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 203 毫秒
1.
Quantum chemistry calculations predict that besides the reported single metal anion Pt, Ni can also mediate the co-conversion of CO2 and CH4 to form [CH3−M(CO2)−H] complex, followed by transformation to C−C coupling product [H3CCOO−M−H] ( A ), hydrogenation products [H3C−M−OCOH] ( B ) and [H3C−M−COOH]. For Pd, a fourth product channel leading to PdCO2…CH4 becomes more competitive. For Ni, the feed order must be CO2 first, as the weaker donor-acceptor interaction between Ni and CH4 increases the C−H activation barrier, which is reduced by [Ni−CO2]. For Ni/Pt, the highly exothermic products A and B are similarly stable with submerged barrier that favors B . The smaller barrier difference between A and B for Ni suggests the C−C coupling product is more competitive in the presence of Ni than Pt. The charge redistribution from M is the driving force for product B channel. This study adds our understanding of single atomic anions to activate CH4 and CO2 simultaneously.  相似文献   

2.
Corona[5]arenes, a novel type of macrocyclic compound that is composed of alternating heteroatoms and para-arylenes, were synthesized efficiently by two distinct methods. In a macrocycle-to-macrocycle transformation approach, S6-corona[3]arene[3]tetrazine underwent sequential SNAr reactions with HS-C6H4-X-C6H4-SH (X=S, CH2, CMe2, SO2, and O) to produce the corresponding corona[3]arene[2]tetrazines. Different corona[3]arene[2]tetrazine compounds were also constructed in a straightforward manner by a one-pot three-component reaction of HS-C6H4-X-C6H4-SH (X=S, CH2, CMe2, SO2, and O) with diethyl 2,5-dimercaptoterephthalate and 2 equiv of 3,6-dichlorotetrazine under very mild conditions. All corona[5]arenes adopted 1,2,4-alternate conformational structures in the crystalline state yielding similar nearly regular pentagonal cavities. Both the cavity size and the electronic property of the acquired macrocycles were fine-tuned by the nature of the bridging element X.  相似文献   

3.
Ascorbate (H2A) is a well-known antioxidant to protect cellular components from free radical damage and has also emerged as a pro-oxidant in cancer therapies. However, such “contradictory” mechanisms underlying H2A oxidation are not well understood. Herein, we report Fe leaching during catalytic H2A oxidation using an Fe−N−C nanozyme as a ferritin mimic and its influence on the selectivity of the oxygen reduction reaction (ORR). Owing to the heterogeneity, the Fe-Nx sites in Fe−N−C primarily catalyzed H2A oxidation and 4 e ORR via an iron-oxo intermediate. Nonetheless, trace O2 produced by marginal N−C sites through 2 e ORR accumulated and attacked Fe-Nx sites, leading to the linear leakage of unstable Fe ions up to 420 ppb when the H2A concentration increased to 2 mM. As a result, a substantial fraction (ca. 40 %) of the N−C sites on Fe−N−C were activated, and a new 2+2 e ORR path was finally enabled, along with Fenton-type H2A oxidation. Consequently, after Fe ions diffused into the bulk solution, the ORR at the N−C sites stopped at H2O2 production, which was the origin of the pro-oxidant effect of H2A.  相似文献   

4.
Corona[5]arenes, a novel type of macrocyclic compound that is composed of alternating heteroatoms and para ‐arylenes, were synthesized efficiently by two distinct methods. In a macrocycle‐to‐macrocycle transformation approach, S6‐corona[3]arene[3]tetrazine underwent sequential SNAr reactions with HS‐C6H4‐X‐C6H4‐SH (X=S, CH2, CMe2, SO2, and O) to produce the corresponding corona[3]arene[2]tetrazines. Different corona[3]arene[2]tetrazine compounds were also constructed in a straightforward manner by a one‐pot three‐component reaction of HS‐C6H4‐X‐C6H4‐SH (X=S, CH2, CMe2, SO2, and O) with diethyl 2,5‐dimercaptoterephthalate and 2 equiv of 3,6‐dichlorotetrazine under very mild conditions. All corona[5]arenes adopted 1,2,4‐alternate conformational structures in the crystalline state yielding similar nearly regular pentagonal cavities. Both the cavity size and the electronic property of the acquired macrocycles were fine‐tuned by the nature of the bridging element X.  相似文献   

5.
The anellation of a 6‐membered ring to the 2,3‐position of corannulene (=dibenzo[ghi,mno]fluoranthene; 1 ) leads to curved aromatic compounds with a significantly higher bowl‐inversion barrier than corannulene (see Fig. 1). If the bridge is −CH2−NR−CH2−, a variety of linkers can be introduced at the N(2) atom, and the corresponding curved aromatics act as versatile building blocks for larger structures (see Scheme). The locked bowl, in combination with an amide bond (see 9 and 10 ), gives rise to corannulene derivatives with chiral ground‐state conformations, which possess the ability to adapt to their chiral environment by shifting their enantiomer equilibrium slightly in favor of one enantiomeric conformer. Rim annulation of corannulene seems to display a significantly lower electron‐withdrawing effect than facial anellation on [5,6]fullerene‐C60Ih, as determined by an investigation of the basicity at the N‐atom of CH2−NR−CH2 (see 4 vs. 15 in Fig. 2).  相似文献   

6.
With the goal of generating anionic analogues to MN2S2 ⋅ Mn(CO)3Br we introduced metallodithiolate ligands, MN2S22− prepared from the Cys-X-Cys biomimetic, ema4− ligand (ema=N,N′-ethylenebis(mercaptoacetamide); M=NiII, [VIV≡O]2+ and FeIII) to Mn(CO)5Br. An unexpected, remarkably stable dimanganese product, (H2N2(CH2C=O(μ-S))2)[Mn(CO)3]2 resulted from loss of M originally residing in the N2S24− pocket, replaced by protonation at the amido nitrogens, generating H2ema2−. Accordingly, the ema ligand has switched its coordination mode from an N2S24− cavity holding a single metal, to a binucleating H2ema2− with bridging sulfurs and carboxamide oxygens within Mn-μ-S-CH2-C-O, 5-membered rings. In situ metal-templating by zinc ions gives quantitative yields of the Mn2 product. By computational studies we compared the conformations of “linear” ema4− to ema4− frozen in the “tight-loop” around single metals, and to the “looser” fold possible for H2ema2− that is the optimal arrangement for binucleation. XRD molecular structures show extensive H-bonding at the amido-nitrogen protons in the solid state.  相似文献   

7.
Crystalline {Cryptand(Na+)}[(COD)RhICl⋅SnII(Pc3−)]⋅2C6H4Cl2 ( 1 ) and {Cryptand(Cs+)}[(COD)RhI⋅SnII(Pc4−)]⋅C6H5CH3 ( 2 ) complexes were obtained via the interaction of [SnII(Pc3−)] and [SnII(Pc4−)]2−, respectively, with organometallic {(COD)RhCl}2 dimer (COD is 1,5-cyclooctadiene). Dissociation of {(COD)RhCl}2 followed by the Rh−Sn binding is observed at the formation of 1 . Elimination of the chlorine atom at the rhodium atom is observed in 2 , and rhodium is additionally coordinated to the imine nitrogen atom of Pc4−. The complexes contain mono- Pc⋅3− and doubly reduced Pc4− species, respectively, that is supported by the data of XRD analysis as well as optical and magnetic properties of 1 and 2 . There is an alternation of C-Nimine bonds in the macrocycles, which gradually increases with increasing negative charge on the macrocycle. The difference between shorter and longer bonds increases from 0.051 Å in Pc3− to 0.075 Å in Pc4−. The formation of 1 is accompanied by an essential blue shift of the Q-band of starting SnPc and the appearance of a new intense band at 1031 nm. The even stronger shift of the Q-band is observed in the spectrum of 2 , but the band in the near-IR range becomes weaker. The value of effective magnetic moment of 1 is 1.76 μB at 300 K corresponding the contribution of the Pc3− radical trianions (S=1/2). Only weak magnetic coupling with the Weise temperature of −3 K is observed in 1 due to weak π–π interaction between the macrocycles in the chains. Paramagnetic Pc3− species additionally monitored by EPR spectroscopy show a strong temperature dependence of g-factor and linewidth of the EPR signal. Complex 2 is diamagnetic and EPR silent.  相似文献   

8.
The dynamic transformations of conformations and aromatic properties of [32]octaphyrins(1.0.1.0.1.0.1.0) through rotating the pyrrolic ring of the macrocycles are demonstrated by theoretical simulations in CH2Cl2 solution. Facile multistep isomerizations involving antiaromatic-Hückel and aromatic-M?bius topologies were also predicted by density functional theory(DFT). The understanding of changes in topologies and aromaticities of free-base expanded porphrins may provide useful information to build new macrocycles with unique properties.  相似文献   

9.
《化学:亚洲杂志》2017,12(8):910-919
Reduction of aluminum(III), gallium(III), and indium(III) phthalocyanine chlorides by sodium fluorenone ketyl in the presence of tetrabutylammonium cations yielded crystalline salts of the type (Bu4N+)2[MIII(HFl−O)(Pc.3−)].−(Br) ⋅ 1.5 C6H4Cl2 [M=Al ( 1 ), Ga ( 2 ); HFl−O=fluoren‐9‐olato anion; Pc=phthalocyanine] and (Bu4N+) [InIIIBr(Pc.3−)].− ⋅ 0.875 C6H4Cl2 ⋅ 0.125 C6H14 ( 3 ). The salts were found to contain Pc.3− radical anions with negatively charged phthalocyanine macrocycles, as evidenced by the presence of intense bands of Pc.3− in the near‐IR region and a noticeable blueshift in both the Q and Soret bands of phthalocyanine. The metal(III) atoms coordinate HFl−O anions in 1 and 2 with short Al−O and Ga−O bond lengths of 1.749(2) and 1.836(6) Å, respectively. The C−O bonds [1.402(3) and 1.391(11) Å in 1 and 2 , respectively] in the HFl−O anions are longer than the same bond in the fluorenone ketyl (1.27–1.31 Å). Salts 1 – 3 show effective magnetic moments of 1.72, 1.66, and 1.79 μB at 300 K, respectively, owing to the presence of unpaired S= 1/2 spins on Pc.3−. These spins are coupled antiferromagnetically with Weiss temperatures of −22, −14, and −30 K for 1 – 3 , respectively. Coupling can occur in the corrugated two‐dimensional phthalocyanine layers of 1 and 2 with an exchange interaction of J /k B=−0.9 and −1.1 K, respectively, and in the π‐stacking {[InIIIBr(Pc.3−)].−}2 dimers of 3 with an exchange interaction of J /k B=−10.8 K. The salts show intense electron paramagnetic resonance (EPR) signals attributed to Pc.3−. It was found that increasing the size of the central metal atom strongly broadened these EPR signals.  相似文献   

10.
The GROMOS96 molecular‐dynamics (MD) program and force field was used to calculate the conformations at 298 K in CHCl3 solution of two hexakis(3‐hydroxyalkanoic acids). One consists of (R)‐3‐hydroxybutanoate (HB) residues only: H−(OCH(Me)−CH2−CO)6−OH ( 1 ). The other one carries the side chains of valine, alanine, and leucine: H−(OCH(CHMe2)CH2−CO−O−CH(Me)−CH2−CO−O−CH(CH2 CHMe2)−CH2−CO)2−OH ( 2 ), with homochiral 3‐hydroxyalkanoate (HA) moieties. In both cases, the conformational equilibria were sampled 2500 times for 25 ns. Other than clusters of arrangements with inter‐residual hydrogen bonding (between the O‐ and C‐terminal OH and COOH groups, and with chain‐bound ester carbonyl O‐atoms; Fig. 6), there are no preferred backbone conformations in which the molecules 1 and 2 spend more than 5% of the time. Specifically, neither the 21‐ nor the 31‐helical conformation of the oligoester backbone (found in stretched fibers, in lamellar crystallites, and in single crystals of polymers PHB and of oligomers OHB) is sampled to any significant extent (Fig. 8 and 9), in spite of the high population, in both oligomers, of the (−)‐synclinal conformation around the C(2)−C(3) bond (angle ϕ2 in Fig. 2). In contrast to β‐oligopeptides, for which strongly preferred secondary structures are found after a few ns, and for which the number of conformations levels off with time, the number of conformational clusters of the corresponding oligoesters found by our force‐field MD calculations increases steadily over the observation time of 25 ns (Fig. 5). Thus, the conclusion from biological and physical‐chemical studies, according to which the PHB chain is extremely flexible, is confirmed by our computational investigation: in CHCl3 solution, the hexakis(3‐hydroxyalkanoate) chain samples its conformational space randomly!  相似文献   

11.
We report a new antiferromagnetic radical-anion salt (RAS) formed from 7,7,8,8-tetracyanquinonedimethane (TCNQ) anion and 2-amino-5-chloro-pyridine cation with the composition of (N−CH3−2-NH2−5Cl−Py)(TCNQ)(CH3CN). The crystallographic data indicates the formation of (TCNQ)2.− radical-anion π-dimers in the synthesized RAS. Unrestricted density functional theory calculations show that the formed π-dimers characterize with strong π-stacking “pancake” interactions, resulting in high electronic coupling, enabling efficient charge transfer properties, but π-dimers cannot be stable in the isolated conditions as a result of strong Coulomb repulsions. In a crystal, where (TCNQ)2.− π-dimers bound in the endless chainlets via supramolecular bonds with (N−CH3−2-NH2−5-Cl−Py)+ cations, the repulsion forces are screened, allowing for specific parallel π-stacking interactions and stable radical-anion dimers formation. Measurements of magnetic susceptibility and magnetization confirm antiferromagnetic properties of RAS, what is in line with the higher stability of ground singlet state of the radical-anion pair, calculated by means of the DFT. Therefore, the reported radical-anion (N−CH3−2-NH2−5Cl−Py)(TCNQ)(CH3CN) solvate has promising applications in novel magnetics with supramolecular structures.  相似文献   

12.
Water is the most sustainable source for H2 production, and the efficient electrocatalytic production of H2 from mixed water/acetonitrile solutions by using two new air-stable nickel(II) pincer complexes, [Ni(κ3-2,6-{Ph2PNR}2(NC5H3)Br2] (R=H I , Me II ) is reported. Hydrogen generation from H2O/CH3CN solutions is initiated at −2 V against Fc+/0, and bulk electrocatalysis studies showed that the catalyst functions with an excellent Faradaic efficiency and a turnover frequency of 160 s−1. A DFT computational investigation of the reduction behavior of I and II revealed a correlation of H2 formation with charge donation from electrons originating in a reduced ligand-localized orbital. As a result, these catalysts are proposed to proceed by a novel mechanism involving electron/proton transfer between a Ni0I species bonded to an anionic PN3P ligand (“L/Ni0I”) and a NiI-hydride (“Ni−H”). Furthermore, these catalysts are able to reduce phenol and acetic acid, more active proton sources, at lower potentials that correlate with the substrate pKa.  相似文献   

13.
Lower-rim mono- and diacylated calix[4]arenes [acyl = C6H5CO, 3,5-(NO2)2C6H3CO] undergo selective adamantylation with 3-Y-1-adamantanols (Y = H, i-Pr, 4-MeC6H4) in trifluoroacetic acid at the free phenolic fragments of the macroring. The reaction provides a convenient preparative route to di-, tri-, and tetraadamantylated calix[4]arenes.  相似文献   

14.
The current library of amidinate ligands has been extended by the synthesis of two novel dimethylamino-substituted alkynylamidinate anions of the composition [Me2N−CH2−C≡C−C(NR)2] (R = iPr, cyclohexyl (Cy)). The unsolvated lithium derivatives Li[Me2N−CH2−C≡C−C(NR)2] ( 1 : R = iPr, 2 : R = Cy) were obtained in good yields by treatment of in situ-prepared Me2N−CH2−C≡C−Li with the respective carbodiimides, R−N=C=N−R. Recrystallization of 1 and 2 from THF afforded the crystalline THF adducts Li[Me2N−CH2−C≡C−C(NR)2] ⋅ nTHF ( 1 a : R = iPr, n=1; 2 a : R = Cy, n=1.5). Precursor 2 was subsequently used to study initial complexation reactions with selected di- and trivalent transition metals. The dark red homoleptic vanadium(III) tris(alkynylamidinate) complex V[Me2N−CH2−C≡C−C(NCy)2]3 ( 3 ) was prepared by reaction of VCl3(THF)3 with 3 equiv. of 2 (75 % yield). A salt-metathesis reaction of 2 with anhydrous FeCl2 in a molar ratio of 2 : 1 afforded the dinuclear homoleptic iron(II) alkynylamidinate complex Fe2[Me2N−CH2−C≡C−C(NCy)2]4 ( 4 ) in 69 % isolated yield. Similarly, treatment of Mo2(OAc)4 with 3 or 4 equiv. of 2 provided the dinuclear, heteroleptic molybdenum(II) amidinate complex Mo2(OAc)[Me2N−CH2−C≡C−C(NCy)2]3 ( 5 ; yellow crystals, 50 % isolated yield). The cyclohexyl-substituted title compounds 2 a , 4 , and 5 were structurally characterized through single-crystal X-ray diffraction studies.  相似文献   

15.
In this study we compare the binding energies of polycoordinated complexes of Zn2+ within cavities composed of model “hard” (H2O, OH) or “soft” (CH3SH, CH3S) ligands. Ab initio supermolecule computations are performed at the HF and MP2 levels using extended basis sets to determine the binding energies and their components as a function of: the number of ligands, ranging from three to six; the net charge of the cavity; and the “hard” versus “soft” character of the ligands. These ab initio computations are used to test the reliability of the SIBFA molecular mechanics procedure, originally formulated and calibrated on the basis of ab initio computations, for such charged systems. The SIBFA intermolecular interaction energies match the corresponding ab initio values using a coreless effective potential split‐valence basis set with a relative error of ≤3%. Extensions to binuclear Zn2+ complexes, such as those that occur in the Zn‐binding sites of Gal4 and β‐lactamase proteins, are performed to test the applicability of the methodology for such systems. © 2000 John Wiley & Sons, Inc. J Comput Chem 21: 1011–1039, 2000  相似文献   

16.
The purpose of this article was to calculate the structures and energetics of CH3O(H2O)n and CH3S(H2O)n in the gas phase; the maximum number of water molecules that can directly interact with the O of CH3O; and when n is larger, we asked how the CH3O and CH3S moiety of CH3O(H2O)n and CH3S(H2O)n changes and how we can reproduce experimental ΔH 0n−1, n. Using the ab initio closed-shell self-consistent field method with the energy gradient technique, we carried out full geometry optimizations with the MP2/aug-cc-pVDZ for CH3O(H2O)n (n=0, 1, 2, 3) and the MP2/6–31+G(d,p) (for n=5, 6). The structures of CH3S(H2O)n (n=0, 1, 2, 3) were fully optimized using MP2/6–31++G(2d,2p). It is predicted that the CH3O(H2O)6 does not exist. We also performed vibrational analysis for all clusters [except CH3O(H2O)6] at the optimized structures to confirm that all vibrational frequencies are real. Those clusters have all real vibrational frequencies and correspond to equilibrium structures. The results show that the above maximum number of water molecules for CH3O is five in the gas phase. For CH3O(H2O)n, when n becomes larger, the C—O bond length becomes longer, the C—H bond lengths become smaller, the HCO bond angles become smaller, the charge on the hydrogen of CH3 becomes more positive, and these values of CH3O(H2O)n approach the corresponding values of CH3OH with the n increment. The C—O bond length of CH3O(H2O)3 is longer than the C—O bond length of CH3O in the gas phase by 0.044 Å at the MP2/aug-cc-pVDZ level of theory. The structure of the CH3S moiety in CH3S(H2O)n does not change with the n increment. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 1138–1144, 1999  相似文献   

17.
Photo-driven CH4 conversion to multi-carbon products and H2 is attractive but challenging, and the development of efficient catalytic systems is critical. Herein, we construct a solar-energy-driven redox cycle for combining CH4 conversion and H2 production using iron ions. A photo-driven iron-induced reaction system was developed, which is efficient at selective coupling of CH4 as well as conversion of benzene and cyclohexane under mild conditions. For CH4 conversion, 94 % C2 selectivity and a C2H6 formation rate of 8.4 μmol h−1 is achieved. Mechanistic studies reveal that CH4 coupling is induced by hydroxyl radical, which is generated by photo-driven intermolecular charge migration of an Fe3+ complex. The delicate coordination structure of the [Fe(H2O)5OH]2+ complex ensures selective C−H bond activation and C−C coupling of CH4. The produced Fe2+ can be used to reduce the potential for electrolytic H2 production, and then turns back into Fe3+, forming an energy-saving and sustainable recyclable system.  相似文献   

18.
2‐Methylideneglutarate mutase is an adenosylcobalamin (coenzyme B12)‐dependent enzyme that catalyses the equilibration of 2‐methylideneglutarate with (R)‐3‐methylitaconate. This reaction is believed to occur via protein‐bound free radicals derived from substrate and product. The stereochemistry of the formation of the methyl group of 3‐methylitaconate has been probed using a `chiral methyl group'. The methyl group in 3‐([2H1,3H]methyl)itaconate derived from either (R)‐ or (S)‐2‐methylidene[3‐2H1,3‐3H1]glutarate was a 50 : 50 mixture of (R)‐ and (S)‐forms. It is concluded that the barrier to rotation about the C−C bond between the methylene radical centre and adjacent C‐atom in the product‐related radical [.CH2CH(O2CC=CH2)CO2] is relatively low, and that the interaction of the radical with cob(II)alamin is minimal. Hence, cob(II)alamin is a spectator of the molecular rearrangement of the substrate radical to product radical.  相似文献   

19.
In achiral rod‐like molecules, a nematic phase is the most disordered liquid crystal phase, which only has one‐directional order in the direction of the molecular long axis. A dumbbell‐shaped molecule (compound 3 : R−C6H10−CH=CH−C6H4−CH=CH−C6H10−R, (R=n C5H11)), and its liquid crystal phase (X phase) are reported, which exhibit high scattering without thermal fluctuation between two nematic phases under a polarized light optical microscope. The X phase was investigated by X‐ray diffraction, scanning electron microscopy, atomic force microscopy, and molecular dynamics simulation. A layered structure was ascertained for which a molecular self‐organization mechanism was postulated in which the super‐structure is based on lateral intermolecular interlocking. A second nematic phase above the X phase consisted of “rice grain”‐shaped particles.  相似文献   

20.
Four silver thiolate clusters, [H3O][(Ag3S3)(BF4)@Ag27(tBuS)18(hfac)6H2O] ⋅ H2O ( 1 ; hfac = hexafluoroacetylacetone), [(Ag3S3)(CF3CO2)@Ag30(tBuS)16(CF3CO2)9(CH3CN)4] ⋅ CF3CO2 ⋅ 4 CH3CN ( 2 ), [(Ag3S3)(MoO4)@Ag30(tBuS)16(CF3CO2)9(CH3CN)4] ⋅ 2 CH3CN ( 3 ), and [(Ag3S3)(CrO4)@Ag30(tBuS)16(CF3CO2)9(CH3CN)4] ⋅ 4 CH3CN ( 4 ), were isolated. They have similar nestlike structures assembled by an [Ag3S3]3− template together with one of the BF4, CF3CO2, MoO42−, or CrO42− anions. Interestingly, the solid-state emissions of 2 – 4 are dependent on the templating anions and are tunable from green to orange and then to red by changing the template from CF3CO2 to MoO42− and to CrO42−, and this may be correlated to the charge transfer between these templates to metal atoms. This work helps to understand the templating role of heteroanions and the relationship between structure and properties.  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