共查询到20条相似文献,搜索用时 15 毫秒
1.
Dr. Benedikt Schwarze Sanja Jelača Linda Welcke Prof. Dr. Danijela Maksimović-Ivanić Prof. Dr. Sanja Mijatović Prof. Dr. Dr. h.c. mult. Evamarie Hey-Hawkins 《ChemMedChem》2019,14(24):2075-2083
Investigations on the antitumor activity of metallacarboranes are sparse in the literature and limited to a handful of ruthena- and molybdacarboranes. In this study, the molybdacarborane fragment [3-(CO)2-closo-3,1,2-MoC2B9H11] was combined with a vector molecule, inspired by the well-known drug tamoxifen or 4,4′-dihydroxytamoxifen (TAM-diOH). The molybdacarborane derivative [3,3-{4-[1,1-bis(4-hydroxyphenyl)but-1-en-2-yl]-2,2′-bipyridine-κ2N,N′}-3-(CO)2-closo-3,1,2-MoC2B9H11] ( 10 ), as well as the ligand itself 4-[1,1-bis(4-hydroxyphenyl)but-1-en-2-yl]-2,2′-bipyridine ( 6 ) showed cytotoxic activities in the low micromolar range against breast adenocarcinoma (MDA-MB-231, MDA-MB-361 and MCF-7), human glioblastoma (LN-229) and human glioma (U-251) cell lines. In addition, compounds 6 and 10 were found to induce senescence and cytodestructive autophagy, lower ROS/RNS levels, but only the molybdacarborane 10 induced a strong increase of nitric oxide (NO) concentration in the MCF-7 cells. 相似文献
2.
3.
《Journal of Sulfur Chemistry》2013,34(4):275-282
Atropisomers of 2,2′-bithiophene-5-tetrahydroisoquinoline, S -8, R -8, and 2,2′-bithiophene-5,5′-tetrahydroisoquinoline, (S,S)-11, (R,S)-11, (R,R)-11, analogs of michellamines have been synthesized in low yield under Stille coupling conditions (Pd0-mediated cross coupling reactions) of stannanes 1 or 2 with non-racemic bromotetrahydroisoquinoline 6. The use of non-racemic iodotetrahydroisoquinoline 7 instead of 6 significantly improves the efficiency of the coupling, affording the atropisomers in moderate yields. 相似文献
4.
In this paper, a copolymer of carbazole (Cz) and 2,2′:5′,2″-terthiophene (TTh) was electropolymerized in 0.1 M sodium perchlorate (NaClO4)/acetonitrile (CH3CN) on glassy carbon electrode. The optimum conditions of resulting homopolymers of Cz, TTh and copolymer of Cz and TTh in the initial feed ratio of [Cz]0/[TTh]0 = 1/10 were characterized by cyclic voltammetry, Fourier-transform infrared-attenuated total reflectance, scanning electron microscopy, energy dispersive X-ray analysis, and electrochemical impedance spectroscopy. Morphological analysis of copolymer shows that a micro-spherical and web-like morphology was formed for copolymer at different initial feed ratios of [Cz]0/[TTh]0 = 1/2, 1/5 and 1/10. The capacitive behavior of the modified electrodes was defined via Nyquist, Bode-magnitude, and Bode-phase plots. The highest low-frequency capacitance (C LF) was obtained as 4.11 mFcm?2 in the initial feed ratio of [Cz]0/[TTh]0 = 1/10. Double-layer capacitance (C dl) and phase angles (θ) were obtained for homopolymer and copolymer systems. The highest C dl was obtained as 2.01 mFcm?2 for the copolymer in the initial feed ratio of [Cz]0/[TTh]0 = 1/2. The highest phase angle of copolymer was obtained as θ = ~75° in the initial feed ratio of [Cz]0/[TTh]0 = 1/1. These capacitance results confirmed that films of copolymer Cz/TTh are promising materials for micro-capacitor applications. 相似文献
5.
《Journal of Sulfur Chemistry》2013,34(6):661-691
The variable transformations of the S?S bond in the presence of metal complexes generally depend upon the different reaction conditions used. In this review, the synthesis and coordination versatility of complexes derived from diaryl and 2,2′ -dipyridyl disulfide have been discussed in the absence or presence of other donor groups. 相似文献
6.
Hai-Hua Li Ning Ma Kai-Hui Li 《Journal of Inorganic and Organometallic Polymers and Materials》2012,22(6):1320-1324
A novel metal–organic framework {[Cu(H2bptc)(cbpy)(H2O)]·(H2O)} n (1) has been hydrothermally synthesized through reaction of 1,1′-biphenyl-2,2′,3,3′-tetracarboxylic acid (H4bptc) with Cu(II) salt in the presence of ancillary nitrogen ligand 5,5′-dimethyl-2,2′-bipyridine (cbpy), and its structure was determined by X-ray diffraction and characterized by elemental analysis, and IR spectrum. The title compound crystallizes in monoclinic space group P21/c. Both right- and left-handed helices are detected in the structure. In the ab-plane, adjacent chains are homochiral and parallel to each other, which are connected together by hydrogen bonds to form a 2D supramolecular structure. The supramolecular layers with opposite chirality are arranged alternatively along the c-axis to form a 3D mesomeric supramolecular structure through π···π interactions. The thermal stability of the complex 1 was studied by thermal gravimetric and differential thermal analysis. 相似文献
7.
M. Yu. Alyapyshev L. I. Tkachenko A. Paulenova A. A. Popova N. E. Borisova 《溶剂提取与离子交换》2013,31(2):138-152
The extraction of americium and lanthanides by diamides of 2,2′-dipyridyl-6,6′-dicarboxilic acid solutions in polar diluents was investigated. The dependence of extraction properties of the diamides on their structure was studied and the distribution ratios of americium and lanthanides from nitric acid solutions were determined. The highest extraction was found for the compounds with the ethyl-group and the alkyl-substituted aryl-group on the amidic nitrogen. The Am/Ln separation factors higher than 10 were achieved for extraction from high acidity media. 相似文献
8.
《Reactive and Functional Polymers》2004,58(2):111-115
Aromatic coupling copolymerization of 5,5′-bis(4-chlorobenzoyl)-2,2′-dimethoxybiphenyl (1) and 5,5′-bis(3-chlorobenzoyl)-2,2′-dimethoxybiphenyl (2) has successfully afforded higher-molecular-weight polyketones than either of the corresponding homopolymerizations. The wholly aromatic copolyketones thus obtained have highly amorphous nature with high glass transition temperatures (Tg). 相似文献
9.
10.
The ozonation of a triple bond in organic compounds is rare in industry. The reaction is much more complicated to run, because ozone is an electrophilic agent and more likely to bind to the double bond compared to the triple bond in an organic compound. This may be a problem because some fine organic substances are very hard to synthesize and can only be solved by the ozonolysis technique of a triple-bond organic compound. In addition, oxidation by using ozone is also preferred because of the cleanliness and effectiveness of the reaction. A novel method for the synthesis of 2,2?-bipyridine was described, which involves ozonolysis of a 2-ethynylpyridine and is supported by a high-pressure reactor. The reaction proceeds via ozone attack on the acetylenic triple bond, cleavage, and free radical formation. The present study establishes the conditions for selective ozonolysis of 2-ethynylpyridine, leading to 2,2?-bipyridine. The 2,2?-bipyridine was characterized by gas chromatography flame ionization detector (GC-FID) and mass spectroscopy techniques. A novel suggested mechanism was used to explain the oxidation process. Analysis showed that the formation of oxidized product is shown when ozone is supplied in excess, which shows that the reaction between 2-ethynylpyridine and ozone is a first-order reaction, whe reby the result of the relationship only depends on the concentration of the starting material, which is 2-ethynylpyridine acts as the limiting reactant. 相似文献
11.
Edwin C. Constable Catherine E. Housecroft Jason R. Price Luca Schweighauser Jennifer A. Zampese 《Inorganic chemistry communications》2010,13(4):495-497
The first example of a copper-catalysed [2 + 3] cycloaddition reaction of a coordinated 4′-azido-2,2′:6′,2″-terpyridine ligand is reported and the solid state structures of the precursor and product are described. 相似文献
12.
F. Karbasi H. Younesi M. Ardjmand A. Safe Kordi S. Yaghmaei H. Qaderi 《Chemical Engineering Communications》2016,203(2):224-235
Batch culture of Azohydromonas lata was investigated for the production of intracellular poly-β-hydroxybutyrate (PHB). In order to determine the C:N value of the culture media for maximizing the microbial productivity of PHB, different concentrations of glucose and ammonium chloride were used as carbon and nitrogen sources, respectively. The optimal temperature and shaking rate was obtained at 30°C and 180 rpm, respectively. The maximum intracellular PHB concentration obtained was 5.09 g/l, which was 20% (w/w) of the cell dry weight (CDW) after 72 h. Also, the synthesis of PHB was growth associated with the C:N ratio of 153.71. The maximum calculated Yp/s was 0.212 (gr/gr) and the specific production rate value after 12 h was 0.264 g/l/h, with 40 and 50 g/l of glucose concentrations, respectively, with 0.5 g/l ammonium chloride kept constant. The chemical composition of the resulting PHB was analyzed by Fourier transform infrared spectroscopy and proton nuclear magnetic resonance spectroscopy. The Leudeking–Piret model was used for kinetic analysis of the PHB production, the statistical analysis of which was modeled by response surface methodology. An artificial neural network technique was applied for modeling of the microbial production of PHB by A. lata as a function of the glucose concentration and CDW, where the minimum mean square error of the model was 0.0012 and 0.0038 for glucose concentrations of 50 g/l and 40 g/l, respectively, when 0.5 g/l ammonium chloride was kept constant. 相似文献
13.
Catalytic oxidative polymerization of 2,2′-dihydroxybiphenyl (DHBP) was performed by using Schiff base polymer-Cu (II) complex
and hydrogen peroxide as catalyst and oxidant, respectively. According to size exclusion chromatography (SEC) analysis, the
number-average molecular weight (M
n), weight-average molecular weight (M
w) and polydispersity index (PDI) values of poly (2,2′-dihydroxybiphenyl) (PDHBP) were found to be 37,500, 90,000 g mol−1 and 2.4, respectively. The thermal degradation kinetics was investigated by thermogravimetric analysis in dynamic nitrogen
atmosphere at four different heating rates: 5, 10, 15 and 20 °C min−1. The derivative thermogravimetry curves of PDHBP showed that its thermal degradation process had one weight-loss step. The
apparent activation energies of thermal decomposition for PDHBP as determined by Tang, Flynn–Wall–Ozawa (FWO), Kissenger–Akahira–Sunose
(KAS), Coats–Redfern (CR) and Invariant kinetic parameter (IKP) methods were 109.1, 109.0, 110.0, 108.4 and 109.8 kJ mol−1, respectively. The mechanism function and pre-exponential factor were determined by master plots and Criado–Malek–Ortega
method. The most likely decomposition process was a D
n
Deceleration type in terms of the CR, master plots and Criado–Malek–Ortega results. 相似文献
14.
《Reactive Polymers》1992,16(2):223-226
Polymer-bound 2,2′-bipyridine-iron complexes were synthesized and found to be novel catalysts for the oxidation of saturated hydrocarbons. Toluene, p-xylene, ethylbenzene, n-propylbenzene and n-butylbenzene were oxidized under an atmosphere of air or oxygen in the absence of solvent at 110°C in the presence of the iron complexes chemically immobillized on polymers, affording corresponding ketones and alcohols in most cases. Except for toluene, the conversion of alkylbenzene was 20% to 35% with the different iron complexes: The selectivity in producing ketone and alcohol in the oxidation of PhR (where R = Et, n-propyl and n-butyl was ∼100%. 相似文献
15.
The kinetics of the thermal decomposition of chalcopyrite concentrate was investigated by means of thermal analysis techniques, Thermogravimetry/Derivative thermogravimetry (TG/DTG) under ambient air conditions in the temperature range of 0–900°C with heating rates of 2, 5, 10, 15, and 20°C min?1. TG and DTG measurements showed that the thermal behavior of chalcopyrite concentrate shows a two-step decomposition. The decomposition mechanism was confirmed using X-ray diffraction (XRD), Scanning Electron Microscope (SEM)/energy-dispersive X-ray spectroscopy (EDS), and Fourier transform infrared spectroscopy (FTIR) analyses. Kinetic parameters were determined from the TG and DTG curves for steps I and II by using two model-free (isoconversional) methods—Flyn–Wall–Ozowa (FWO) and Kissinger–Akahira–Sunose (KAS). The kinetic parameters consisting of Ea, A, and g(α) models of the materials were determined. The average activation energies (Ea) obtained from both models for the decomposition of chalcopyrite concentrate were 72.55 and 300.77 kJ mol?1 and the pre-exponential factors (A) were 15.07 and 29.39 for steps I and II, respectively. The most probable kinetic model for the decomposition of chalcopyrite concentrate is an first-order mechanism, i.e., chemical reaction [g(α) = (?ln(1?α))], and an Avrami–Eroeyev equation mechanism, i.e., nucleation and growth for n = 2 [g(α) = (?ln(1?α)1/2)], for steps I and II, respectively. 相似文献
16.
概述了2,2′,6,6′-四氯双酚A的应用和合成工艺。详细研究了双酚A在二氯乙烷溶剂中经氯化反应合成2,2′,6,6′-四氯双酚A,得到了最佳合成工艺条件,收率可达915%,为工业化生产提供了可靠的基础数据 相似文献
17.
《Catalysis Reviews》2013,55(3-4):255-318
A critical review of the kinetics and selectivity of the Fischer–Tropsch synthesis (FTS) is given. The focus is on reaction mechanisms and kinetics of the water–gas shift and Fischer–Tropsch (FT) reactions. New developments in the product selectivity as well as the overall kinetics are reviewed. It is concluded that the development of rate equations for the FTS should be based on realistic mechanistic schemes. Qualitatively, there is agreement that the product distribution is affected by the occurrence of secondary reactions (hydrogenation, isomerization, reinsertion, and hydrogenolysis). At high CO and H2O pressures, the most important secondary reaction is readsorption of olefins, resulting in initiation of chain growth processes. Secondary hydrogenation of α-olefins may occur and depends on the catalytic system and the process conditions. The rates of the secondary reactions increase exponentially with chain length. Much controversy exists about whether these chain-length dependencies stem from differences in physisorption, solubility, or diffusivity. Preferential physisorption of longer hydrocarbons and increase of the solubility with chain length influences the product distribution and results in a decreasing olefin-to-paraffin ratio with increasing chain length. Process development and reactor design should be based on reliable kinetic expressions and detailed selectivity models. 相似文献
18.
《Inorganic chemistry communications》2003,6(10):1332-1334
The first 1-D double-helical-chain coordination polymer, [(H2O)Cu(BPDC)] (2,2′-biphenyldicarboxylate), based upon the binuclear square pyramidal copper(II)-pair motifs has been synthesized under hydrothermal conditions and characterized by single crystal X-ray diffraction technique. [(H2O)Cu(BPDC)] crystallized in orthorhombic crystal system, space group Cmca with unit cell dimensions a=21.073(6) Å, b=7.118(2) Å and c=17.670(5) Å, Z=8. 相似文献
19.
Edwin C. Constable Ana Hernandez Redondo Catherine E. Housecroft Markus Neuburger 《Inorganic chemistry communications》2010,13(1):70-73
Two new aldehyde-decorated tpy and bpy-containing ruthenium(II) complexes, [Ru(1)(bpy)2][PF6]2 and [Ru(1)(tpy)Cl][PF6] in which 1 is 5,5′-bis(4-formylphenyl)-2,2′-bipyridine, have been prepared and fully characterized. The packing in both solid state structures involves extensive Oaldehyde···HCpyridine contacts, but π-stacking interactions are important only between [Ru(1)(tpy)Cl]+ cations. 相似文献
20.
Hien Y. Hoang Renat Maratovich Akhmadullin Farida Yunusovna Akhmadullina Rustem Kayumovich Zakirov Dinh Nhi Bui Alfiia Garipova Akhmadullina 《Journal of Sulfur Chemistry》2018,39(2):130-139
In this paper, the intermediate and final reaction products of catalytic oxidation of inorganic sulfides in the presence of oxygen dissolved in the kerosene fraction and 3,3′,5,5′-tetra-tert-butyl-4,4′-stilbenequinone were investigated. The thiosulfate and sulfate are major products of the oxidation of sodium sulfide under these conditions. The intermediate and final products in the catalytic oxidation of sulfide sulfur do not affect the rate of its oxidation. The yield of catalytic oxidation products depends on the nature of the sulfide and on the pH of the solution. The catalytic cycle for sulfide oxidation in the presence of 3,3′,5,5′-tetra-tert-butyl-4,4′-stilbenequinone is shown. The role of 3,3′,5,5′-tetra-tert-butyl-4,4′-stilbenequinone is to create a new and a more effective way of electron transfer from the reducing agent (sulfide) to the oxidant (oxygen). 相似文献