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 359 毫秒
1.
We have employed group theory and picosecond time-resolved fluorescence isotropy and anisotropy spectroscopy methods to explore the excitation transfers within an isolated C-phycocyanin (C-PC) hexamer (αβ)6PCL27RC, situated at the end of the rod proximal to the core of the pycobilisome (PBS) in the cyanobacterium Anabaena variabilis. The group-theory results imply that excitation energy transfer between two trimers occurs between the lowest exciton level of each trimer. The excitation energy-transfer process might occur at a rate of 10–20 ps, and it may be described by an exciton hopping-like Förster transfer mechanism. Dynamic components of 45–50 ps are assigned to the excitation transfer from β155-PCB chromophores to the exciton states of dimers, which consist of two neighbouring monomers of the same trimer in an isolated C-PC hexamer.  相似文献   

2.
Phototransformations of autofluorescent proteins are applied in high‐resolution microscopy and in studying cellular transport, but they are detrimental when accidentally occurring in blinking or photobleaching (BL). Here, we investigate the kinetics of phototransformations of a photoactivatable green fluorescent protein (GFP) in confocal microscopy. Photoconversion (PC) is achieved by excitation of the barely present anionic chromophore state Req? in the GFP mutant Thr203Val. Besides the shift of the equilibrium between the neutral chromophore state RH and Req?, the photoconverted anionic chromophore RPC? exhibits a reduced fluorescence lifetime τfl=2.2 ns. In fluorescence lifetime imaging microscopy, τfl is found to depend, however, on the excitation conditions and history. The underlying photochemistry is described by the kinetic scheme of consecutive reactions, Req?→RPC?→Pdark, in which the anionic chromophore species and the dark protein Pdark are coupled by PC and BL. Time‐correlated single‐photon‐counting detection in a confocal geometry of freely diffusing species is used to compute the quantum yields for PC and BL, ΦPC and ΦBL. The assessed values are ΦPC=5.5×10?4 and ΦBL>1×10?5. Based on these values, PC provokes misinterpretation in fluorescence resonance energy transfer experiments and is responsible for spectroscopic peculiarities in single‐molecule detection.  相似文献   

3.
Cyclic voltammetric studies of clusters (C5H5-C2C6 H4-R-p)Co2(CO)6-n Ln[n=0,2; L=PPh3, P(OEt)3] and (RCH2C)2Co2(CO4) (PPh3)2 on Pt electrode are described. The primary reduction (0 / ?1) and oxidation (+ 1 / 0) steps are considered as a mono-electron process for all clusters. For the clusters (C5H5C2C6H4-R-p)Co2(CO)6, a good linear relation between reduction potential Epred and Hammett constant σp of R in the clusters is found. For the clusters (RC2R')Co2(CO)4L2, their radical anions are extremely unstable at room temperature and fragment into a series of mononuclear species, one of which is (RC2R')Co(CO)2PPh3. The reaction of radical anions of (RC2R')Co2(CO)6–n (PPh3)n(n=0,2) with PPh3 also produces mononuclear species (RC2R')Co(CO)2PPh3 which has been detected by means of cyclic voltammetry and ESR. The influence of R on redox properties of clusters is discussed.  相似文献   

4.
In this study, cyclic poly(3‐hexylthiophene‐2,5‐diyl) (c‐P3HT) with a controlled Mn was synthesized by the intramolecular cyclization of α‐bromo‐ω‐ethynyl‐functionalized P3HT via the Sonogashira coupling reaction. The effect of the cyclic structure, which does not have terminal groups of polymers, on the photoelectric conversion characteristics was investigated in comparison to linear P3HT (l‐P3HT). c‐P3HT was successfully synthesized with Mn ≈ 17,000, dispersity ≈ 1.2, and regioregularity ≈ 99%. The hole mobility was determined to be 5.1 × 10?4 cm2 V?1 s?1 by time‐of‐flight (TOF) experiment. This was comparable to that of l‐P3HT of 5.6 × 10?4 cm2 V?1 s?1. Organic solar cell systems were fabricated with each polymer by blending them with [6,6]‐phenyl‐C71‐butyric acid methyl ester (PC71BM). The l‐P3HT:PC71BM system showed a dispersive TOF photocurrent profile for electron transport, whereas a nondispersive profile was observed for c‐P3HT:PC71BM. In addition, an amount of collected electrons in c‐P3HT:PC71BM was greater than that in l‐P3HT:PC71BM for TOF experiments. The photoelectric conversion characteristics were improved by using c‐P3HT rather than l‐P3HT (power conversion efficiency [PCE] = 4.05% vs 3.23%), reflecting the nondispersive transport and the improvement of electron collection. PCEs will be much improved by applying this cyclic concept to highly‐efficient OSC polymers. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2019 , 57, 266–271  相似文献   

5.
Thermal trasfomations of vinylcyclopropane (VCP) radical cations (RC) in X-ray irradiated frozen Freon matrices, CFCl2CF2Cl and CFCl3, were studed by ESR. Radical processes involving VCP.+ in very rarefied and moderately thickened gaseous VCP were simulated. Monomolecular cleavage of the cyclopropane ring ofgauche-VCP.+ (1) occurs to give the more thermally stable distonic radical cationdist(0.90)-C5H8 .+ (3). As the density of VCP increases RC3 adds at the double bond ofanti-VCP to give the distonic RC,.CH2CH2CHCH(CH2)3CHCHCH2 + (5). Under the same conditions, the less thermally stableanti-VCP.+(2) undergoes monomolecular isomerization into RC1 or reacts withanti-VCP with the rearrangement (as in the condensed phase) to give its distonic form,dist(90.0)-C5H8 .+ (4). The MNDO-UHF method was adapted for quantum-chemical analysis of the constants of isotropic hyperfine coupling with1H and13C nuclei in neutral and charged hydrocabon radicals, since the standard version of this method inadequately reproduces the structural parameters of low-symmetry (C 1,C s) paramagnetic species. A quantum-chemical analysis of the radiospectroscopic information and of the stereoelectronic control of thermal transformations of conformers of RC1 and2 into their structurally nonequivalent distonic forms3 and4, respectively, was carried out.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 212–235, February, 1995.This work was carried out with the financial support of the Russian Foundation for Basic Research (Project No. 93-03-04075).  相似文献   

6.
Mass spectra of substituted benchrotrenyls RC6H5Cr(CO)3 where R?H, F, CI, I, CH3, OCH3, COOCH3, C2H5, N(CH3)2, NH2, C6H5, C(CH3)3, p-C6H4NH2, CH2C6H5, CH2CH2C6H5), 1,3,5-(CH3)3C6H3Cr(CO)3 and 1,2,3,5-(CH3)4C6H2Cr(CO)3 have been studied. It has been found that for monosubstituted benchrotrenyls there is a linear dependence of the parameter log [Cr]+/[RC6H5Cr]+) on the number of degrees of freedom of the [RC6H5Cr]+ ion. Decarbonylation of the molecular ions is not affected by the nature of the substituent R. The results are interpreted in terms of the quasi-equilibrium theory of mass spectra.  相似文献   

7.
Three novel norcantharidin acylamide acids (L1?N‐thiadiazole norcantharidin acylamide acid, C10H11N3O4S; L2?N‐thiazole norcantharidin acylamide acid, C11H12N2O4S and L3?N‐benzothiazole norcantharidin acylamide acid, C15H14N2O4S) were synthesized by the reactions of norcantharidin (NCTD?7‐oxabicyclo[2,2,1]heptane‐2,3‐dicarboxylic acid anhydride, C8H8O4) with 2‐amino‐1,3,4‐thiadiazole (C2H3N3S), 2‐aminothiazole (C3H4N2S) and 2‐aminobenzothiazole (C7H6N2S), respectively. Their structures were characterized by elemental analysis, IR, and NMR. The inhibition rates of L3 was much higher than those of L1 and L2 against human hepatoma cells SMMC7721 cell lines in vitro. The interaction between the compounds and DNA was studied by means of fluorescence quenching studies and viscosity measurements. The emission intensities decreased obviously with increasing concentration of the compounds in the fluorescence quenching experiments. The linear Stern‐Volmer quenching constant Ksq values were 0.62 (L1), 0.55 (L2) and 1.08 (L3), respectively. The binding abilities followed the trend from high to low were L3, L1 and L2, respectively. The results of viscosity measurements showed that L1 and L2 might bind to DNA via partial intercalation, while L3 bound mainly in intercalation.  相似文献   

8.
In this article, a method for quantitative determination of phytochelatins (PC n being the classic example) and other thiol-containing compounds in mixed standard solution and plant tissues is presented. Thiols were converted to fluorescent derivatives by precolumn derivatization with monobromobimane. The results showed that PC n and other thiol-containing compounds in standard mixed solutions were rapidly separated within 15 min by using a ACN 0.1% trifluoroacetic acid binary gradient elution. Glutathione was representatively selected to test the precision of this method. The calibration curve was linear in the range of 1.25–160 ng μl−1 (regression coefficient r 2=0.9999). It was confirmed that this method was rapid, simple, highly sensitive, stable, and had the property of simultaneous determination of PC n and other thiol-containing compounds. This method was applied to determine PC n and other thiol-containing compounds in a Cd hyperaccumulator Sedum alfredii in response to Cd. It was found that no PC n was detected in any tissue at any Cd treatment, suggesting that Cd hyperaccumulation and detoxification in this plant is not based on PC synthesis. Translated from Journal of Nanjing University (Natural Science Edition), 2005, 41(3) (in Chinese)  相似文献   

9.
The features in phosphorus heterocumulene yilides Ph3PCCX [X = NPh, NC(O)Ph, C(CN)CO · (OMe), O, S] with the ylide carbon geometry close to sp-hybridization are studied using ab initio methods. Estimation of structural parameters of a new member in the Ph3PCα=Cβ=NC(O)Ph series on the B3LYP/6-31G(d,p) level and together with experimental data for related ylide (X = NPh) gave the following values: r(PCα) 1.665 Å, r(Cα=Cβ) 1.237 Å, ω(PCαCβ) 145.2°. Nonvalent interactions of sp λ orbital of the ylide carbon atom bearing extra negative charge with coplanar π-electrons of the nearest Cβ=N bond are found close to those of nitrogen lone pair with the Cα=Cβ bond and leading to up to 10° nonlinearity of Cα-Cβ=N triad of the phosphorus iminoketene ylides. Chemical nonequivalence of the PCipso bonds in the triphenylphosphonium fragment, relation of \(^1 J_{PC^\alpha } \) spin-spin coupling and antisymmetric bonding vibration νas(C=C=X) to the structure of the carbanion, and inductive hyperconjugative influence of atom (or group) X on the ylide carbon geometry are discussed.  相似文献   

10.
The effective catalytic activity of organoaluminum compounds for the monohydroboration of carbodiimides has been demonstrated. Two aluminum complexes, 2 and 3 , were synthesized and characterized. The efficient catalytic performances of four aluminum hydride complexes L1AlH2 (L1=HC(CMeNAr)2, Ar=2,6-Et2C6H3; 1 ), L2AlH2(NMe3) (L2=o-C6H4F(CH=N-Ar), Ar=2,6-Et2C6H3; 2 ), L3AlH (L3=2,6-bis(1-methylethyl)-N-(2-pyridinylmethylene)phenylamine; 3 ), and L4AlH(NMe3) (L4=o-C6H4(N-Dipp)(CH=N-Dipp), Dipp=2,6-iPr2C6H3; 4 ), and an aluminum alkyl complex L1AlMe2 ( 5 ) were used for the monohydroboration of carbodiimides investigated under solvent-free and mild conditions. Compounds 1 – 3 and 5 can produce monohydroborated N-borylformamidine, whereas 4 can afford the C-borylformamidine product. A suggested mechanism of this reaction was explored, and the aluminum formamidinate compound 6 was characterized by single-crystal X-ray, also a stoichiometric reaction was investigated.  相似文献   

11.
Phytochelatins (PC) were described earlier to play a role in metal detoxification in Chlamydomonas reinhardtii but were not clearly identified. The focus of this case study was to identify PC synthesized by C. reinhardtii exposed to Cd. Only low intracellular concentrations of cadmium (85 nmol g−1 fresh weight) were sufficient to cause significant changes in thiol peptide pools. Thus, results showed a progressive decline of the glutathione content, accompanied by an induction of phytochelatins. Not only canonic phytochelatins but for the first time also the iso-phytochelatins CysPC n and PC2Ala were identified in this unicellular green alga using electrospray ionization quadrupole time-of-flight tandem mass spectrometry. Additionally, CysPC n desGly, PC n desGly, CysPC n Glu, and PC2Glu were found throughout MS analysis. Also, low abundant PCs could be detected due to the high sample preconcentration combined with little sample amounts (0.3 μL min−1) necessary for electrospray. Identified PCs had a maximum number of 5 γ-glutamyl cysteine (γ-GluCys) units. Thiol peptides of higher molecular masses suggesting PC n with n > 5 could be identified as intermolecular oxidation products of smaller PCs. Thiols may easily be oxidized. Therefore, PCs were reduced prior to MS analysis. Dithiothreitol and tris(2-carboxyethyl) phosphine were compared concerning their reduction effort.   相似文献   

12.
The instability of metal and metalloid complexes during analytical processes has always been an issue of an uncertainty regarding their speciation in plant extracts. Two different speciation protocols were compared regarding the analysis of arsenic phytochelatin (AsIIIPC) complexes in fresh plant material. As the final step for separation/detection both methods used RP-HPLC simultaneously coupled to ICP-MS and ES-MS. However, one method was the often used off-line approach using two-dimensional separation, i.e. a pre-cleaning step using size-exclusion chromatography with subsequent fraction collection and freeze-drying prior to the analysis using RP-HPLC–ICP-MS and/or ES-MS. This approach revealed that less than 2% of the total arsenic was bound to peptides such as phytochelatins in the root extract of an arsenate exposed Thunbergia alata, whereas the direct on-line method showed that 83% of arsenic was bound to peptides, mainly as AsIIIPC3 and (GS)AsIIIPC2. Key analytical factors were identified which destabilise the AsIIIPCs. The low pH of the mobile phase (0.1% formic acid) using RP-HPLC–ICP-MS/ES-MS stabilises the arsenic peptide complexes in the plant extract as well as the free peptide concentration, as shown by the kinetic disintegration study of the model compound AsIII(GS)3 at pH 2.2 and 3.8. But only short half-lives of only a few hours were determined for the arsenic glutathione complex. Although AsIIIPC3 showed a ten times higher half-life (23 h) in a plant extract, the pre-cleaning step with subsequent fractionation in a mobile phase of pH 5.6 contributes to the destabilisation of the arsenic peptides in the off-line method. Furthermore, it was found that during a freeze-drying process more than 90% of an AsIIIPC3 complex and smaller free peptides such as PC2 and PC3 can be lost. Although the two-dimensional off-line method has been used successfully for other metal complexes, it is concluded here that the fractionation and the subsequent freeze-drying were responsible for the loss of arsenic phytochelatin complexes during the analysis. Hence, the on-line HPLC–ICP-MS/ES-MS is the preferred method for such unstable peptide complexes. Since freeze-drying has been found to be undesirable for sample storage other methods for sample handling needed to be investigated. Hence, the storage of the fresh plant at low temperature was tested. We can report for the first time a storage method which successfully conserves the integrity of the labile arsenic phytochelatin complexes: quantitative recovery of AsIIIPC3 in a formic acid extract of a Thunbergia alata exposed for 24 h to 1 mg Asv L−1 was found when the fresh plant was stored for 21 days at 193 K. Figure On-line HPLC–ICP-MS/ES-MS (bottom) is the preferred method for MS determination of unstable arsenic peptide complexes in plant extracts, since this avoids fractionation and subsequent freeze-drying that are responsible for loss of arsenic phytochelatin complexes in the 2D off-line method (top) Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

13.
The competitive binding of Cu2+, Cd2+ and Pb2+ in ternary mixtures by the phytochelatins PC2 and PC5 (PCn; (γGlu‐Cys)n‐Gly, n=2 and 5) are examined by voltammetry, which allows one to follow the displacements of the voltammetric signals induced by the competitive binding among the metal ions towards PC2 or PC5 complexation, and direct injection positive‐mode Electrospray Ionization Mass Spectrometry (ESI‐MS), in order to obtain the stoichiometries of the complexes. Voltammetric data are analyzed by Gaussian Peak Adjustment (GPA), which is a recently developed multivariate analysis method for nonlinear electrochemical data. Different complexes have been detected or deduced their presence, depending on the experimental conditions. Ternary complexes CuCdPCn and CdPbPCn were detected for both PC2 and PC5, while the ternary CuPbPCn complex was only detected for PC2. Some of the complexes have been only detected by ESI‐MS because in some cases voltammetry data could not be totally resolved, even by using GPA. The quaternary CdPbCuPCn complex has been detected for PC5, but for PC2 data are not so conclusive. In summary, the signal evolution for mixed CdCuPCn complexes is quite different. These observations could be a reflection of an antagonistic effect for the case of PC2 and a synergetic one for PC5.  相似文献   

14.
[Pd(cod)(cotl)]ClO4 (cod = 1,5-cyclooctadiene, cotl = cyclooctenyl, C8H13 ) undergoes substitutions with multidentate N-heterocycles: 1,3-bis(benzimidazolyl)benzene (L1), 1,3-bis(1-methylbenzimidazol-2-yl)benzene (L2), 2,6-bis(benzimidazolyl)pyridine (L3) and 2,6-bis(1-methylbenzimidazol-2-yl)pyridine (L4) to yield mono/binuclear complexes: [Pd(cotl)(L1)(OClO3)], [Pd(cotl)(L)]ClO4 (L = L2 or L3) and [Pd(cotl)2(L4)](ClO4)2. Dihalobridged binuclear complexes [PdX(cotl)]2 (X = Cl or Br) undergo halogen bridge cleavages with the multidentate N-heterocycles to form binuclear complexes of the type [PdX(cotl)2L] (X = Cl or Br; L = L1, L2, L3 or L4). The complexes were characterized by elemental analyses, 1H-, 13C-n.m.r., i.r., far-i.r. and FAB-mass spectral studies.  相似文献   

15.
Complexes of nickel(II) aryl carboxylates with a general formula Ni(RC6H4COO)2L2 where R=H, p-CH3. p-Cl, m- & p-NO2; and L = morpholine and piperidine; have been prepared by the interaction of nickel(II) aryl carboxylates with a large excess of appropriate amine. Unlike parent anhydrous nickel(II) aryl carboxylates all these complexes are soluble in common organic solvents.  相似文献   

16.
We report here the synthesis of new C,N‐chelated chlorostannylenes and germylenes L3MCl (M=Sn( 1 ), Ge ( 2 )) and L4MCl (M=Sn( 3 ), Ge ( 4 )) containing sterically demanding C,N‐chelating ligands L3, 4 (L3=[2,4‐di‐tBu‐6‐(Et2NCH2)C6H2]?; L4=[2,4‐di‐tBu‐6‐{(C6H3‐2′,6′‐iPr2)N=CH}C6H2]?). Reductions of 1 – 4 yielded three‐coordinate C,N‐chelated distannynes and digermynes [L3, 4M ]2 for the first time ( 5 : L3, M=Sn, 6 : L3, M=Ge, 7 : L4, M=Sn, 8 : L4, M=Ge). For comparison, the four‐coordinate distannyne [L5Sn]2 ( 10 ) stabilized by N,C,N‐chelate L5 (L5=[2,6‐{(C6H3‐2′,6′‐Me2)N?CH}2C6H3]?) was prepared by the reduction of chlorostannylene L5SnCl ( 9 ). Hence, we highlight the role of donor‐driven stabilization of tetrynes. Compounds 1 – 10 were characterized by means of elemental analysis, NMR spectroscopy, and in the case of 1 , 2 , 5 – 7 , and 10 , also by single‐crystal X‐ray diffraction analysis. The bonding situation in either three‐ or four‐coordinate distannynes 5 , 7 , and 10 was evaluated by DFT calculations. DFT calculations were also used to compare the nature of the metal–metal bond in three‐coordinate C,N‐chelating distannyne [L3Sn]2 ( 5 ) and related digermyme [L3Ge]2 ( 6 ).  相似文献   

17.
Plasticised membranes using 2-[{(2-hydroxyphenyl)imino}methyl]-phenol (L1) and 2-[{(3-hydroxyphenyl)imino}methyl]-phenol (L2), have been prepared and investigated as Cu2+ ion-selective sensors. Effect of various plasticisers, namely, dibutyl phthalate (DBP), dibutyl sebacate (DBS), benzyl acetate (BA), o-nitrophenyloctylether (o-NPOE) and anion excluders, oleic acid (OA) and sodium tetraphenylborate (NaTPB) was studied and improved performance was observed in several instances. Optimum performance was observed with membranes of (L1) having composition L1 : DBS : OA : PVC in the ratio of 6 : 54 : 10 : 30 (w/w, %). The sensor works satisfactorily in the concentration range 3.2 × 10?8–1.0 × 10?1 mol L?1 with a Nernstian slope of 29.5 ± 0.5 mV decade?1 of a cu2+ . The detection limit of the proposed sensor is 2.0 × 10?8 mol L?1 (1.27 ng mL?1). Wide pH range (3.0–8.5), fast response time (7 s), sufficient (up to 25% v/v) non-aqueous tolerance and adequate shelf life (3 months) indicate the utility of the proposed sensor. The potentiometric selectivity coefficients as determined by matched potential method indicate selective response for Cu2+ ions over various interfering ions, and therefore could be successfully used for the determination of copper in edible oils, tomato plant material and river water.  相似文献   

18.
The complex Mo(CO)2L2 [L = S2CNEt2] reacts with acetylenes to yield both Mo(CO)(RC2R')L2 and Mo(RC2R')2L2, with diazenes giving Mo(RN2R')L2 and Mo(RC2R')2L2, with diazenes giving Mo(RN2R')L2 and Mo(RN2R')2L2, and with CO and PPh3 to form Mo(CO)3L2 and Mo(CO)2(PPh3)L2.  相似文献   

19.
The title compund, [Cu2(OH)2(C22H25N3)2](ClO4)2, is a copper(II) dimer, with two [CuL]2+ units [L is bis(6‐methyl‐2‐pyridylmethyl)(2‐phenylethyl)amine] bridged by hydroxide groups to define the {[CuL](μ‐OH)2[CuL]}2+ cation. Charge balance is provided by perchlorate counter‐anions. The cation has a crystallographic inversion centre halfway between the CuII ions, which are separated by 3.0161 (8) Å. The central core of the cation is an almost regular Cu2O2 parallelogram of sides 1.931 (2) and 1.935 (2) Å, with a Cu—O—Cu angle of 102.55 (11)°. The coordination geometry around each CuII centre can be best described as a square‐based pyramid, with three N atoms from L ligands and two hydroxide O atoms completing the coordination environment. Each cationic unit is hydrogen bonded to two perchlorate anions by means of hydroxide–perchlorate O—H...O interactions.  相似文献   

20.
5‐[(Imidazol‐1‐yl)methyl]benzene‐1,3‐dicarboxylic acid (H2L) was synthesized and the dimethylformamide‐ and dimethylacetamide‐solvated structures of its adducts with CuII, namely catena‐poly[[copper(II)‐bis[μ‐3‐carboxy‐5‐[(imidazol‐1‐yl)methyl]benzoato]] dimethylformamide disolvate], {[Cu(C12H9N2O4)2]·2C3H7NO}n, (I), and catena‐poly[[copper(II)‐bis[μ‐3‐carboxy‐5‐[(imidazol‐1‐yl)methyl]benzoato]] dimethylacetamide disolvate], {[Cu(C12H9N2O4)2]·2C4H9NO}n, (II), the formation of which are associated with mono‐deprotonation of H2L. The two structures are isomorphous and isometric. They consist of one‐dimensional coordination polymers of the organic ligand with CuII in a 2:1 ratio, [Cu(μ‐HL)2]n, crystallizing as the dimethylformamide (DMF) or dimethylacetamide (DMA) disolvates. The CuII cations are characterized by a coordination number of six, being located on centres of crystallographic inversion. In the polymeric chains, each CuII cation is linked to four neighbouring HL ligands, and the organic ligand is coordinated via Cu—O and Cu—N bonds to two CuII cations. In the corresponding crystal structures of (I) and (II), the coordination chains, aligned parallel to the c axis, are further interlinked by strong hydrogen bonds between the noncoordinated carboxy groups in one array and the coordinated carboxylate groups of neighbouring chains. Molecules of DMF and DMA (disordered) are accommodated at the interface between adjacent polymeric assemblies. This report provides the first structural evidence for the formation of coordination polymers with H2Lvia multiple metal–ligand bonds through both carboxylate and imidazole groups.  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