首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   220篇
  免费   178篇
  国内免费   1篇
  2024年   2篇
  2023年   3篇
  2022年   2篇
  2021年   3篇
  2020年   23篇
  2019年   24篇
  2018年   34篇
  2017年   25篇
  2016年   38篇
  2015年   48篇
  2014年   48篇
  2013年   33篇
  2012年   8篇
  2011年   17篇
  2010年   1篇
  2009年   6篇
  2008年   7篇
  2007年   9篇
  2006年   11篇
  2005年   3篇
  2004年   3篇
  2003年   4篇
  2002年   3篇
  2001年   4篇
  2000年   4篇
  1999年   2篇
  1997年   2篇
  1996年   2篇
  1995年   4篇
  1994年   1篇
  1993年   2篇
  1992年   2篇
  1991年   5篇
  1990年   3篇
  1989年   1篇
  1986年   1篇
  1985年   3篇
  1984年   4篇
  1983年   1篇
  1982年   1篇
  1981年   1篇
  1980年   1篇
排序方式: 共有399条查询结果,搜索用时 46 毫秒
281.
Carbon‐based hole transport material (HTM)‐free perovskite solar cells (PSCs) have shown much promise for practical applications because of their high stability and low cost. However, the efficiencies of this kind of PSCs are still relatively low, especially for the simplest paintable carbon‐based PSCs, in comparison with the organic HTM‐based PSCs. This can be imputed to the perovskite deposition methods that are not very suitable for this kind of devices. A solvent engineering strategy based on two‐step sequential method is exploited to prepare a high‐quality perovskite layer for the paintable carbon‐based PSCs in which the solvent for CH3NH3I (MAI) solution at the second step is changed from isopropanol (IPA) to a mixed solvent of IPA/Cyclohexane (CYHEX). This mixed solvent not only accelerates the conversion of PbI2 to CH3NH3PbI3 but also suppresses the Ostwald ripening process resulting in a high‐quality perovskite layer, e.g., pure phase, even surface, and compact capping layer. The paintable carbon‐based PSCs fabricated from IPA/CYHEX solvent exhibits a considerable enhancement in photovoltaic performance and performance reproducibility in comparison with that from pure IPA, especially on fill factor (FF), owing mainly to the better contact of perovskite/carbon interface, lower trap density in perovskite, higher light absorption ability, and faster charge transport of perovskite layer. As a result, the highest power conversion efficiency (PCE) of 14.38% is obtained, which is a record value for carbon‐based HTM‐free PSCs. Furthermore, a PCE of as high as 10% is achieved for the large area device (1 cm2), also the highest of its kind.  相似文献   
282.
283.
Currently studied carbon nanotube‐silicon (CNT‐Si) solar cells are based on relatively small active areas (typically <0.15 cm2); increasing the active area generally leads to reduced power conversion efficiencies. This study reports CNT‐Si solar cells with active areas of more than 2 cm2 for single cells, yet still achieving cell efficiencies of about 10%, which is the first time for CNT‐Si solar cells with an active area more than 1 cm2 to reach the level for real applications. In this work, a controlled number of flattened highly conductive CNT strips is added, in simple arrangement, to form a CNT‐Si solar cell with CNT strips in which the middle film makes heterojunctions with Si while the top strips act as self‐similar top electrodes, like conventional metal grids. The CNT strips, directly condensed from as‐grown CNT films, not only improve the CNT‐Si junctions, but also enhance the conductivity of top electrodes without introducing contact barrier when the CNT strips are added onto the film. This property may facilitate the development of large‐area high‐performance CNT or graphene‐Si solar cells.  相似文献   
284.
Cytochrome c 552 (Cyt-c 552) and its redox partner ba 3 -oxidase from Thermus thermophilus possess structural differences compared with Horse heart cytochrome c (cyt-c)/cytochrome c oxidase (CcO) system, where the recognition between partners and the electron transfer (ET) process is initiated via electrostatic interactions. We demonstrated in a previous study by surface-enhanced resonance Raman (SERR) spectroscopy that roughened silver electrodes coated with uncharged mixed self-assembled monolayers HS–(CH2) n –CH3/HS–(CH2) n + 1–OH 50/50, n = 5, 10 or 15, was a good model to mimic the Cyt-c 552 redox partner. All the adsorbed molecules are well oriented on such biomimetic electrodes and transfer one electron during the redox process. The present work focuses on the kinetic part of the heterogeneous ET process of Cyt-c 552 adsorbed onto electrodes coated with such mixed SAMs of different alkyl chain length. For that purpose, two complementary methods were combined. Firstly cyclic voltammetry shows that the ET between the adsorbed Cyt-c 552 and the biomimetic electrode is direct and reversible. Furthermore, it allows the estimation of both the density surface coverage of adsorbed Cyt-c 552 and the kinetic constants values. Secondly, time-resolved SERR (TR-SERR) spectroscopy showed that the ET process occurs without conformational change of the Cyt-c 552 heme group and allows the determination of kinetic constants. Results show that the kinetic constant values obtained by TR-SERR spectroscopy could be compared to those obtained from cyclic voltammetry. They are estimated at 200, 150 and 40 s−1 for the ET of Cyt-c 552 adsorbed onto electrodes coated with mixed SAMs HS–(CH2) n –CH3/HS–(CH2) n + 1–OH 50/50, n = 5, 10 or 15, respectively. Presented at the joint biannual meeting of the SFB-GEIMM-GRIP, Anglet France, 14–19 October, 2006.  相似文献   
285.
High energy density, fast recharging ability, and sustained cycle life are the primary requisite of supercapacitors (SCs); these necessities can be fulfilled by engineering a smart current collector with hierarchical combination of different active materials. This study reports a multicomponent design of hierarchical zinc cobalt sulfide (ZCS) hollow nanotube arrays wrapped with interlaced ultrathin Ni(OH)2 nanoflakes for high‐performance electrodes. The ZCS exhibits a unique pentagonal cross‐section and a rough surface that facilitates the deposition of Ni(OH)2 nanoflakes with a thickness of 7.5 nm. The ZCS/Ni(OH)2 hierarchical electrode exhibits a high specific capacitance of 2156 F g?1 and excellent cyclic stability with 94% retention over 3000 cycles. This is attributed to enhanced redox reactions, the direct growth of arrays on 3D porous foam acting as a “superhighway” for electron transport, and the increased availability of electrochemical active sites provided by the ultrathin Ni(OH)2 flakes that also sustain the stability of the electrode by sacrificing themselves during long charge/discharge cycles. Symmetric SCs are assembled to achieve high energy density of 74.93 W h kg?1 and exhibit superior cyclic stability of 78% retention with 81% coulombic efficiency over 10 000 cycles.  相似文献   
286.
This study presents a battery concept with a “mediator‐ion” solid electrolyte for the development of next‐generation electrochemical energy storage technologies. The active anode and cathode materials in a single cell can be in the solid, liquid, or gaseous form, which are separated by a sodium‐ion solid‐electrolyte separator. The uniqueness of this mediator‐ion strategy is that the redox reactions at the anode and the cathode are sustained by a shuttling of a mediator sodium ion between the anolyte and the catholyte through the solid‐state electrolyte. Use of the solid‐electrolyte separator circumvents the chemical‐crossover problem between the anode and the cathode, overcomes the dendrite‐problem when employing metal‐anodes, and offers the possibility of using different liquid electrolytes at the anode and the cathode in a single cell. The battery concept is demonstrated with two low‐cost metal anodes (zinc and iron), two liquid cathodes (bromine and potassium ferricyanide), and one gaseous cathode (air/O2) with a sodium‐ion solid electrolyte. This novel battery strategy with a mediator‐ion solid electrolyte is applicable to a wide range of electrochemical energy storage systems with a variety of cathodes, anodes, and mediator‐ion solid electrolytes.  相似文献   
287.
The selectivity of electrodes of solar cells is a critical factor that can limit the overall efficiency. If the selectivity of an electrode is not sufficient both electrons and holes recombine at its surface. In materials with poor transport properties such as in organic solar cells, these surface recombination currents are accompanied by large gradients of the quasi‐Fermi energies as the driving force. Experimental results from current–voltage characteristics, advanced photo‐ and electroluminescence as well as charge extraction of three different photoactive materials are shown and compared to drift‐diffusion simulations. It can be concluded that in cases of electrodes with reduced selectivity the decrease of the open‐circuit voltage can be divided into two distinct contributions, the reduction of the overall steady‐state charge carrier density and the gradients of the quasi‐Fermi energies. The results clearly show that for photoactive layers with poor transport properties, the gradient of the quasi‐Fermi energy in the vicinity of the contact is the main contribution to the loss in open‐circuit voltage. For imbalanced mobilities, this gives rise to the phenomenon that it is more challenging to realize a selective contact for the less mobile charge carrier, i.e., the hole contact in most organic solar cells.  相似文献   
288.
Na‐ion technology is increasingly studied as a low‐cost solution for grid storage applications. Many positive electrode materials have been reported, mainly among layered oxides and polyanionic compounds. The vanadium oxy/flurophosphate solid solution Na3V2(PO4)2F3‐y O2y (0 ≤ y ≤ 1), in particular, has proven the ability to deliver ≈500 Wh kg‐1, operating on the V3+/V4+ (y = 0) or V4+/V5+ redox couples (y = 1). This paper reports here on a significant increase in specific energy by enabling sodium insertion into Na3V2(PO4)2FO2 to reach Na4V2(PO4)2FO2 upon discharge. This occurs at ≈1.6 V and increases the theoretical specific energy to 600 Wh kg?1, rivaling that of several Li‐ion battery cathodes. This improvement is achieved by the judicious modification of the composition either as O for F substitution, or Al for V substitution, both of which disrupt Na‐ion ordering and thereby enable insertion of the 4th Na. This paper furthermore shows from operando X‐Ray Diffraction (XRD) that this energy is obtained in the cycling range Na4V2(PO4)2FO2–NaV2(PO4)2FO2 with a very small overall volume change of 1.7%, which is one of the smallest volume changes for Na‐ion cathodes and which is a crucial requisite for stable long‐term cycling.  相似文献   
289.
290.
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