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.
《合成通讯》2013,43(11):1447-1453
Abstract

An efficient preparation of 1‐[79Br]bromo‐2‐fluoroethylene, [79Br]BrHC?CHF, was carried out by a three‐step procedure: (a) natural 1‐bromo‐2‐fluoroethylene, BrHC?CHF, was iodinated to 1‐fluoro‐2‐iodoethylene, FHC?CHI; (b) 1‐fluoro‐2‐iodoethylene was 79Br2‐brominated to 1,2‐di[79Br]bromo‐1‐fluoro‐2‐iodoethane, [79Br]BrFCHCH[79Br]BrI; and (c) 1,2‐di[79Br]bromo‐1‐fluoro‐2‐iodoethane was dehalogenated to 1‐[79Br]bromo‐2‐fluoroethylene, [79Br]BrHC?CHF. The yield of isolated product, on a 2‐mmol scale, was 62% with respect to 79Br2.  相似文献   

2.
刘勉  叶蕴华 《中国化学》2002,20(11):1347-1353
IntroductionCyclicpeptides ,whichareconstrainedconforma tionallyandmoreresistanttoproteasedigestionsthantheirlinearprecursors ,havebeenofgreatinterestassynthetictargetsbothaspotentialdrugleadsandasmodelsforcon formationalanalysis .1 4 Currentmethodsforsynt…  相似文献   

3.
Hybrid peptides composed of α‐ and β‐amino acids have recently emerged as new class of peptide foldamers. Comparatively, γ‐ and hybrid γ‐peptides composed of γ4‐amino acids are less studied than their β‐counterparts. However, recent investigations reveal that γ4‐amino acids have a higher propensity to fold into ordered helical structures. As amino acid side‐chain functional groups play a crucial role in the biological context, the objective of this study was to investigate efficient synthesis of γ4‐residues with functional proteinogenic side‐chains and their structural analysis in hybrid‐peptide sequences. Here, the efficient and enantiopure synthesis of various N‐ and C‐terminal free‐γ4‐residues, starting from the benzyl esters (COOBzl) of N‐Cbz‐protected (E)α,β‐unsaturated γ‐amino acids through multiple hydrogenolysis and double‐bond reduction in a single‐pot catalytic hydrogenation is reported. The crystal conformations of eight unprotected γ4‐amino acids (γ4‐Val, γ4‐Leu, γ4‐Ile, γ4‐Thr(OtBu), γ4‐Tyr, γ4‐Asp(OtBu), γ4‐Glu(OtBu), and γ‐Aib) reveals that these amino acids adopted a helix favoring gauche conformations along the central Cγ? Cβ bond. To study the behavior of γ4‐residues with functional side chains in peptide sequences, two short hybrid γ‐peptides P1 (Ac‐Aib‐γ4‐Asn‐Aib‐γ4‐Leu‐Aib‐γ4‐Leu‐CONH2) and P2 (Ac‐Aib‐γ4‐Ser‐Aib‐γ4‐Val‐Aib‐γ4‐Val‐CONH2) were designed, synthesized on solid phase, and their 12‐helical conformation in single crystals were studied. Remarkably, the γ4‐Asn residue in P1 facilitates the tetrameric helical aggregations through interhelical H bonding between the side‐chain amide groups. Furthermore, the hydroxyl side‐chain of γ4‐Ser in P2 is involved in the interhelical H bonding with the backbone amide group. In addition, the analysis of 87 γ4‐residues in peptide single‐crystals reveal that the γ4‐residues in 12‐helices are more ordered as compared with the 10/12‐ and 12/14‐helices.  相似文献   

4.
A series of neutral Ir(III)‐based heteroleptic complexes with a formula of [Ir(η2‐(CN))22‐(SS))] ((CN) = ppy, (SS) = Et2NCS2 ( 2a ), MeOCS2 ( 2b ), EtOCS2 ( 2c ), iPrOCS2 ( 2d ); (CN) = tpy, (SS) = Et2NCS2 ( 3a ), MeOCS2 ( 3b ), EtOCS2 ( 3c ), iPrOCS2 ( 3d ); (CN) = epb , (SS) = Et2NCS2 ( 4a ), MeOCS2 ( 4a ), EtOCS2 ( 4a ); ppyH = 2‐phenylpyridine; tpyH = 2‐(4′‐tolyl)pyridine; epbH = ethyl 4‐(2′‐pyridyl)benzate) was synthesized and characterized. The crystal structure of complex 2d was also determined. The electron‐releasing substituents on (CN) or (SS) blueshift λmax values.  相似文献   

5.
Four cyclic peptides, diandrine A–D ( 1 – 4 ), were isolated from the MeOH extract of Formosan Drymaria diandra. Their structures were elucidated by chemical and spectroscopic analyses as cyclo(‐Gly1‐Pro2‐Trp3‐Pro4‐Tyr5‐Phe6‐), cyclo(‐Gly1‐Pro2‐Leu3‐Pro4‐Leu5‐Trp6‐Ser7‐Ser8‐), cyclo(Gly1‐Gly2‐Pro3‐Tyr4‐Trp5‐Pro6‐), and cyclo(Gly1‐Gly2‐Pro3‐Tyr4‐Trp5‐Pro6‐), respectively. Compounds 3 and 4 were stable conformational isomers. Cyclopeptide 1 showed a selective inhibitory effect on collagen‐induced platelet aggregation with an IC50 value of 44.2 μM .  相似文献   

6.
2N‐(3‐phenyl‐allyl‐)(5‐phenyl‐[1,3,4] thiadiazol‐2‐yl) amine was studied by means of the 1H, 13C, 15N NMR spectroscopy and DFT calculations. On the basis of the one‐dimensional 1H, 13C, 15N‐NMR and two‐dimensional 1H‐13C HMQC, 1H‐13C HMBC, 1H‐15N HMQC, 1H1H NOESY, 1H1H COSY correlation spectra the amine‐type and the imine‐type tautomers have been determined in the solution. Variety of structural forms including: biradical, ionic–biradical, and ionic structures of the amine‐type a and of the imine‐type b , c tautomers exist in the solution. According to the DFT computations the differences in the total energy between a and b , a and c , and b and c tautomers are equal to 1.5 kJ/mol, 1.2 kJ/mol, and 0.3 kJ/mol, respectively.  相似文献   

7.
New bridged trialkoxysilanes bearing acetoxymethyl groups were synthesized by double hydrosilylation of 1,6‐diacetoxy‐2,4‐butadiyne, using two equivalents of triethoxysilane and a metal catalyst. With a Ru catalyst, the reaction proceeded via anti‐addition to provide BTES‐Ac‐a as a single isomer, while a similar reaction with a Pt or Rh catalyst provided an isomeric mixture of syn ‐adducts BTES‐Ac‐b. Reverse osmosis (RO) silica membranes were prepared by the sol–gel process with BTES‐Ac‐a and BTES‐Ac‐b and the membranes were examined with respect to water desalination using a 2000 ppm NaCl aqueous solution. NaCl rejection of the membranes increased to reach 96% at the early stage of the RO experiments. However, the rejection decreased gradually to 85% after 70 and 200 h for BTES‐Ac‐a and BTES‐Ac‐b, respectively, due to hydrolytic decomposition of the silica network during the experiments. In contrast, a membrane prepared from copolymerization of BTES‐Ac‐a with ethane‐bridged bistrialkoxysilane (BTES‐E1) showed improved stability towards hydrolysis with stable NaCl rejection of 96% with higher water permeance (3.5 × 1013 m3 m2 s−1 Pa−1) than that of a membrane prepared by homopolymerization of BTES‐E1 (2.7 × 1014 m3 m2 s−1 Pa−1) reported previously.  相似文献   

8.
The classical synthesis, followed by purification of the steroidal A‐ring Δ1‐olefin, 5α‐androst‐1‐en‐17‐one ( 5 ), from the Δ1‐3‐keto enone, (5α,17β)‐3‐oxo‐5‐androst‐1‐en‐17‐yl acetate ( 1 ), through a strategy involving the reaction of Δ1‐3‐hydroxy allylic alcohol, 3β‐hydroxy‐5α‐androst‐1‐en‐17β‐yl acetate ( 2 ), with SOCl2, was revisited in order to prepare and biologically evaluate 5 as aromatase inhibitor for breast cancer treatment. Surprisingly, the followed strategy also afforded the isomeric Δ2‐olefin 6 as a by‐product, which could only be detected on the basis of NMR analysis. Optimization of the purification and detection procedures allowed us to reach 96% purity required for biological assays of compound 5 . The same synthetic strategy was applied, using the Δ4‐3‐keto enone, 3‐oxoandrost‐4‐en‐17β‐yl acetate ( 8 ), as starting material, to prepare the potent aromatase inhibitor Δ4‐olefin, androst‐4‐en‐17‐one ( 15 ). Unexpectedly, a different aromatase inhibitor, the Δ3,5‐diene, androst‐3,5‐dien‐17‐one ( 12 ), was formed. To overcome this drawback, another strategy was developed for the preparation of 15 from 8 . The data now presented show the unequal reactivity of the two steroidal A‐ring Δ1‐ and Δ4‐3‐hydroxy allylic alcohol intermediates, 3β‐hydroxy‐5α‐androst‐1‐en‐17β‐yl acetate ( 2 ) and 3β‐hydroxyandrost‐4‐en‐17β‐yl acetate ( 9 ), towards SOCl2, and provides a new strategy for the preparation of the aromatase inhibitor 12 . Additionally, a new pathway to prepare compound 15 was achieved, which avoids the formation of undesirable by‐products.  相似文献   

9.
A new cyclopeptide, clausenain I ( 1 ), has been isolated by a multi‐step chromatography procedure from Clausena anisum‐olens. Its structure was elucidated as cyclo (‐Gly1‐Ile2‐Ile3‐Val4‐Leu5‐Ile6‐Ile7‐Leu8‐Leu9‐) by extensive 2D‐NMR spectroscopic methods and chemical evidence. It is the first time that a natural cyclic peptide has been isolated from the genus Clausena.  相似文献   

10.
The tricyclic azoalkanes, (1α,4α,4aα,7aα)‐4,4a,5,6,7,7a‐hexahydro‐1,4,8,8‐tetramethyl‐1,4‐methano‐1H‐cyclopenta[d]pyridazine ( 1c ), (1α,4α,4aα,6aα)‐4,4a,5,6,6a‐pentahydro‐1,4,7,7‐tetramethyl‐1,4‐methano‐1H‐cyclobuta[d]pyridazine ( 1d ), (1α,4α,4aα,6aα)‐4,4a,6a‐trihydro‐1,4,7,7‐tetramethyl‐1,4‐methano‐1H‐cyclobuta[d]pyridazine ( 1e ), and (1α,4α,4aα,5aα)‐4,4a,5,5a‐tetrahydro‐1,4,6,6‐tetramethyl‐1,4‐methano‐1H‐cyclopropa[d]pyridazine ( 1f ), as well as the corresponding housanes, the 2,3,3,4‐tetramethyl‐substituted tricyclo[3.3.0.02,4]octane ( 2c ), tricyclo[3.2.0.02,4]heptane ( 2d ), and tricyclo[3.2.0.02,4]hept‐6‐ene ( 2e ), were subjected to γ‐irradiation in Freon matrices. The reaction products were identified with the use of ESR and, in part, ENDOR spectroscopy. As expected, the strain on the C‐framework increases on going from the cyclopentane‐annelated azoalkanes and housanes ( 1c and 2c ) to those annelated by cyclobutane ( 1d and 2d ), by cyclobutene ( 1e and 2e ), and by cyclopropane ( 1f ). Accordingly, the products obtained from 1c and 2c in all three Freons used, CFCl3, CF3CCl3, and CF2ClCFCl2, were the radical cations 3c .+ and 2c .+ of 2,3,4,4‐tetramethylbicyclo[3.3.0]oct‐2‐ene and 2,3,3,4‐tetramethylbicyclo[3.3.0]octane‐2,4‐diyl, respectively. In CFCl3 and CF3CCl3 matrices, 1d and 2d yielded analogous products, namely the radical cations 3d .+ and 2d .+ of 2,3,4,4‐tetramethylbicyclo[3.2.0]hept‐2‐ene and 2,3,3,4‐tetramethylbicyclo[3.2.0]heptane‐2,4‐diyl. The radical cations 3c .+ and 3d .+ and 2c .+ and 2d .+ correspond to their non‐annelated counterparts 3a .+ and 3b .+, and 2a .+ and 2b .+ generated previously under the same conditions from 2,3‐diazabicyclo[2.2.1]hept‐2‐ene ( 1a ) and bicyclo[2.1.0]pentane ( 2a ), as well as from their 1,4‐dimethyl derivatives ( 1b and 2b ). However, in a CF2ClCFCl2 matrix, both 1d and 2d gave the radical cation 4d .+ of 2,3,3,4‐tetramethylcyclohepta‐1,4‐diene. Starting from 1e and 2e , the radical cations 4e .+ and 4e′ .+ of the isomeric 1,2,7,7‐ and 1,6,7,7‐tetramethylcyclohepta‐1,3,5‐trienes appeared as the corresponding products, while 1f was converted into the radical cation 4f .+ of 1,5,6,6‐tetramethylcyclohexa‐1,4‐diene which readily lost a proton to yield the corresponding cyclohexadienyl radical 4f .. Reaction mechanisms leading to the pertinent radical cations are discussed.  相似文献   

11.
Polythiophenes with reactive Zincke salt structure, P4ThPy+DNP(Cl?)‐a and P5ThPy+DNP(Cl?)‐a , were synthesized by the oxidation polymerization of oligothiophenes, such as 3'‐(4‐N‐(2,4‐dinitrophenyl)pyridinium chloride)?2,2':5',2'';5'',2'''‐quarterthiophene ( 4ThPy+DNP(Cl?) ) and 4''‐(4‐N‐(2,4‐dinitrophenyl)pyridinium chloride)?2,2';5',2'';5'',2''';5''',2''''‐quinquethiophene ( 5ThPy+DNP(Cl?) ), with iron(III) chloride. The reaction of P5ThPy+DNP(Cl?)‐a with R‐NH2 [R = n‐hexyl (Hex) and phenyl (Ph)] substituted the 2,4‐dinitrophenyl group into the R group with the elimination of 2,4‐dinitroaniline to yield P5ThPy+R(Cl?) . Similarly, model compounds, 4ThPy+R(Cl?) and 5ThPy+R(Cl?) (R = Hex and Ph), were also synthesized. In contrast to the photoluminescent 4ThPy and 5ThPy , the compounds P4ThPy+DNP(Cl?)‐a , P5ThPy+DNP(Cl?)‐a , and P5ThPy+R(Cl?) showed no photoluminescence because their internal pyridinium rings acted as quenchers. Cyclic voltammetry measurements suggested that P4ThPy+DNP(Cl?)‐a , P5ThPy+DNP(Cl?)‐a , and P5ThPy+R(Cl?) received an electrochemical reduction of the pyridinium and 2,4‐dinitrophenyl groups and oxidation of the polymer backbone. P4ThPy+DNP(Cl?)‐a and P5ThPy+DNP(Cl?)‐a were electrically conductive (ρ = 3.0 × 10 ? 6 S cm ? 1 and 2.1 × 10 ? 6 S cm ? 1, respectively) in the nondoped state. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 481–492  相似文献   

12.
1H‐1, 3‐Benzazaphospholes react with M(CO)5(THF) (M = Cr, Mo, W) to give thermally and relatively air stable η1‐(1H‐1, 3‐Benzazaphosphole‐P)M(CO)5 complexes. The 1H‐ and 13C‐NMR‐data are in accordance with the preservation of the phosphaaromatic π‐system of the ligand. The strong upfield 31P coordination shift, particularly of the Mo and W complexes, forms a contrast to the downfield‐shifts of phosphine‐M(CO)5 complexes and classifies benzazaphospholes as weak donor but efficient acceptor ligands. Nickelocene reacts as organometallic species with metalation of the NH‐function. The resulting ambident 1, 3‐benzazaphospholide anions prefer a μ2‐coordination of the η5‐CpNi‐fragment at phosphorus to coordination at nitrogen or a η3‐heteroallyl‐η5‐CpNi‐semisandwich structure. This is shown by characteristic NMR data and the crystal structure analysis of a η5‐CpNi‐benzazaphospholide. The latter is a P‐bridging dimer with a planar Ni2P2 ring and trans‐configuration of the two planar heterocyclic phosphido ligands arranged perpendicular to the four‐membered ring.  相似文献   

13.
The three component condensation reactions involving 2‐methylbenzothiazole or 2,5‐dimethylbenzothiazole, dialkyl acetylenedicarboxylate, and 2,6‐dimethyl phenol or 2,6‐di‐tert‐butylphenol constitute a novel and one‐pot synthesis of tetrahydropyrrolo[1,2‐a ]‐2‐methylbenzothiazoles‐3‐spiro‐1‐cyclohexa‐2,5‐dien‐4‐one‐4,5‐dicarboxylate derivatives in good yields. The reactions proceeded at room temperature without using any catalyst. This method is very useful to functionalize benzothiazole derivatives in a one‐pot operation.  相似文献   

14.
The reaction of 1,4‐dihydro‐1,4‐epoxynaphthalene with cyclooctatetraene at 130±5° for 14 days gave the four products 2a,3,3a,4,9,9a,10,10a‐octahydro‐4,9‐epoxy‐3,10‐ethenocyclobuta[b]anthracene ( 13 ), 25‐oxanonacyclo[10.10.2.25,9.114,21.02,11.03,10.04,6.013,22.015,20]heptacosa‐7,15,17,19,23,26‐hexaene ( 14 ), 5,5a,6,6a,6b,6c,12a,12b,12c,13,13a,14‐dodecahydro‐5,14‐epoxy‐6,13‐ethenocycloocta[3′,4′]cyclobuta[1′,2′:3,4]cyclobuta[1,2‐b]anthracene ( 15 ) and bis‐adduct 16 . The structures of the products were determined by spectroscopic methods. It was observed that adduct 14 undergoes a Cope rearrangement. The Cope rearrangement of this adduct was investigated in the temperature range of ?85° to 100° by NMR spectroscopy.  相似文献   

15.
The novel title organic salt, 4C5H7N2+·C24H8O84−·8H2O, was obtained from the reaction of perylene‐3,4,9,10‐tetracarboxylic acid (H4ptca) with 4‐aminopyridine (4‐ap). The asymmetric unit contains half a perylene‐3,4,9,10‐tetracarboxylate (ptca4−) anion with twofold symmetry, two 4‐aminopyridinium (4‐Hap+) cations and four water molecules. Strong N—H...O hydrogen bonds connect each ptca4− anion with four 4‐Hap+ cations to form a one‐dimensional linear chain along the [010] direction, decorated by additional 4‐Hap+ cations attached by weak N—H...O hydrogen bonds to the ptca4− anions. Intermolecular O—H...O interactions of water molecules with ptca4− and 4‐Hap+ ions complete the three‐dimensional hydrogen‐bonding network. From the viewpoint of topology, each ptca4− anion acts as a 16‐connected node by hydrogen bonding to six 4‐Hap+ cations and ten water molecules to yield a highly connected hydrogen‐bonding framework. π–π interactions between 4‐Hap+ cations, and between 4‐Hap+ cations and ptca4− anions, further stabilize the three‐dimensional hydrogen‐bonding network.  相似文献   

16.
A new class of half‐sandwich (η6p‐cymene) ruthenium(II) complexes supported by 2‐aminofluorene derivatives [Ru(η6p‐cymene)(Cl)(L)] ( L  = 2‐(((9H‐fluoren‐2‐yl)imino)methyl)phenol ( L 1 ), 2‐(((9H‐fluoren‐2‐yl)imino)methyl)‐3‐methoxyphenol ( L 2 ), 1‐(((9H‐fluoren‐2‐yl)imino)methyl)naphthalene‐2‐ol ( L 3 ) and N‐((1H‐pyrrol‐2‐yl)methylene)‐9H‐fluorene‐2‐amine ( L 4 )) were synthesized. All compounds were fully characterized by analytical and spectroscopic techniques (IR, UV–Vis, NMR) and also by mass spectrometry. The solid state molecular structures of the complexes [Ru(η6p‐cymene)(Cl)(L2)], [Ru(η6p‐cymene)(Cl)(L3)] and [Ru(η6p‐cymene)(Cl)(L4)] revealed that the 2‐aminofluorene and p‐cymene moieties coordinate to ruthenium(II) in a three‐legged piano‐stool geometry. The synthesized complexes were used as catalysts for the dehydrogenative coupling of benzyl alcohol with a range of amines (aliphatic, aromatic and heterocyclic). The reactions were carried out under thermal heating, ultrasound and microwave assistance, using solvent or solvent free conditions, and the catalytic performance was optimized regarding the solvent, the type of base, the catalyst loading and the temperature. Moderately high to very high isolated yields were obtained using [Ru(η6p‐cymene)(Cl)(L4)] at 1 mol%. In general, microwave irradiation produced better yields than the other two techniques irrespective of the nature of the substituents.  相似文献   

17.
The reaction of di(alkyn‐1‐yl)vinylsilanes R1(H2C═CH)Si(C≡C―R)2 (R1 = Me ( 1 ), Ph ( 2 ); R = Bu (a), Ph (b), Me2HSi (c)) at 25°C with 1 equiv. of 9‐borabicyclo[3.3.1]nonane (9‐BBN) affords 1‐silacyclopent‐2‐ene derivatives ( 3a , 3b , 3c , 4a , 4b ), bearing one Si―C≡C―R function readily available for further transformations. These compounds are formed by consecutive 1,2‐hydroboration followed by intramolecular 1,1‐carboboration. Treated with a further equivalent of 9‐BBN in benzene they are converted at relatively high temperature (80–100°C) into 1‐alkenyl‐1‐silacyclopent‐2‐ene derivatives ( 5a , 5b 6a , 6b ) as a result of 1,2‐hydroboration of the Si―C≡C―R function. Protodeborylation of the 9‐BBN‐substituted 1‐silacyclopent‐2‐ene derivatives 3 , 4 , 5 , 6 , using acetic acid in excess, proceeds smoothly to give the novel 1‐silacyclopent‐2‐ene ( 7 , 8 , 9 , 10 ). The solution‐state structural assignment of all new compounds, i.e. di(alkyn‐1‐yl)vinylsilanes and 1‐silacyclopent‐2‐ene derivatives, was carried out using multinuclear magnetic resonance techniques (1H, 13C, 11B, 29Si NMR). The gas phase structures of some examples were calculated and optimized by density functional theory methods (B3LYP/6‐311+G/(d,p) level of theory), and 29Si NMR parameters were calculated (chemical shifts δ29Si and coupling constants nJ(29Si,13C)). Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

18.
Palladium(II) Complexes of 1,1,3,3,5,5‐Hexakis(dimethylamino)‐λ5‐[1,3,5]triphosphinine 1,1,3,3,5,5‐Hexakis(dimethylamino)‐1λ5‐3λ5‐5λ5‐[1,3,5]triphosphinine ( 5 ) reacts with (benzonitrile)2PdCl2 to give the chelate complex dichloro(dodeca‐N‐methyl‐1λ5,3λ5,5λ5‐1,3,5‐triphosphinine‐1,1,3,3,5,5‐hexaamin‐C2,C4)palladium ( 6 ). In a pyridine‐d5 solution of 6 the complex dichloro(dodeca‐N‐methyl‐1λ5,3λ5,5λ5‐1,3,5‐triphosphinine‐1,1,3,3,5,5‐hexaamin‐C2)((2H5)pyridine‐N)palladium ( 7 ) is formed. The solute 7 could not be isolated as a solid, because elimination of the solvent regenerates 6 quantitatively. Properties, nmr and ir spectra of 6 and 7 are reported. 6 is characterized by the results of an X‐ray structural analysis.  相似文献   

19.
《化学:亚洲杂志》2017,12(9):1013-1019
Apart from sensing and imaging, luminogens with aggregation‐induced emission (AIE) are also interesting for photosensitizing. The photosensitizing behavior and bacteria‐killing performance of a pyridinium‐substituted tetraphenylethylene with an alkyne group ( TPE‐A‐Py+ ) is reported herein. Interestingly, TPE‐A‐Py+ exhibits higher photosensitizing efficiency than TPE‐Py+ (without alkyne group) when I was used as a counteranion. This is well explained by the fact that the ΔΕ ST between the excited singlet state (S 1) and triplet state (T 1) was lower for TPE‐A‐Py+ than for TPE‐Py+ , according to theoretical calculations. Moreover, replacement of I with other anions (PF6, N(SO2CF3)2 and BPh4) led to a decrease of photosensitizing efficiency for TPE‐A‐Py+ . Notably, TPE‐A‐Py+ could be used as an efficient photosensitizer to photo‐inactivate ampicillin‐resistant (ampr) E. coli at low concentration under white‐light irradiation very quickly.  相似文献   

20.
Coordination polymers are a thriving class of functional solid‐state materials and there have been noticeable efforts and progress toward designing periodic functional structures with desired geometrical attributes and chemical properties for targeted applications. Self‐assembly of metal ions and organic ligands is one of the most efficient and widely utilized methods for the construction of CPs under hydro(solvo)thermal conditions. 2‐(Pyridin‐3‐yl)‐1H‐imidazole‐4,5‐dicarboxylate (HPIDC2−) has been proven to be an excellent multidentate ligand due to its multiple deprotonation and coordination modes. Crystals of poly[aquabis[μ3‐5‐carboxy‐2‐(pyridin‐3‐yl)‐1H‐imidazole‐4‐carboxylato‐κ5N1,O5:N3,O4:N2]copper(II)dicopper(I)], [CuIICuI2(C10H5N3O4)2(H2O)]n, (I), were obtained from 2‐(pyridin‐3‐yl)‐1H‐imidazole‐4,5‐dicarboxylic acid (H3PIDC) and copper(II) chloride under hydrothermal conditions. The asymmetric unit consists of one independent CuII ion, two CuI ions, two HPIDC2− ligands and one coordinated water molecule. The CuII centre displays a square‐pyramidal geometry (CuN2O3), with two N,O‐chelating HPIDC2− ligands occupying the basal plane in a trans geometry and one O atom from a coordinated water molecule in the axial position. The CuI atoms adopt three‐coordinated Y‐shaped coordinations. In each [CuN2O] unit, deprotonated HPIDC2− acts as an N,O‐chelating ligand, and a symmetry‐equivalent HPIDC2− ligand acts as an N‐atom donor via the pyridine group. The HPIDC2− ligands in the polymer serve as T‐shaped 3‐connectors and adopt a μ3‐κ2N,O2N′,O′:κN′′‐coordination mode, linking one CuII and two CuI cations. The Cu cations are arranged in one‐dimensional –Cu1–Cu2–Cu3– chains along the [001] direction. Further crosslinking of these chains by HPIDC2− ligands along the b axis in a –Cu2–HPIDC2−–Cu3–HPIDC2−–Cu1– sequence results in a two‐dimensional polymer in the (100) plane. The resulting (2,3)‐connected net has a (123)2(12)3 topology. Powder X‐ray diffraction confirmed the phase purity for (I), and susceptibilty measurements indicated a very weak ferromagnetic behaviour. A thermogravimetric analysis shows the loss of the apical aqua ligand before decomposition of the title compound.  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