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 78 毫秒
1.
A new artificial photosynthetic triad array, a [60]fullerene–triosmium cluster/zinc–porphyrin/boron–dipyrrin complex ( 1 , Os3C60/ZnP/Bodipy), has been prepared by decarbonylation of Os3(CO)8(CN(CH2)3Si(OEt)3)(μ3‐η222‐C60) ( 6 ) with Me3NO/MeCN and subsequent reaction with the isocyanide ligand CNZnP/Bodipy ( 5 ) containing zinc porphyrin (ZnP) and boron dipyrrin (Bodipy) moieties. Triad 1 has been characterized by various spectroscopic methods (MS, NMR, IR, UV/Vis, photoluminescence, and transient absorption spectroscopy). The electrochemical properties of 1 in chlorobenzene (CB) have been examined by cyclic voltammetry; the general feature of the cyclic voltammogram of 1 is nine reversible one‐electron redox couples, that is, the sum of those of 5 and 6 . DFT has been applied to study the molecular and electronic structures of 1 . On the basis of fluorescence‐lifetime measurements and transient absorption spectroscopic data, 1 undergoes an efficient energy transfer from Bodipy to ZnP and a fast electron transfer from ZnP to C60; the detailed kinetics involved in both events have been elucidated. The SAM of triad 1 ( 1 /ITO; ITO=indium–tin oxide) has been prepared by immersion of an ITO electrode in a CB solution of 1 and diazabicyclo‐octane (2:1 equiv), and characterized by UV/Vis absorption spectroscopy, water contact angle, X‐ray photoelectron spectroscopy, and cyclic voltammetry. The photoelectrochemical properties of 1 /ITO have been investigated by a standard three‐electrode system in the presence of an ascorbic acid sacrificial electron donor. The quantum yield of the photoelectrochemical cell has been estimated to be 29 % based on the number of photons absorbed by the chromophores. Our triad 1 is unique when compared to previously reported photoinduced electron‐transfer arrays, in that C60 is linked by π bonding with little perturbation of the C60 electron delocalization.  相似文献   

2.
Raw oil shale, kerogen (demineralized shale) and carbonaceous residues from kerogen pyrolysis in the range 350–700°C (at 50°C intervals) were studied by laser ablation Fourier transform ion cyclotron resonance mass spectrometry using the fundamental frequency of Nd: YAG laser (1064 nm). Normally, pyrolysis of the raw materials produces oil and the resulting residues have decreased hydrogen to carbon ratios and exhibit relative increases in aromatic carbons. Raw shale and kerogen give positive-ion spectra with mainly protonated species of m/z 100–400. Laser ablation positive-ion mass spectra of the pyrolysis products of the kerogen show the presence of C60, C70 and other fullerene ions with a distribution of higher mass fullerene ions up to m/z 4000. Using high laser powers (100–3000 MW cm?2), the residue from pyrolysis at 350°C initially did not produce any fullerene ions (apart from traces of C60 and C70), but after continued ablation a cavity was formed in the target and a wide distribution of fullerene ions was obtained with subsequent laser pulses. Residues obtained from the pyrolysis of kerogen at 400–500°C produced fullerene ions at both low (4–200 kW cm?2) and high laser powers. The 550°C pyrolysis residue gave only small amounts of C60 and C70 positive ions at low laser power whereas residues from the pyrolysis of kerogen above 550°C did not give fullerene ions over a wide range of laser powers. It is proposed from the above results that the changes in the aromatic nature of the kerogen residues with increasing pyrolysis temperature are directly related to the ease of fullerene formation. This is possibly due to the formation of large polycyclic aromatic systems at pyrolysis temperatures above 400°C, formed in the residues. It should be noted that the shale samples (raw or pyrolysed) did not generate fullerene ions under any of the conditions employed in these experiments.  相似文献   

3.
We investigate the interface between a C60 fullerite film, C60F36, and diamond (100) by using core‐level photoemission spectroscopy, cyclic voltammetry (CV), and high‐resolution electron energy loss spectroscopy (HREELS). We show that C60 can be covalently bonded to reconstructed C(100)‐2×1 and that the bonded interface is sufficiently robust to exhibit characteristic C60 redox peaks in solution. The bare diamond surface can be passivated against oxidation and hydrogenation by covalently bound C60. However, C60F36 is not as stable as C60 and desorbs below 300 °C (the latter species being stable up to 500 °C on the diamond surface). Neither C60 fullerite nor C60F36 form reactive interfaces on the hydrogenated surface—they both desorb below 300 °C. The surface transfer doping process of hydrogenated diamond by C60F36 is the most evident one among all the adsorbate systems studied (with a coverage‐dependent band bending induced by C60F36).  相似文献   

4.
Fullerols of C60 and of C70 [C60(OH)n, C70(OH)m], water-soluble fullerene derivatives, unlike some other fullerene derivatives (such as C60 (C4H6O), C60 (C3H7N) and C60 [C(COOEt)2]x), do not result in excited triplet state but in ionization via monophotonic process in aqueous solutions with 248 nm laser. The quantum yields of formation of hydrated electron (Φe ) are determined to be 0.08 and 0.11 for fullerols of C60 and of C70 respectively at room temperature (ca. 15°C) with KI solution used as reference. By laser flash photolysis and oxidation of sulfate radical anion SO4 , the fullerol radical cation or neutral radical of C60 is confirmed to be existent and the transient absorption spectra of fullerol radical cation of C70 are observed for the first time. Project supported by the National Natural Science Foundation of China  相似文献   

5.
Poly(trifluoromethyl)fullerene S6‐C60(CF3)12 was reduced by sodium fluorenone ketyl in the presence of (PPN)Cl (PPN=bis(triphenylphosphine)iminium) to afford the salt (PPN)[C60(CF3)12] ( 1 ), which contains C60(CF3)12.? radical anions. In the crystal structure of 1 , C60(CF3)12.? layers alternate with the PPN+ cations. There are short F ??? F contacts between C60(CF3)12.? radical anions within the layers but no C ??? C contacts. DFT calculations revealed that the negative charge on C60(CF3)12.? is distributed mainly between sp2 carbon and fluorine atoms, whereas spin density is localized mainly on the fullerene‐cage sp2 carbon atoms. IR and UV/Vis/NIR spectra in the solid state and solution showed characteristic changes relative to those of neutral S6‐C60(CF3)12 due to the formation of radical anions. The solid‐state electronic spectrum of 1 exhibits a single broad band at 738 nm attributed to C60(CF3)12.?. Crystals of 1 show a narrow EPR signal with g=2.0025 (ΔH=0.45 mT) at 300 K. The temperature dependence of the integral intensity follows the Curie–Weiss law with a negative Weiss temperature of ?11.8 K (30–300 K) indicating antiferromagnetic interaction of spins. This dependence was approximated by the Heisenberg model for one‐dimensional chains of antiferromagnetically interacting spins with exchange interaction J/kB=?9.1 K. It was assumed that magnetic interaction between the C60(CF3)12.? spins in the layers is mediated by short F ??? F contacts.  相似文献   

6.
利用聚二甲基二烯丙基氯化铵(PDDA)非共价修饰的碳纳米管(CNTs)与PtCl62-之间的自发氧化还原作用, 制备了Pt 纳米颗粒(Pt NPs)/CNTs-PDDA复合催化剂. PDDA在该催化剂中具有三种作用: (1) 作为金属前驱体PtCl62-还原为Pt NPs 的还原剂; (2) 作为原位产生的Pt NPs 的稳定剂; (3) 在CNTs 表面形成保护膜抑制CNTs 在甲醇电催化氧化过程中的腐蚀. 采用傅里叶变换红外(FTIR)光谱、热重分析和拉曼光谱对CNTs-PDDA进行了表征, 表明PDDA通过π-π作用已成功覆盖在CNTs 表面, 并且修饰过程没有导致CNTs 结构的破坏. 采用透射电子显微镜(TEM)对Pt NPs/CNTs-PDDA 催化剂进行了表征, 结果表明, Pt NPs 均匀地分散在CNTs上, 平均粒径约2 nm, 且粒径分布范围窄. 用循环伏安法、计时电流法进一步考察了Pt NPs/CNTs-PDDA催化剂在酸性介质中对甲醇的电催化氧化的性能. 电化学测试结果表明, 与原始CNTs 负载的Pt NPs催化剂相比,Pt NPs/CNTs-PDDA催化剂具有更高的电化学活性表面积、电催化质量比活性和稳定性.  相似文献   

7.
The structure of the title compound, [Pt(C6H5)2(C6H12N3P)2] or [Pt(Ph)2(PTA)2] (where Ph is phenyl and PTA is 1,3,5‐tri­aza‐7‐phosphaadamantane), is discussed. Selected geom­etric parameters are: Pt—P = 2.2888 (16) and 2.2944 (17) Å, Pt—C = 2.052 (5) and 2.064 (6) Å, C—Pt—C = 84.6 (2)° and P—Pt—P = 99.28 (6)°. The effective cone angle for the PTA ligands was calculated as 113°.  相似文献   

8.
A10(PO4)6(OH)2 (A = Ca and Sr)-supported Pt catalysts were prepared and their catalytic activity in NO reduction were investigated. The Sr10(PO4)6(OH)2-supported catalyst had high catalytic activity in the C3H6?CNO?CO2 reaction; the activity was higher than that of the ??-Al2O3-supported catalyst at 300 °C. The basicity of the apatite supports would affect the chemical state of Pt on catalyst, resulting in promotion of NO reduction.  相似文献   

9.
Summary Hydrophilic endohedral 133Xe-fullerenols, [133Xe@C60(OH)xand 133Xe@C70(OH)x], were synthesized from hydrophobic endohedral 133Xe-fullerenes. The yield of endohedral 133Xe-fullerenols extracted in water was about 40% and 23% for C60and C70, respectively. The products stored in 0.9% NaCl solution at 20 °C were stable enough to be used in nuclear medicine.  相似文献   

10.
Wang  Qing  Liu  Ying  Xu  Fang  Liu  Qi  Cui  Da 《Journal of Thermal Analysis and Calorimetry》2019,136(4):1631-1643

The cleavage behavior of covalent bonds in Xilinguole (XLGL) lignite and changes in chemical structure of lignite and its chars during low-temperature pyrolysis were investigated by thermogravimetric (TG) analysis and Fourier-transform infrared (FTIR) spectroscopy. Based on the TG and differential thermogravimetric (DTG) analysis results, the cleavage of different types of chemical bonds in lignite occurred mainly at four certain temperatures, 170 °C, 376 °C, 432 °C, and 521 °C. The latter three were selected as the final pyrolysis temperatures of chars evaluated in this study. The FTIR analysis results indicate that thermal treatment increased the relative content of two and three adjacent H deformation structures but decreased that of four adjacent H deformation structure. This was caused by the cleavage of Cal–Cal and Car–Cal bonds. The oxygen-containing functional groups in lignite are dominated by C–O and C–OH groups with a lower chemical reactivity than C=O–C and conjugated C=O groups. Moreover, XLGL lignite has the highest ratio of CH2/CH3 which declines with increasing temperature, indicating the decrease in the length of aliphatic chains and increase in the degree of branching of aliphatic side chains. This change mainly resulted from the cleavage of Cal–O, Cal–Cal, and Car–Cal bonds. Furthermore, XLGL lignite and its chars contain five specific hydrogen bonds: OH–N, cyclic OH, OH–ether O, OH–OH, and OH–π hydrogen bonds. The relative content of OH–OH hydrogen bond was the highest, indicating that OH–OH hydrogen bond has the highest thermal stability.

  相似文献   

11.
Single-atom catalysts (SACs) have emerged as promising materials in heterogeneous catalysis. Previous studies reported controversial results about the relative level in activity for SACs and nanoparticles (NPs). These works have focused on the effect of metal atom arrangement, without considering the oxidation state of the SACs. Here, we immobilized Pt single atoms on defective ceria and controlled the oxidation state of Pt SACs, from highly oxidized (Pt0: 16.6 at %) to highly metallic states (Pt0: 83.8 at %). The Pt SACs with controlled oxidation states were then employed for oxidation of CO, CH4, or NO, and their activities compared with those of Pt NPs. The highly oxidized Pt SACs presented poorer activities than Pt NPs, whereas metallic Pt SACs showed higher activities. The Pt SAC reduced at 300 °C showed the highest activity for all the oxidations. The Pt SACs with controlled oxidation states revealed a crucial missing link between activity and SACs.  相似文献   

12.
Podlike nitrogen‐doped carbon nanotubes encapsulating FeNi alloy nanoparticles (Pod(N)‐FeNi) were prepared by the direct pyrolysis of organometallic precursors. Cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and Tafel polarization measurements revealed their excellent electrocatalytic activities in the I?/I3? redox reaction of dye‐sensitized solar cells (DSSCs). This is suggested to arise from the modification of the surface electronic properties of the carbon by the encapsulated metal alloy nanoparticles (NPs). Sequential scanning with EIS and CV further showed the high electrochemical stability of the Pod(N)‐FeNi composite. DSSCs with Pod(N)‐FeNi as the counter electrode (CE) presented a power conversion efficiency of 8.82 %, which is superior to that of the control device with sputtered Pt as the CE. The Pod(N)‐FeNi composite thus shows promise as an environmentally friendly, low‐cost, and highly efficient CE material for DSSCs.  相似文献   

13.
Ni(La)-hydroxide films were prepared from aqueous colloidal solutions containing nickel sulfate and lanthanum acetate in the molar ratio 10:1. Two types of film were made by heating for 15 and 60?min at 300?°C. Thermogravimetry (TG) and X-ray diffraction (XRD) reveal that both films consist of NiO (bunsenite 40%) nanoparticles (particle size?~30?Å), the remainder being amorphous. IR spectroscopy showed that the amorphous phase comprised the α(II)-Ni(OH)2 phase incorporating SO4 2?, carboxylate and water species. Cyclic voltammetry (CV) in a 0.1?M LiOH electrolyte combined with in situ UV-VIS spectroscopy revealed that the colouring/bleaching changes, as a function of applied potential, differed considerably for the two types of film. Ex situ IR spectroelectrochemical measurements at near-grazing incidence angle conditions using P-polarised light (NGIA IR) were performed for films heated for 60?min in 0.1?M LiOH and 0.1?M tetramethylammonium hydroxide (TMAH) electrolytes and cycled 1402 and 1802 times. During the oxidation/reduction cycles the α(II)-Ni(OH)2 phase transforms to the γ(III)-NiOOH phase, while the β(II)-Ni(OH)2 did not develop. This explains the high cycling stability of Ni(La)-hydroxide films. The incorporation of TMA+ ions was observed from the ν(CH3) stretching band intensities in the IR spectra of cycled films.  相似文献   

14.
The templated borate, [C9H14N] · [B5O6(OH)4], was synthesized under hydrothermal conditions. Single crystal X‐ray diffraction techonology reveals that it crystallizes in the triclinic system, space group P$\bar{1}$ (No. 2). The material was also characterized by element analysis, Fourier transform infrared spectroscopy (FTIR), powder X‐ray diffraction (PXRD), thermogravimetric and differential thermal analysis (TG‐DTA), and luminescence spectroscopy. The compound consisted of isolated pentaborate [B5O6(OH)4] and N‐butylpyridinium cations [C9H14N]+. The [B5O6(OH)4] anions are connected together by hydrogen bonds to form a three‐dimensional framework, in which [C9H14N]+ cations are located in. [C9H14N] · [B5O6(OH)4] exhibits tunable luminescence emission at 415–458 nm by means of heating treatment from 100 to 300 °C.  相似文献   

15.
Tungsten carbide and graphitic carbon (WC/GC) composite has been synthesized by a simple solid-state pyrolysis method from an in situ route. The results indicate that the synthesized sample has a large specific surface area (S BET) of 198 m2 g?1, and the WC nanoparticles (NPs) with a narrow particle size are well dispersed on the graphitic carbon. After loading Pt nanoparticles, the prepared Pt/WC/GC catalyst exhibits a mass activity of 416.1 mA mg?1 Pt toward methanol electrooxidation, which is much higher than that of commercial Pt/C (JM) (231.2 mA mg?1 Pt). Moreover, the onset potential is 100 mV more negative than that on Pt/C (JM) electrocatalyst. In addition, the Pt/WC/GC catalyst has stronger resistance to CO poisoning than the commercial Pt/C (JM). Its superior electrochemical performance could be attributed not only to the synergistic effect between Pt and WC NPs but also to the excellent electrical conductivity of GC and proper porous structure for desirable mass transportation in a porous electrode.  相似文献   

16.
The equilibrium geometric parameters and the energetic characteristics of fullerenol molecules C60(OH)20 and C60(OH)18 and fullerenol-like inorganic clusters B12(OH) 12 2? , Si20O30(OH)20, and Ti20O30(OH)20 and their derivatives C60(OH)20 ? n (OLi) n , C60(OH)18 ? n (OLi) n , B12(OH)12 ? n (OLi) n 2? , Si20O30(OH)20 ? n (OLi) n , and Ti20O30(OH)20 ? n (OLi) n , in which the H atoms of all or one-half of hydroxyl groups are replaced by Li atoms, have been calculated by the density functional theory B3LYP/6-31G* method. It has been found that partial energies ??E(H/Li) per single substitution of Li for H in the reactions C60(OH)20 + nLiAc ?? C60(OH)20 ? n (OLi) n + nHAc and C60(OH)18 + nLiAc ?? C60(OH)18 ? n (OLi) n + nHAc (Ac is acetate) and the energies averaged over the entire series of changes in n, as well as over its first and second halves, do not exceed a few kilocalories per mole. It has been predicted that at least one-half (or more than one-half) of OH groups can be replaced by OLi without noticeable changes in energy; however, with a further increase in n, substitutions become endothermic and require ever-increasing energy inputs. In the completely hydroxylated closo-dodecaborane dianion with an icosahedral [B12] cage and more polar B-O-H bonds, analogous H/Li substitutions are slightly exothermic so that the reaction can proceed somewhat smoother and further (toward larger n values) than in fullerenols, other conditions being the same. In the inorganic clusters Si20O30(OH)20 and Ti20O30(OH)20 with the [Si20] and [Ti20] cages, respectively, and with even more polar Si-O-H and Ti-O-H moieties, the substitutions are even more exothermic (their partial energies ??E(H/Li) increase to 4?C6 kcal/mol). For sodium and potassium analogues, the qualitative pattern persists, but H/Na and H/K substitutions are somewhat less exothermic than the H/Li substitutions. The results are compared to the data of previous calculations of stepwise H/Li and H/Na substitutions in the reactions C60(OH)24 + nLAc ?? C60(OH)24 ? n (OL) n + nHAc (L = Li, Na).  相似文献   

17.
Summary The water-soluble fullerene derivative C60(OH)x was radiolabeled with 67GaCl3. The labeling yields were determined by radio-PLC. The effects of pH, reaction time, temperature and the amount of C60(OH)x on the labeling yields were studied. The stability of 67Ga-C60(OH)x was also examined. The results showed that the labeling yields could reach 97% under the best labeling conditions and the radiochemical purity of 67Ga-C60(OH)x solution kept at 37 °C remained at 88% after 212 hours. The biodistribution studies of 67Ga-C60(OH)x in mice showed a high localization of 67Ga-C60(OH)x in the bone marrow, bone, liver and spleen with slow clearance and a negligible accumulation in the blood. These data suggest that the water-soluble C60(OH)x, having the same properties as microcolloids, may be used as a carrier of drug system for lymphatic targeting.  相似文献   

18.
The paper describes the synthesis of geometrical isomers and diastereomers of Pt(II) bischelates with diastereomeric hydroxy-amino acids threonine (threo-α-amino-β-hydroxybutyric acid CH3C*H(OH)C*H(NH2)COOH=ThrH) and allothreonine (erythro-α-amino-β-hydroxybutyric acid=alloThrH) containing two asymmetric carbon atoms C*: cis-,trans-[Pt(S-Thr)2], cis-,trans-[Pt(RThr)(S-Thr)], cis-,trans-[Pt(R-alloThr)(S-alloThr)] (where R and S are the absolute configurations of the asymmetric carbon atom bonded to the carboxyl group). 195Pt NMR spectroscopy is used to investigate the successive phases of the synthesis of the stereoisomeric Pt(II) complexes with threonine. The synthesized complexes are studied by 1H, 13C, 195Pt NMR spectroscopy, IR spectroscopy, and single crystal XRD.  相似文献   

19.
Analytically pure C60H18 is obtained by a Ru3 cluster complexation and decomplexation method. The crystal structure of C60H18 consists of one flattened hemisphere, to which all 18 hydrogen atoms are symmetrically bonded, and one curved hemisphere akin to C60. A benzenoid ring in the flattened hemisphere is isolated from the residual π systems by a belt composed of sp3‐hybridized CH units. The average out‐of‐plane distances for carbon atoms attached to the benzenoid ring (0.14 Å) is substantially larger than that found in C60F18 (0.06 Å). Several long C(sp3)?C(sp3) single bond lengths [1.61(3)–1.65(3) Å] are observed for C60H18. The reaction of [Ru3(CO)12] and C60H18 produces [Ru3(CO)93‐η222‐C60H18)] ( 1 ), where the Ru3 triangle is regiospecifically linked to the hexagon opposite to the benzenoid ring. Compound 1 is the first transition metal complex of a polyhydrofullerene (fullerane). C60H18 and 1 have been characterized by 1H and 13C NMR, UV/Vis, and mass spectroscopies. The HOMO–LUMO gap of C60H18 is evaluated to be 1.51 V by cyclic voltammetry.  相似文献   

20.
9,10-Dihydroanthracene (C14H12 reacts with platinum fulleride C60Pt at 513-623 K under anaerobic conditions to form anthracene and hydrofullerenes identified by IR and mass spectroscopy. The Pt 4f 7/2 binding energy of platinum in the initial fulleride (72.4 eV) indicates partial charge transfer form Pt to C60, which agrees with the results of X-ray fluorescence spectroscopic study of C60Pt.  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