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   25篇
  免费   1篇
化学   13篇
力学   4篇
物理学   9篇
  2012年   1篇
  2011年   1篇
  2010年   3篇
  2008年   3篇
  2007年   1篇
  2006年   2篇
  2005年   4篇
  2004年   1篇
  2002年   4篇
  2000年   1篇
  1997年   2篇
  1996年   1篇
  1983年   1篇
  1982年   1篇
排序方式: 共有26条查询结果,搜索用时 0 毫秒
21.
A compact broadband vibrational sum-frequency spectroscopy (SFG) apparatus is described to study molecules at surfaces and interfaces. Using an étalon as the frequency narrowing device, the visible pulse has a time-asymmetric profile that allows the user to deeply suppress nonresonant background signals that hinder detection of molecular vibrational resonances. Several features of the spectrometer that, in aggregate, improve signal-to-noise ratios by a large factor are described. The spectrometer features a series of interchangeable prealigned sample holders for different applications. Examples of applications are presented where nonresonant suppression greatly improves the ability to study adsorbates on single-crystal surfaces as a function of rotation about the azimuth, and where the rapid data acquisition abilities of the spectrometer are used to study electrochemical transformations on single-crystal electrodes.  相似文献   
22.
We have observed long-lived OH-stretch (nu(OH)) excitations (v = 1) in water during ultrafast laser ablation by a mid-infrared pulse tuned to the nu(OH) absorption maximum. The spectrum of excitations is measured using incoherent anti-Stokes Raman spectroscopy. Relative to the equilibrium water spectrum, these excitations evidence a narrowed (100 cm(-1) fwhm) and blue-shifted (3600 cm(-1) peak) transition. The excited-state lifetime is T1 > 200 ps, compared to 0.2 ps in bulk water. In the early stages of the ablation process, the water mean density decreases rapidly, which breaks up the hydrogen bonding. The long-lived species is attributed to nu(OH) excitations on water molecules associated with interfaces, having broken hydrogen bonds which cannot be rapidly reformed as in the liquid state.  相似文献   
23.
The potential-dependent (2x2)-3CO-->(radical19x radical19)R23.4 degrees-13CO adlayer phase transition on Pt(111) with 0.1M H(2)SO(4) electrolyte was studied using femtosecond broadband multiplex sum frequency generation (SFG) spectroscopy combined with linear scan voltammetry. Across the phase boundary the SFG atop intensity jumps, and at the same time the SFG spectrum of threefold CO sites is transformed into a bridge site spectrum with a small decrease in integrated SFG intensity. The SFG atop intensity jump and three fold-to-bridge intensity drop are noticeably different from what would be expected for these structures on the basis of coverage alone. This occurs because the SFG signal is sensitive to both the coverage and changes in the local field that result from a changing adlayer structure. We derive an equation that allows us to correct the SFG intensities for these effects using information derived from infrared absorption-reflection spectroscopy (IRAS) and second-harmonic generation (SHG) measurements. With this correction, the SFG results agree well with what would be expected for a transition between perfect adlattices. A small (approximately 20%) discrepancy in the SFG determination of atop coverage is attributed to either a small amount of surface disorder or uncertainties in the SFG, SHG, and IRAS measurements. SFG is also used to examine the reversibility hysteresis and kinetics of the phase transition and its dependence on electrolyte composition. The phase transition is reversible with an approximately 150 mV anodic overpotential and the forward (2x2)-->(radical19x radical19) transition is slower than the reverse. Repeated cycles of phase transition indicate that the 25 microm electrolyte layer used here does not appreciably distort the potential-coverage relationships.  相似文献   
24.
Simulations of self-assembled monolayers (SAMs) are performed to interpret experimental measurements of ultrafast approximately 1 GPa (volume compression deltaV approximately 0.1) planar shock compression dynamics probed by vibrational sum-frequency generation (SFG) spectroscopy (Lagutchev, A. S.; Patterson, J. E.; Huang, W.; Dlott, D. D. J. Phys. Chem. B 2005, 109, XXXX). The SAMs investigated are octadecanethiol (ODT) and pentadecanethiol (PDT) on Au(111) and Ag(111) substrates, and benzyl mercaptan (BMT) on Au(111). In the alkane SAMs, SFG is sensitive to the instantaneous orientation of the terminal methyl; in BMT it is sensitive to the phenyl orientation. Computed structures of alkane SAMs are in good agreement with experiment. In alkanes, the energies of gauche defects increase with increasing number and depth below the methyl plane, with the exception of ODT/Au where both single and double gauche defects at the two uppermost dihedrals have similar energies. Simulations of isothermal uniaxial compression of SAM lattices show that chain and methyl tilting is predominant in PDT/Au, ODT/Ag and PDT/Ag, whereas single and double gauche defect formation is predominant in ODT/Au. Time-resolved shock data showing transient SFG signal loss of ODT/Au and PDT/Au are fit by calculations of the terminal group orientations as a function of deltaV and their contributions to the SFG hyperpolarizability. The highly elastic response of PDT/Au results from shock-generated methyl and chain tilting. The viscoelastic response of ODT/Au results from shock generation of single and double gauche defects. Isothermal compression simulations help explain and fit the time dependence of shock spectra but generally underestimate the magnitude of SFG signal loss because they do not include effects of high-strain-rate dynamics and shock front and surface irregularities.  相似文献   
25.
Deàk JC  Iwaki LK  Dlott DD 《Optics letters》1997,22(23):1796-1798
A high-power (50-MW), kilohertz, picosecond, mid-IR optical parametric amplifier that is pumped by an amplified Ti:sapphire laser and also produces a fixed-frequency visible pulse is described. Mid-IR pulse energies of 40-55 microJ with 0.6-0.8-ps durations and 35-cm (-1) bandwidths are reported in the 3650- 2800-cm (-1) range. The combination of picosecond mid-IR and visible pulses is useful for two-color spectroscopies, which require simultaneous time and frequency resolution. To illustrate the above, we present vibrational relaxation data for the polyatomic molecule nitromethane, using time-resolved infrared Raman spectroscopy.  相似文献   
26.
A method employing photochemical hole burning, previously developed to measure the distribution of Raman enhancement factors on a nanostructured substrate for surface‐enhanced Raman scattering, is used to compare the enhancement distributions of benzenethiol adsorbed on substrates optimized for 532 nm laser excitation consisting of close‐packed (CP) or nonclose‐packed (NCP) nanospheres. The ensemble‐averaged Raman enhancement factor was 2.8 times smaller for the NCP substrate. The measured distributions revealed additional information. For instance, 92% of the molecules on the CP substrate and 93.6% of the molecules on the NCP substrate had Raman enhancements below average. The minimum enhancements on both substrates were ~104, but on the NCP substrate the maximum enhancement was 1.2 × 108, whereas on the CP substrate the maximum was 2 × 1010. The Ag‐coated nanospheres form hemisphere‐on‐cylinder mushroom‐like structures on both lattices, but on the NCP lattice, one third of the molecules are on the flat regions between the mushrooms. The flats on the NCP lattice have enhancements of ~104, showing they are part of a resonant plasmonic structure. The highest NCP enhancements of ~108 are tentatively associated with regions at the bases of the mushrooms, whereas the highest CP enhancements of 2 × 1010 are tentatively associated with gaps between nanospheres where 0.0025% of the molecules reside. Copyright © 2011 John Wiley & Sons, Ltd.  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