtBu‐C6H4O] PdMe(Py); 2e : [6‐SiMe3–2‐PPh2‐C6H3O]PdMe(Py); 2f : [2‐tBu‐6‐(Ph2P=O)‐C6H3O]PdMe(Py); 2g : [6‐SiMe3–2‐(Ph2P=O)‐C6H3S]PdMe(Py)) bearing phosphine (oxide)‐(thio) phenolate ligand have been efficiently synthesized and characterized. The solid‐state structures of complexes 2d , 2f and 2g have been further confirmed by single‐crystal X‐ray diffraction, which revealed a square‐planar geometry of palladium center. In the presence of B(C6F5)3, these complexes can be used as catalysts to polymerize norbornene (NB) with relatively high yields, producing vinyl‐addition polymers. Interestingly, 2a /B(C6F5)3 system catalyzed the polymerization of NB in living polymerization manner at high temperature (polydispersity index 1.07, Mn up to 1.5 × 104). The co‐polymerization of NB and polar monomers was also studied using catalysts 2a and 2f . All the obtained co‐polymers could dissolve in common solvent.  相似文献   

20.
The di(1,3,7-trimethylpurine-2,6-dione) dihydrogen 12-tungstosilicate (C8N4O2H11)2H2SiW12O40 (I) was synthesized and studied using chemical methods, IR and 1H NMR spectroscopy, X-ray phase analysis, and the thermogravimetry 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