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.
Both hydrogen (H2) and copper ions (Cu+) can be used as anti-cancer treatments. However, the continuous generation of H2 molecules and Cu+ in specific sites of tumors is challenging. Here we anchored Cu2+ on carbon photocatalyst (Cu@CDCN) to allow the continuous generation of H2 and hydrogen peroxide (H2O2) in tumors using the two-electron process of visible water splitting. The photocatalytic process also generated redox-active Cu-carbon centers. Meanwhile, the Cu2+ residues reacted with H2O2 (the obstacle to the photocatalytic process) to accelerate the two-electron process of water splitting and cuprous ion (Cu+) generation, in which the Cu2+ residue promoted a pro-oxidant effect with glutathione through metal-reducing actions. Both H2 and Cu+ induced mitochondrial dysfunction and intracellular redox homeostasis destruction, which enabled hydrogen therapy and cuproptosis to inhibit cancer cell growth and suppress tumor growth. Our research is the first attempt to integrate hydrogen therapy and cuproptosis using metal-enhanced visible solar water splitting in nanomedicine, which may provide a safe and effective cancer treatment.  相似文献   

2.
A series of Cu+ complexes with ligands that feature varying numbers of benzimidazole/thioether donors and methylene or ethylene linkers between the central nitrogen atom and the thioether sulfur atoms have been spectroscopically and electrochemically characterized. Cyclic voltammetry measurements indicated that the highest Cu2+/Cu+ redox potentials correspond to sulfur‐rich coordination environments, with values decreasing as the thioether donors are replaced by nitrogen‐donating benzimidazoles. Both Cu2+ and Cu+ complexes were studied by DFT. Their electronic properties were determined by analyzing their frontier orbitals, relative energies, and the contributions to the orbitals involved in redox processes, which revealed that the HOMOs of the more sulfur‐rich copper complexes, particularly those with methylene linkers (? N? CH2? S? ), show significant aromatic thioether character. Thus, the theoretically predicted initial oxidation at the sulfur atom of the methylene‐bridged ligands agrees with the experimentally determined oxidation waves in the voltammograms of the NS3‐ and N2S2‐type ligands as being ligand‐based, as opposed to the copper‐based processes of the ethylene‐bridged Cu+ complexes. The electrochemical and theoretical results are consistent with our previously reported mechanistic proposal for Cu2+‐promoted oxidative C? S bond cleavage, which in this work resulted in the isolation and complete characterization (including by X‐ray crystallography) of the decomposition products of two ligands employed, further supporting the novel reactivity pathway invoked. The combined results raise the possibility that the reactions of copper–thioether complexes in chemical and biochemical systems occur with redox participation of the sulfur atom.  相似文献   

3.
A new ratiometric fluorescent sensor ( 1 ) for Cu2+ based on 4,4‐difluoro‐4‐bora‐3a,4a‐diaza‐s‐indacene (BODIPY) with di(2‐picolyl)amine (DPA) as ion recognition subunit has been synthesized and investigated in this work. The binding abilities of 1 towards different metal ions such as alkali and alkaline earth metal ions (Na+, K+, Mg2+, Ca2+) and other metal ions ( Ba2+, Zn2+, Cd2+, Fe2+, Fe3+, Pb2+, Ni2+, Co2+, Hg2+, Ag+) have been examined by UV‐vis and fluorescence spectroscopies. 1 displays high selectivity for Cu2+ among all test metal ions and a ~10‐fold fluorescence enhancement in I582/I558 upon excitation at visible excitation wavelength. The binding mode of 1 and Cu2+ is a 1:1 stoichiometry determined via studies of Job plot, the nonlinear fitting of the fluorometric titration and ESI mass.  相似文献   

4.
In this research, we successfully synthesized and fully characterized the new compound 5,8,13,16,21,24‐hex‐(triisopropylsilyl)ethynyl)‐6,23‐dihydro‐6,7,14,15,22,23‐hexaza‐trianthrylene ( HHATA , brown color in a mixed solvent of CH2Cl2/CH3CN 1:1, v/v, weakly blue fluorescent), which can be easily oxidized to 5,8,13,16,21,24‐hex‐(triisopropylsilyl)ethynyl)‐6,7,14,15,22,23‐hexazatrianthrylene ( HATA ) (yellow color in CH2Cl2/CH3CN 1:1, v/v), red fluorescent) by Cu2+ ions. This reaction only proceeds efficiently in the presence of Cu2+ ions when compared with other common metal ions such as Fe3+, Co2+, Mn2+, Hg2+, Ni2+, Pb2+, Ag+, Mg2+, Ca2+, K+, Na+, and Li+. Our result suggests that this reaction can be developed as an effective method for the detection of Cu2+ ions.  相似文献   

5.
The syntheses, structures, and characterization of six Ln3+–Cu2+–glycine (Hgly) coordination polymers are described in this paper. They represent three types of structures. Type I (Ln=La ( 1 ), Pr ( 2 ), and Sm ( 3 )) is a 1D catenarian polymer comprising [Ln2] nodes bridged by four cis‐Cu(gly)2 linkers. Type II (Ln=Eu ( 4 ) and Dy ( 5 )) is a 2D open framework with a 44‐net, composed of novel [Ln6Cu22] cluster nodes linked by trans‐Cu(gly)2 linkers. Furthermore, the inner structures of the [Ln6Cu22] nodes, and the connection mode between the nodes and linkers are slightly different for 4 and 5 . Type III (Ln=Er ( 6 )) is a 3D open framework with a novel 36?418?53?6 topology, made up of [Er6Cu24] cluster nodes and trans‐Cu(gly)2 linkers. The rich variety of the resulting structures owes itself mainly to the interselection between the dynamic control of metalloligands and cationic components. A transition from frequency dependence to frequency independence is observed in the field‐induced magnetization lag for 1 – 3 . The frequency dependence at low temperatures may come from the antiferromagnetic Cu? Cu interaction through the [Ln2] nodes, whereas the frequency independence may be due to the disappearance of the antiferromagnetic Cu? Cu interaction at high temperatures.  相似文献   

6.
The intrinsic binding ability of 7 natural peptides (oxytocin, arg8‐vasopressin, bradykinin, angiotensin‐I, substance‐P, somatostatin, and neurotensin) with copper in 2 different oxidation states (CuI/II) derived from different Cu+/2+ precursor sources have been investigated for their charge‐dependent binding characteristics. The peptide‐CuI/II complexes, [M − (n‐1)H + nCuI] and [M − (2n‐1)H + nCuII], are prepared/generated by the reaction of peptides with CuI solution/Cu‐target and CuSO4 solution and are analyzed by using matrix‐assisted laser desorption/ionization (MALDI) time‐of‐flight mass spectrometry. The MALDI mass spectra of both [M − (n‐1)H + nCuI] and [M − (2n‐1)H + nCuII] complexes show no mass shift due to the loss of ─H atoms in the main chain ─NH of these peptides by Cu+ and Cu2+ deprotonation. The measured m/z value indicates the reduction of CuI/II oxidation state into Cu0 during MALDI processes. The number and relative abundance of Cu+ bound to the peptides are greater compared with the Cu2+ bound peptides. Oxytocin, arg8‐vasopressin, bradykinin, substance‐P, and somatostatin show the binding of 5Cu+, and angiotensin‐I and neurotensin show the binding of 7Cu+ from both CuI and Cu targets, while bradykinin shows the binding of 2Cu2+, oxytocin, arg8‐vasopressin, angiotensin‐I, and substance‐P; somatostatin shows the binding of 3Cu2+; and neurotensin shows 4Cu2+ binding. The binding of more Cu+ with these small peptides signifies that the bonding characteristics of both Cu+ and Cu2+ are different. The amino acid residues responsible for the binding of both Cu+ and Cu2+ in these peptides have been identified based on the density functional theory computed binding energy values of Cu+ and the fragment transformation method predicted binding preference of Cu2+ for individual amino acids.  相似文献   

7.
We report a new approach to create metal‐binding site in a series of metal–organic frameworks (MOFs), where tetratopic carboxylate linker, 4′,4′′,4′′′,4′′′′‐methanetetrayltetrabiphenyl‐4‐carboxylic acid, is partially replaced by a tritopic carboxylate linker, tris(4‐carboxybiphenyl)amine, in combination with monotopic linkers, formic acid, trifluoroacetic acid, benzoic acid, isonicotinic acid, 4‐chlorobenzoic acid, and 4‐nitrobenzoic acid, respectively. The distance between these paired‐up linkers can be precisely controlled, ranging from 5.4 to 10.8 Å, where a variety of metals, Mg2+, Al3+, Cr3+, Mn2+, Fe3+, Co2+, Ni2+, Cu2+, Zn2+, Ag+, Cd2+ and Pb2+, can be placed in. The distribution of these metal‐binding sites across a single crystal is visualized by 3D tomography of laser scanning confocal microscopy with a resolution of 10 nm. The binding affinity between the metal and its binding‐site in MOF can be varied in a large range (observed binding constants, Kobs from 1.56×102 to 1.70×104 L mol?1), in aqueous solution. The fluorescence of these crystals can be used to detect biomarkers, such as cysteine, homocysteine and glutathione, with ultrahigh sensitivity and without the interference of urine, through the dissociation of metal ions from their binding sites.  相似文献   

8.
A 2D coordination compound {[Cu2(HL)(N3)]?ClO4} ( 1 ; H3L=2,6‐bis(hydroxyethyliminoethyl)‐4‐methyl phenol) was synthesized and characterized by single‐crystal X‐ray diffraction to be a polymer in the crystalline state. Each [Cu2(HL)(N3)]+ species is connected to its adjacent unit by a bridging alkoxide oxygen atom of the ligand to form a helical propagation along the crystallographic a axis. The adjacent helical frameworks are connected by a ligand alcoholic oxygen atom along the crystallographic b axis to produce pleated 2D sheets. In solution, 1 dissociates into [Cu2(HL)2(H3L)]?2H2O ( 2 ); the monomer displays high selectivity for Zn2+ and can be used in HEPES buffer (pH 7.4) as a zinc ion selective luminescent probe for biological application. The system shows a nearly 19‐fold Zn2+‐selective chelation‐enhanced fluorescence response in the working buffer. Application of 2 to cultured living cells (B16F10 mouse melanoma and A375 human melanoma) and rat hippocampal slices was also studied by fluorescence microscopy.  相似文献   

9.
This work reports a new electrochemical monitoring platform for sensitive detection of Cu2+ coupling click chemistry with nanogold‐functionalized PAMAM dendrimer (AuNP‐PAMAM). The system involved an alkyne‐modified carbon electrode and an azide‐functionalized AuNP‐PAMAM. Initially, the added Cu2+ was reduced to Cu+ by the ascorbate, and then the azide‐modified AuNP‐PAMAM was covalently conjugated to the electrode via Cu+‐catalyzed azide‐alkyne click reaction. The carried AuNPs accompanying PAMAM dendrimer could be directly monitored by stripping voltammetry after acidic pretreatment. By introduction of high‐loading PAMAM dendrimer with gold nanoparticles, as low as 2.8 pM Cu2+ (ppt) could be detected, which was 125‐fold lower than that of gold nanoparticle‐based labeling strategy. The method exhibited high specificity toward target Cu2+ against other potentially interfering ions, and was applicable for monitoring Cu2+ in drinking water with satisfactory results.  相似文献   

10.
Five metal–organic frameworks (MOFs) formed by [WS4Cux]x?2 secondary building units (SBUs) and multi‐pyridyl ligands are presented. The [WS4Cux]x?2 SBUs function as network vertexes showing various geometries and connectivities. Compound 1 contains one‐dimensional channels formed in fourfold interpenetrating diamondoid networks with a hexanuclear [WS4Cu5]3+ unit as SBU, which shows square‐pyramidal geometry and acts as a tetrahedral node. Compound 2 contains brick‐wall‐like layer also with a hexanuclear [WS4Cu5]3+ unit as SBU. The [WS4Cu5]3+ unit in 2 is a new type of [WS4Cux]x?2 cluster unit in which the five Cu+ ions are in one plane with the W atom, forming a planar unit. Compound 3 shows a nanotubular structure with a pentanuclear [WS4Cu4]2+ unit as SBU, which is saddle‐shaped and acts as a tetrahedral node. Compound 4 contains large cages formed between two interpenetrated (10,3)‐a networks also with a pentanuclear [WS4Cu4]2+ unit acting as a triangular node. The [WS4Cu4]2+ unit in 4 is isomeric to that in 3 and first observed in a MOF. Compound 5 contains zigzag chains with a tetrahedral [WS4Cu3]+ unit as SBU, which acts as a V‐shaped connector. The influence of synthesis conditions including temperature, ligand, anions of CuI salts, and the ratio of [NH4]2WS4 to CuI salt on the formation of these [WS4Cux]x?2‐based MOFs were also studied. Porous MOF 3 is stable upon removal and exchange of the solvent guests, and when accommodating different solvent molecules, it exhibits specific colors depending on the polarity of incorporated solvent, that is, it shows a rare solvatochromic effect and has interesting prospects in sensing applications.  相似文献   

11.
As the first example of a photocatalytic system for splitting water without additional cocatalysts and photosensitizers, the comparatively cost‐effective Cu2I2‐based MOF, Cu‐I‐bpy (bpy=4,4′‐bipyridine) exhibited highly efficient photocatalytic hydrogen production (7.09 mmol g−1 h−1). Density functional theory (DFT) calculations established the electronic structures of Cu‐I‐bpy with a narrow band gap of 2.05 eV, indicating its semiconductive behavior, which is consistent with the experimental value of 2.00 eV. The proposed mechanism demonstrates that Cu2I2 clusters of Cu‐I‐bpy serve as photoelectron generators to accelerate the copper(I) hydride interaction, providing redox reaction sites for hydrogen evolution. The highly stable cocatalyst‐free and self‐sensitized Cu‐I‐bpy provides new insights into the future design of cost‐effective d10‐based MOFs for highly efficient and long‐term solar fuels production.  相似文献   

12.
The crystal structures of tricopper decavanadate tetracosa­hydrate, (I), and copper tetra­sodium decavanadate tricosa­hydrate, (II), have been determined by single‐crystal X‐ray diffraction. Both compounds exhibit a catenary structure consisting of [V10O28]6− anions linked by Cu2+ cations in (I) or by Na+ cations in (II). Compound (II) also contains a polymeric linear array of edge‐sharing [Na(OH2)6]+ and [Cu(OH2)6]2+ octahedra. In both compounds, the [V10O28]6− ions lie about inversion centres and the Cu2+ ions in (I) also lie about inversion centers.  相似文献   

13.
Designing small peptides that are capable of binding Cu2+ ions mainly through the side‐chain functionalities is a hard task because the amide nitrogen atoms strongly compete for Cu2+ ion coordination. However, the design of such peptides is important for obtaining biomimetic small systems of metalloenyzmes as well as for the development of artificial systems. With this in mind, a cyclic decapeptide, C‐Asp, which contained three His residues and one Asp residue, and its linear derivative, O‐Asp, were synthesized. The C‐Asp peptide has two Pro? Gly β‐turn‐inducer units and, as a result of cyclization, and as shown by CD spectroscopy, its backbone is constrained into a more defined conformation than O‐Asp, which is linear and contains a single Pro? Gly unit. A detailed potentiometric, mass spectrometric, and spectroscopic study (UV/Vis, CD, and EPR spectroscopy) showed that at a 1:1 Cu2+/peptide ratio, both peptides formed a major [CuHL]2+ species in the pH range 5.0–7.5 (C‐Asp) and 5.5–7.0 (O‐Asp). The corrected stability constants of the protonated species (log K*CuH(O?Asp)=9.28 and log K*CuH(C?Asp)=10.79) indicate that the cyclic peptide binds Cu2+ ions with higher affinity. In addition, the calculated value of Keff shows that this higher affinity for Cu2+ ions prevails at all pH values, not only for a 1:1 ratio but even for a 2:1 ratio. The spectroscopic data of both [CuHL]2+ species are consistent with the exclusive coordination of Cu2+ ions by the side‐chain functionalities of the three His residues and the Asp residue in a square‐planar or square‐pyramidal geometry. Nonetheless, although these data show that, upon metal coordination, both peptides adopt a similar fold, the larger conformational constraints that are present in the cyclic scaffold results in different behaviour for both [CuHL]2+ species. CD and NMR analysis revealed the formation of a more rigid structure and a slower Cu2+‐exchange rate for [CuH(C‐Asp)]2+ compared to [CuH(O‐Asp]2+. This detailed comparative study shows that cyclization has a remarkable effect on the Cu2+‐coordination properties of the C‐Asp peptide, which binds Cu2+ ions with higher affinity at all pH values, stabilizes the [CuHL]2+ species in a wider pH range, and has a slower Cu2+‐exchange rate compared to O‐Asp.  相似文献   

14.
The catalytic activity of Pt nanoparticles (PtNPs) with different sizes and shapes was investigated in a photocatalytic hydrogen‐evolution system composed of the 9‐mesityl‐10‐methylacridinium ion (Acr+–Mes: photocatalyst) and dihydronicotinamide adenine dinucleotide (NADH: electron donor), based on rates of hydrogen evolution and electron transfer from one‐electron‐reduced species of Acr+–Mes (Acr.–Mes) to PtNPs. Cubic PtNPs with a diameter of (6.3±0.6) nm exhibited the maximum catalytic activity. The observed hydrogen‐evolution rate was virtually the same as the rate of electron transfer from Acr.–Mes to PtNPs. The rate constant of electron transfer (ket) increased linearly with increasing proton concentration. When H+ was replaced by D+, the inverse kinetic isotope effect was observed for the electron‐transfer rate constant (ket(H)/ket(D)=0.47). The linear dependence of ket on proton concentration together with the observed inverse kinetic isotope effect suggests that proton‐coupled electron transfer from Acr.–Mes to PtNPs to form the Pt? H bond is the rate‐determining step for catalytic hydrogen evolution. When FeNPs were used instead of PtNPs, hydrogen evolution was also observed, although the hydrogen‐evolution efficiency was significantly lower than that of PtNPs because of the much slower electron transfer from Acr.–Mes to FeNPs.  相似文献   

15.
《Electroanalysis》2006,18(18):1827-1832
Studies of nitric oxide (NO) release from S‐nitrosoglutathione (GSNO) decomposition by Cu2+ in the presence of reducing agents were performed using a nickel porphyrin and Nafion‐coated microsensor in order to compare the efficiency of sodium hydrosulfite (Na2S2O4) and sodium borohydride (NaBH4) to that of the most abundant endogenous reducer, glutathione (GSH). When it was mixed to Cu(NO3)2 and added to equimolar concentration of GSNO, each reducing agent caused a NO release (measured in terms of oxidation current) but only NaBH4 induced a proportional rise if its concentration doubled and that of Cu2+ remained constant. For Na2S2O4, there was a mild increase and for GSH, no change. Furthermore, when Cu2+ concentrations ranging from 0.5 to 5 μM were mixed with 2 μM reducing agent and added to 2 μM GSNO, the NO oxidation current linearly increased with NaBH4 and was constant with Na2S2O4. Concerning GSH, Cu2+ dose‐dependently increased the NO release from GSNO only if the Cu2+‐to‐reducer ratio was ≤1. However, GSH formed the catalytic species Cu+ even in excess of Cu2+ and GSNO as indicated by suppression of the Cu2+/GSH‐induced NO release when the Cu+ chelator neocuproine was added to GSNO. This work shows that, among the 3 reducing agents, only NaBH4 allows Cu2+ to dose‐dependently increase the NO release from GSNO for Cu2+‐to‐reducer ratios ranging from 0.25 to 2.5. Despite this good effectiveness, excess of NaBH4 compared to both Cu2+ and GSNO seems to be required for optimal NO release.  相似文献   

16.
The metal–organic framework (MOF) [Pd(2‐pymo)2]n (2‐pymo=2‐pyrimidinolate) was used as catalyst in the hydrogenation of 1‐octene. During catalytic hydrogenation, the changes at the metal nodes and linkers of the MOF were investigated by in situ X‐ray absorption spectroscopy (XAS) and IR spectroscopy. With the help of extended X‐ray absorption fine structure and X‐ray absorption near edge structure data, Quick‐XAS, and IR spectroscopy, detailed insights into the catalytic relevance of Pd2+/Pd0 in the hydrogenation of 1‐octene could be achieved. Shortly after exposure of the catalyst to H2 and simultaneously with the hydrogenation of 1‐octene, the aromatic rings of the linker molecules are hydrogenated rapidly. Up to this point, the MOF structure remained intact. After completion of linker hydrogenation, the linkers were also protonated. When half of the linker molecules were protonated, the onset of reduction of the Pd2+ centers to Pd0 was observed and the hydrogenation activity decreased, followed by fast reduction of the palladium centers and collapse of the MOF structure. Major fractions of Pd0 are only observed when the hydrogenation of 1‐octene is almost finished. Consequently, the Pd2+ nodes of the MOF [Pd(2‐pymo)2]n are identified as active centers in the hydrogenation of 1‐octene.  相似文献   

17.
Double‐stranded copper(II) string complexes of varying nuclearity, from di‐ to tetranuclear species, have been prepared by the CuII‐mediated self‐assembly of a novel family of linear homo‐ and heteropolytopic ligands that contain two outer oxamato and either zero ( 1 b ), one ( 2 b ), or two ( 3 b ) inner oxamidato donor groups separated by rigid 2‐methyl‐1,3‐phenylene spacers. The X‐ray crystal structures of these CuIIn complexes (n=2 ( 1 d ), 3 ( 2 d ), and 4 ( 3 d )) show a linear array of metal atoms with an overall twisted coordination geometry for both the outer CuN2O2 and inner CuN4 chromophores. Two such nonplanar allsyn bridging ligands 1 b – 3 b in an anti arrangement clamp around the metal centers with alternating M and P helical chiralities to afford an overall double meso‐helicate‐type architecture for 1 d – 3 d . Variable‐temperature (2.0–300 K) magnetic susceptibility and variable‐field (0–5.0 T) magnetization measurements for 1 d – 3 d show the occurrence of S=nSCu (n=2–4) high‐spin ground states that arise from the moderate ferromagnetic coupling between the unpaired electrons of the linearly disposed CuII ions (SCu=1/2) through the two anti m‐phenylenediamidate‐type bridges (J values in the range of +15.0 to 16.8 cm?1). Density functional theory (DFT) calculations for 1 d – 3 d evidence a sign alternation of the spin density in the meta‐substituted phenylene spacers in agreement with a spin polarization exchange mechanism along the linear metal array with overall intermetallic distances between terminal metal centers in the range of 0.7–2.2 nm. Cyclic voltammetry (CV) and rotating‐disk electrode (RDE) electrochemical measurements for 1 d – 3 d show several reversible or quasireversible one‐ or two‐electron steps that involve the consecutive metal‐centered oxidation of the inner and outer CuII ions (SCu=1/2) to diamagnetic CuIII ones (SCu=0) at relatively low formal potentials (E values in the range of +0.14 to 0.25 V and of +0.43 to 0.67 V vs. SCE, respectively). Further developments may be envisaged for this family of oligo‐m‐phenyleneoxalamide copper(II) double mesocates as electroswitchable ferromagnetic ‘metal–organic wires’ (MOWs) on the basis of their unique ferromagnetic and multicenter redox behaviors.  相似文献   

18.
The title compound, with nominal formula Cu2ScZr(PO4)3, has a beige coloration and displays fast Cu+ cation conduction at elevated temperatures. It adopts a NASICON‐type structure in the space group Rc. The examined crystal was an obverse–reverse twin with approximately equal twin components. The [ScIIIZrIV(PO4)3]2− framework is composed of corner‐sharing Sc/ZrO6 octahedra and PO4 tetrahedra. The Sc and Zr atoms are disordered on one atomic site on a crystallographic threefold axis. The P atom of the phosphate group lies on a crystallographic twofold axis. Nonframework Cu+ cations occupy three positions. Two of the Cu+ positions generate an approximately circular distribution around a site of symmetry, referred to as the M1 site in the NASICON‐type structure. The other Cu+ position is situated close to the twofold symmetric M2 site, displaced into a position with a distorted square‐based pyramidal coordination geometry. The structure has been determined at 100, 200 and 300 K. Changes in the refined site‐occupancy factors of the Cu+ positions suggest increased mobility of Cu+ around the circular orbit close to the M1 site at room temperature, but no movement into or out of the M2 site. Free refinement of the Cu site‐occupancy factors suggests that the formula of the crystal is Cu1.92(1)ScZr(PO4)3, which is consistent with the low‐level presence of Cu2+ exclusively in the M2 site.  相似文献   

19.
The synthesis of stable porous materials with appropriate pore size and shape for desired applications remains challenging. In this work a combined experimental/computational approach has been undertaken to tune the stability under various conditions and the adsorption behavior of a series of MOFs by subtle control of both the nature of the metal center (Co2+, Cu2+, and Zn2+) and the pore surface by the functionalization of the organic linkers with amido and N‐oxide groups. In this context, six isoreticular MOFs based on T‐shaped ligands and paddle‐wheel units with ScD0.33 topology have been synthesized. Their stabilities have been systematically investigated along with their ability to adsorb a wide range of gases (N2, CO2, CH4, CO, H2, light hydrocarbons (C1–C4)) and vapors (alcohols and water). This study has revealed that the MOF frameworks based on Cu2+ are more stable than their Co2+ and Zn2+ analogues, and that the N‐oxide ligand endows the MOFs with a higher affinity for CO2 leading to excellent selectivity for this gas over other species.  相似文献   

20.
《化学:亚洲杂志》2017,12(20):2727-2733
Hydrogen production by catalytic water splitting using sunlight holds great promise for clean and sustainable energy source. Despite the efforts made in the past decades, challenges still exist in pursuing solid catalysts with light‐harvesting capacity, large surface areas and efficient utilities of the photogenerated carrier, at the same time. Here, a multiple structure design strategy leading to highly enhanced photocatalytic performance on hydrogen production from water splitting in Dion–Jacobson perovskites KCa2Nan ‐3Nbn O3n +1 is described. Specifically, chemical doping (N/Nb4+) of the parent oxides via ammoniation improved the ability of sunlight harvesting efficiently; subsequent liquid exfoliation of the doped perovskites yielded ultrathin [Ca2Nan ‐3Nbn O3n +1] nanosheets with greatly increased surface areas. Significantly, the maximum hydrogen evolution appears in the n =4 nanosheets, which suggests the most favorable thickness for charge separation in such perovskite‐type catalysts. The optimized black N/Nb4+‐[Ca2NaNb4O13] nanosheets show greatly enhanced photocatalytic performance, as high as 973 μmol h−1 with Pt loading, on hydrogen evolution from water splitting. As a proof‐of‐concept, this work highlights the feasibility of combining various chemical strategies towards better catalysts and precise thickness control of two‐dimensional materials.  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