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   1003篇
  免费   70篇
  国内免费   37篇
儿科学   1篇
妇产科学   2篇
基础医学   137篇
口腔科学   20篇
临床医学   54篇
内科学   266篇
皮肤病学   14篇
神经病学   33篇
特种医学   86篇
外国民族医学   1篇
外科学   29篇
综合类   119篇
预防医学   42篇
眼科学   24篇
药学   181篇
中国医学   61篇
肿瘤学   40篇
  2023年   15篇
  2022年   37篇
  2021年   78篇
  2020年   40篇
  2019年   35篇
  2018年   23篇
  2017年   26篇
  2016年   49篇
  2015年   63篇
  2014年   80篇
  2013年   100篇
  2012年   80篇
  2011年   83篇
  2010年   61篇
  2009年   56篇
  2008年   51篇
  2007年   41篇
  2006年   24篇
  2005年   25篇
  2004年   17篇
  2003年   13篇
  2002年   20篇
  2001年   13篇
  2000年   7篇
  1999年   5篇
  1998年   10篇
  1997年   12篇
  1996年   4篇
  1995年   9篇
  1994年   3篇
  1993年   8篇
  1992年   5篇
  1991年   2篇
  1990年   2篇
  1989年   1篇
  1988年   1篇
  1987年   1篇
  1986年   1篇
  1984年   1篇
  1983年   2篇
  1982年   1篇
  1981年   2篇
  1979年   1篇
  1978年   2篇
排序方式: 共有1110条查询结果,搜索用时 468 毫秒
31.
32.
The first step of photosynthesis in plants is the absorption of sunlight by pigments in the antenna complexes of photosystem II (PSII), followed by transfer of the nascent excitation energy to the reaction centers, where long-term storage as chemical energy is initiated. Quantum mechanical mechanisms must be invoked to explain the transport of excitation within individual antenna. However, it is unclear how these mechanisms influence transfer across assemblies of antenna and thus the photochemical yield at reaction centers in the functional thylakoid membrane. Here, we model light harvesting at the several-hundred-nanometer scale of the PSII membrane, while preserving the dominant quantum effects previously observed in individual complexes. We show that excitation moves diffusively through the antenna with a diffusion length of 50 nm until it reaches a reaction center, where charge separation serves as an energetic trap. The diffusion length is a single parameter that incorporates the enhancing effect of excited state delocalization on individual rates of energy transfer as well as the complex kinetics that arise due to energy transfer and loss by decay to the ground state. The diffusion length determines PSII’s high quantum efficiency in ideal conditions, as well as how it is altered by the membrane morphology and the closure of reaction centers. We anticipate that the model will be useful in resolving the nonphotochemical quenching mechanisms that PSII employs in conditions of high light stress.The first step of photosynthesis is light harvesting, the absorption and conversion of sunlight into chemical energy. In photosynthetic organisms, the functional units of light harvesting are self-assembled arrays of pigment–protein complexes called photosystems. Antenna complexes absorb and transfer the nascent excitation energy to reaction centers, where long-term storage as chemical energy is initiated (1). In plants, photosystem II (PSII) flexibly responds to changes in sunlight intensity on a seconds to minutes time scale. In dim light, under ideal conditions, PSII harvests light with a >80% quantum efficiency (2), whereas, in intense sunlight PSII dissipates excess absorbed light safely as heat via nonphotochemical quenching pathways (3). The ability of PSII to switch between efficient and dissipative states is important for optimal plant fitness in natural sunlight conditions (4). Understanding how PSII’s function arises from the structure of its constituent pigment−protein complexes is a prerequisite for systematically engineering the light-harvesting apparatus in crops (57) and could be useful for designing artificial materials with the same flexible properties (8, 9).Recent advances have established structure−function relationships within individual pigment−protein complexes, but not how these relationships affect the functioning of the dynamic PSII (grana) membrane (10). Electron microscopy and fitting of atomic resolution structures (11) place the pigment−protein complexes in the grana membrane in close proximity, enabling long-range transport. Indeed, connectivity of excitation between different PSII reaction centers has been discussed since 1964 (12), suggesting that the functional unit for PSII must involve a large area of the membrane. Two limiting cases have been used to model PSII light harvesting: The lake model assumes perfect connectivity between reaction centers across the membrane; alternatively, the membrane can be described as a collection of disconnected “puddles” of pigments that each contain one reaction center (1, 13). At present, however, resolving the spatiotemporal dynamics within the grana membrane on the relevant length (tens to hundreds of nanometers) and time (1 ps to 1 ns) scales experimentally is not possible. Structure-based modeling of the grana membrane, however, can access this wide range of length and time scales.The dense packing of the major light-harvesting antenna (LHCII, discs), which is a trimeric complex, and PSII supercomplexes (PSII-S, pills) in the grana membrane is shown in Fig. 1 A and B. PSII-S is a multiprotein complex (14) that contains the PSII core reaction center dimer, along with several minor light-harvesting complexes and LHCII (Fig. 1A, Inset). Electronic excited states in LHCII and PSII-S are delocalized over several pigments (1517), making conventional Förster theory inadequate to describe the excitation dynamics. On the protein length scale, generalized Förster (18, 19) calculations between domains of tightly coupled chlorophylls agree very well with more exact methods [e.g., the zeroth-order functional expansion of the quantum-state diffusion model (ZOFE) approximation to non-Markovian quantum state diffusion (20)] for simulating the excitation population dynamics (21, 22). This agreement suggests that the primary quantum phenomenon involved in PSII energy transfer is the site basis coherence that arises from excited states delocalized across a few (approximately three to four) pigments.Open in a separate windowFig. 1.Accurate simulation of chlorophyll fluorescence dynamics from thylakoid membranes using structure-based modeling of energy transfer in PSII. (A and B) The representative mixed (A) and segregated (B) membrane morphologies generated using Monte Carlo simulations and used throughout this work. PSII-S are indicated by the light teal pills, and LHCII, which are trimeric complexes, are indicated by the light grey-green circles. The segregated membrane forms PSII-S arrays and LHCII pools. As shown schematically in A (Bottom) existing crystal structures of PSII-S (14) and LHCII (24) were overlaid on these membrane patches to establish the locations of all chlorophyll pigments. The light teal and light grey-green dashed lines outline the excluded area associated with PSII-S and LHCII trimers respectively, in the Monte Carlo simulations. The chlorophyll pigments are indicated in green, and the protein is depicted by the grey cartoon ribbon. PSII-S is a twofold symmetric dimer of pigment−protein complexes that are outlined by black lines. LHCII-S (strongly bound LHCII), CP26, CP29, CP43, and CP47 are antenna proteins, and RC indicates the reaction center. The inhomogeneously averaged rates of energy transfer between strongly coupled clusters of pigments were calculated using generalized Förster theory. (C) Simulated fluorescence decay of the mixed membrane (solid black line) and the PSII component of experimental fluorescence decay data from thylakoid membranes from ref. 26 (red, dotted line). Inset shows the lifetime components and amplitudes of the simulated decay as calculated using our model with a Gaussian convolution (σ = 20 ps) (black line) or by fitting to three exponential decays (green bars).Here, we construct a generalized Förster model for the ∼104 pigments covering the few-hundred-nanometer length scale of the grana membrane that correctly incorporates the dynamics occurring within and between complexes on the picosecond time scale. We show how delocalized excited states, or excitons, in individual complexes affect light harvesting on the membrane length scale. The formation of excitons is sufficient to explain the high quantum efficiency of PSII in dim light. The model, by being an accurate representation of the complex kinetic network that underlies PSII light harvesting, provides mechanistic explanations for long-observed biological phenomena and sets the stage for developing a better understanding of PSII light harvesting in high light conditions.  相似文献   
33.
首次以柠檬酸为碳源、甘氨酸为修饰剂,通过高温热解法一步合成修饰碳纳米粒。经过正丁醇萃取纯化,碳纳米粒的粒径更加均一,荧光强度更大,且性能得到了改善。最终制得的碳纳米粒呈棕黄色,在380 nm激发波长下,最大发射波长出现在480 nm附近,其荧光量子产率高达47%。氯霉素对碳纳米粒的荧光具有显著的猝灭效应,并呈现一定的规律性,据此建立了对氯霉素含量测定的新方法。该方法简便快捷,易于操作,线性关系良好(r=0.999 7),回收率在99%~101%(RSD=0.3%),显示了碳纳米粒在药物检测方面的潜在应用前景。  相似文献   
34.
This perspective discusses progress in the theory of bimolecular reaction dynamics in the gas phase. The examples selected show that definitive quantum dynamical computations are providing insights into the detailed mechanisms of chemical reactions.  相似文献   
35.
目的 构建一种基于新型纳米材料量子点的流式微球技术,用于高通量的DNA检测,实现DNA的快速、低成本检测。方法 将能与靶DNA一端杂交的探针DNA(P1)标记在微球表面,与DNA配对杂交后,加入能与靶DNA另一端杂交的量子点标记的探针DNA(P2),组装成一种流式微球-探针P1、靶DNA、量子点-探针P2的夹心复合结构,通过测定杂交前后平均荧光强度的变化检测DNA的浓度。结果 该新型检测方法可以区分出完全互补、单碱基错配及完全非互补的DNA(P?<0.05),且随着DNA浓度的升高,检测到的平均荧光强度逐渐增强(P?<0.05),对DNA的最低检出限可达0.2?nmol/L。结论 新型量子点流式微球技术检测DNA具有高敏感性和高特异性、分析速度快、操作简单等优点,是一种非常重要的具有潜力的新型临床检测方法。  相似文献   
36.
We report the observation of anomalously robust valley polarization and valley coherence in bilayer WS2. The polarization of the photoluminescence from bilayer WS2 follows that of the excitation source with both circular and linear polarization, and remains even at room temperature. The near-unity circular polarization of the luminescence reveals the coupling of spin, layer, and valley degree of freedom in bilayer system, and the linearly polarized photoluminescence manifests quantum coherence between the two inequivalent band extrema in momentum space, namely, the valley quantum coherence in atomically thin bilayer WS2. This observation provides insight into quantum manipulation in atomically thin semiconductors.Tungsten sulfide WS2, part of the family of group VI transition metal dichalcogenides (TMDCs), is a layered compound with buckled hexagonal lattice. As WS2 thins to atomically thin layers, WS2 films undergo a transition from indirect gap in bulk form to direct gap at monolayer level with the band edge located at energy-degenerate valleys (K, K′) at the corners of the Brillouin zone (13). Like the case of its sister compound, monolayer MoS2, the valley degree of freedom of monolayer WS2 could be presumably addressed through nonzero but contrasting Berry curvatures and orbital magnetic moments that arise from the lack of spatial inversion symmetry at monolayers (3, 4). The valley polarization could be realized by the control of the polarization of optical field through valley-selective interband optical selection rules at K and K′ valleys as illustrated in Fig. 1A (46). In monolayer WS2, both the top of the valence bands and the bottom of the conduction bands are constructed primarily by the d orbits of tungsten atoms, which are remarkably shaped by spin–orbit coupling (SOC). The giant spin–orbit coupling splits the valence bands around the K (K′) valley by 0.4 eV, and the conduction band is nearly spin degenerated (7). As a result of time-reversal symmetry, the spin splitting has opposite signs at the K and K′ valleys. Namely, the Kramer’s doublet |K ↑ ? and |K′ ↓ ? is separated from the other doublet |K′ ↑ ? and |K ↓ ? by the SOC splitting of 0.4 eV. The spin and valley are strongly coupled at K (K′) valleys, and this coupling significantly suppresses spin and valley relaxations as both spin and valley indices have to be changed simultaneously.Open in a separate windowFig. 1.(A) Schematic of valley-dependent optical selection rules and the Zeeman-like spin splitting in the valence bands of monolayer WS2. (B) Diagram of spin–layer–valley coupling in 2H stacked bilayer WS2. Interlayer hopping is suppressed in bilayer WS2 owing to the coupling of spin, valley, and layer degrees of freedom.In addition to the spin and valley degrees of freedom, in bilayer WS2 there exists an extra index: layer polarization that indicates the carriers’ location, either up-layer or down-layer. Bilayer WS2 follows the Bernal packing order and the spatial inversion symmetry is recovered: each layer is 180° in plane rotation of the other with the tungsten atoms of a given layer sitting exactly on top of the S atoms of the other layer. The layer rotation symmetry switches K and K′ valleys, but leaves the spin unchanged, which results in a sign change for the spin–valley coupling from layer to layer (Fig. 1B). From the simple spatial symmetry point of view, one might expect that the valley-dependent physics fades at bilayers owing to inversion symmetry, as the precedent of bilayer MoS2 (8). Nevertheless, the inversion symmetry becomes subtle if the coupling of spin, valley, and layer indices is taken into account. Note that the spin–valley coupling strength in WS2 is around 0.4 eV (the counterpart in MoS2 ∼ 0.16 eV), which is significantly higher than the interlayer hopping energy (∼0.1 eV); the interlayer coupling at K and K′ valleys in WS2 is greatly suppressed as indicated in Fig. 1B (7, 9). Consequently, bilayer WS2 can be regarded as decoupled layers and it may inherit the valley physics demonstrated in monolayer TMDCs. In addition, the interplay of spin, valley, and layer degrees of freedom opens an unprecedented channel toward manipulations of quantum states.Here we report a systemic study of the polarization-resolved photoluminescence (PL) experiments on bilayer WS2. The polarization of PL inherits that of excitations up to room temperature, no matter whether it is circularly or linearly polarized. The experiments demonstrate the valley polarization and valley coherence in bilayer WS2 as a result of the coupling of spin, valley, and layer degrees of freedom. Surprisingly, the valley polarization and valley coherence in bilayer WS2 are anomalously robust compared with monolayer WS2.For comparison, we first perform polarization-resolved photoluminescence measurements on monolayer WS2. Fig. 2A shows the photoluminescence spectrum from monolayer WS2 at 10 K. The PL is dominated by the emission from band-edge excitons, so-called “A” exciton at K and K′ valleys. The excitons carry a clear circular dichroism under near-resonant excitation (2.088 eV) with circular polarization as a result of valley-selective optical selection rules, where the left-handed (right handed) polarization corresponds to the interband optical transition at K (K′) valley. The PL follows the helicity of the circularly polarized excitation optical field. To characterize the polarization of the luminescence spectra, we define a degree of circular polarization as P=I(σ+)I(σ)I(σ+)+I(σ), where I(σ±) is the intensity of the right- (left-) handed circular-polarization component. The luminescence spectra display a contrasting polarization for excitation with opposite helicities: P = 0.4 under σ+ excitation and P = −0.4 under σ− excitation on the most representative monolayer. For simplicity, only the PL under σ+ excitation is shown. The degree of circular polarization P is insensitive to PL energy throughout the whole luminescence as shown in Fig. 2A, Inset. These behaviors are fully expected in the mechanism of valley-selective optical selection rules (3, 4). The degree of circular polarization decays with increasing temperature and drops to 10% at room temperature (Fig. 2B). It decreases as the excitation energy shifts from the near-resonance energy of 2.088 to 2.331 eV as illustrated in Fig. 2C. The peak position of A exciton emission at band edges shifts from 2.04 eV at 10 K to 1.98 eV at room temperature. The energy difference between the PL peak and the near-resonance excitation (2.088 eV) is around 100 meV at room temperature, which is much smaller than the value 290 meV for the low temperature off-resonance excitation at 2.331 eV. However, the observed polarization for off-resonance excitation at 10 K (P = 16%) is much higher than the near-resonance condition at room temperature (P = 10%). It clearly shows that the depolarization cannot be attributed to single process, namely the off-resonance excitation or band-edge phonon scattering only (10).Open in a separate windowFig. 2.Photoluminescence of monolayer WS2 under circularly polarized excitation. (A) Polarization resolved luminescence spectra with σ+ detection (red) and σ− detection (black) under near-resonant σ+ excitation (2.088 eV) at 10 K. Peak A is the excitonic transition at band edges of K (K′) valleys. Opposite helicity of PL is observed under σ− excitation. Inset presents the degree of the circular polarization at the prominent PL peak. (B) The degree of the circular polarization as a function of temperature. The curve (red) is a fit following a Boltzmann distribution where the intervalley scattering by phonons is assumed. (C) Photoluminescence spectrum under off-resonant σ+ excitation (2.33 eV) at 10 K. The red (black) curve denotes the PL circular components of σ+ (σ−).Next we study the PL from bilayer WS2. Fig. 3 shows the PL spectrum from bilayer WS2. The peak labeled as “I” denotes the interband optical transition from the indirect band gap, and the peak A corresponds to the exciton emission from direct band transition at K and K′ valleys. Although bilayer WS2 has an indirect gap, the direct interband optical transition at K and K′ valleys dominates the integrated PL intensity as the prerequisite of phonon/defect scattering is waived for direct band emission and the direct gap is just slightly larger than the indirect band gap in bilayers. Fig. 3A displays surprisingly robust PL circular dichroism of A exciton emission under circularly polarized excitations of 2.088 eV (resonance) and 2.331 eV (off resonance). The degree of circular polarization of A exciton emission under near-resonant σ± excitation is near unity (around 95%) at 10 K and preserves around 60% at room temperature. In contrast, the emission originating from indirect band gap is unpolarized in all experimental conditions.Open in a separate windowFig. 3.Photoluminescence of bilayer WS2 under circularly polarized excitations. (A) Polarization-resolved luminescence spectra with components of σ+ (red) and σ− (black) under near-resonant σ+ excitation (2.088 eV) at 10 K. Peak A is recognized as the excitonic transition at band edge of direct gap. Peak I originates from the indirect band-gap emission, showing no polarization. Inset presents the circular polarization of the A excitonic transition around the PL peak. Opposite helicity of PL is observed under σ− excitation. (B) The degree of circular polarization as a function of temperature (black). The curve (red) is a fit following a Boltzman distribution where the intervalley scattering by phonons is assumed. (C) Photoluminescence spectrum of components of σ+ (red) and σ− (black) under off-resonance σ+ excitation (2.33 eV) at 10 K. A nonzero circular polarization P is only observed at emissions from A excitons.To exclude the potential cause of charge trapping or substrate charging effect, we study the polarization-resolved PL of bilayer WS2 with an out-plane electric field. Fig. 4A shows the evolution of PL spectra in a field-effect-transistor-like device under circularly polarized excitations of 2.088 eV and an electric gate at 10 K. The PL spectra dominated by A exciton show negligible change under the gate bias in the range of −40 to 20 V. The electric-conductance measurements show that the bilayer WS2 stays at the electrically intrinsic state under the above bias range. The PL spectra can be safely recognized as emissions from free excitons. As the gate bias switches further to the positive side (>20 V), the PL intensity decreases, and the emission from electron-bounded exciton “X,” the so-called trion emerges and gradually raises its weight in the PL spectrum (11, 12). The electron–exciton binding energy is found to be 45 meV. Given only one trion peak in PL spectra, the interlayer trion (formed by exciton and electron/hole in different layers) and intralayer trion (exciton and electron/hole in the same layer) could not be distinguished due to the broad spectral width (13). Both the free exciton and trion show slight red shifts with negative bias, presumably as a result of quantum-confined stark effect (14). At all of the bias conditions, the degree of circular polarization of the free exciton and trion stays unchanged within the experiment sensitivity as shown in Fig. 4C.Open in a separate windowFig. 4.Electric-doping-dependent photoluminescence spectrum of bilayer WS2 field-effect transistor. (A) Luminescence spectra of bilayer WS2 at different gate voltage under near-resonant σ+ excitation (2.088 eV) at 10 K. X and X denote neutral exciton and trion, respectively. Green curve is a fitting consisting of two Lorentzian peak fits (peak I and X) and one Gaussian peak fit (peak X). (B) Intensity of exciton and trion emissions versus gate. (Upper) The gate-dependent integral PL intensity consisting of exciton (X) and trion (X). (Lower) The ratio of the integral PL intensity of exciton versus that of trion, as a function of the gate voltage. (C) Degree of circular polarization of exciton (X, red) and trion (X, blue) versus the gate.It is also unlikely that the high polarization in bilayers results from the isolation of the top layer from the environments, as similar behaviors are observed in monolayer and bilayer WS2 embedded in polymethyl methaccrylate (PMMA) matrix or capped with a 20-nm-thick SiO2 deposition. The insensitivity of the circular-polarization degree on bias and environments rules out the possibility that the effects of Coulomb screening, charge traps, or charge transfers with substrates are the major causes for the robust circular dichroism in bilayers against monolayers.One potential cause may result from the shorter lifetime of excitons at K (K′) valley for bilayer system. The band gap shifts from K and K′ points of the Brillouin zone in monolayers to the indirect gap between the top of the valence band at Γ points and the bottom of the conduction band in the middle of K and Γ points in bilayers. Combining our time-resolved pump-probe reflectance experiments (Supporting Information) and the observed relative PL strength between monolayer and bilayer (10:1), we infer the exciton lifetime at K (K′) valleys around 10 ps, a fraction of that at monolayers. If we assume (i) the PL circular polarization P=P01+2ττk, where P0 is the theoretical limit of PL polarization, and τk and τ denote the valley lifetime and exciton lifetime respectively; and (ii) the valley lifetime is the same for both monolayers and bilayers, the shorter exciton lifetime will lead to significantly higher PL polarization. However, the difference in exciton lifetime between bilayers and monolayers is not overwhelming enough to be the major cause of robust polarization observed in the time-integrated PL in bilayers.In monolayer WS2 under circularly polarized resonant excitations, the depolarization mainly comes from the K ? K′ intervalley scattering. In bilayers, the depolarization could be either via K ? K′ intervalley scattering within the layer in a similar way as in monolayers, or via interlayer hopping, which also requires spin flip. As we discussed above, the interlayer hopping at K valley is suppressed in WS2 as a result of strong SOC in WS2 and spin–layer–valley coupling, which were experimentally proved by the circular dichroism in PL from bilayers. The robust polarization in bilayers implies that the intervalley scattering within a layer is diminished compared with that in monolayers. There are two prerequisites for intervalley scattering within layers: conservation of crystal momentum and spin flip of holes. The crystal momentum conservation could be satisfied with the involvement of phonons at K points in the Brillouin zone or atomic size defects, presumably sharing the similar strength in monolayers and bilayers. Spin-flip process could be realized by three different spin scattering mechanisms, namely D’yakonov–Perel (DP) mechanism (15), Elliot–Yaffet (EY) mechanism (16), and Bir–Aronov–Pikus (BAP) mechanism (17, 18). The DP mechanism acts through a Lamor precession driven by electron wavevector k dependent spin–orbit coupling. It is thought to be negligible for spin flip along out-plane direction as the mirror symmetry with respect to the plane of W atoms secures a zero out-plane crystal electric field. Another possible driving force behind the DP mechanism could be the asymmetry owing to the interface with the substrate. This can be excluded by the similar behaviors, where the monolayers and bilayers WS2 are embedded in PMMA matrix or capped with a thin layer of SiO2. The negligible effect of electric gating on polarization also implies that the DP mechanism is weak in monolayer and bilayer WS2; the EY mechanism originates from scattering with phonons and defects. Its strength in bilayers and monolayers is likely to be at similar scale, and bilayers even have more low-frequency collective vibrational modes (19). Therefore, EY mechanism is unlikely to be the cause here; the BAP mechanism originates from the electron–hole exchange interaction. In monolayer and bilayer TMDCs, the optical features are dominated by the Wannier type, yet giant excitonic effect, and the exciton-binding energy in such intrinsic 2D semiconductors is estimated to be 0.6 ∼ 1 eV (20, 21). This giant exciton-binding energy indicates a mixture of electron and hole wavefunctions and, consequently, strong exchange interaction, which may contribute to the spin flip and intervalley scattering (5, 22). As the conduction band has a band mixing at K points, the spin flip of the electron would be a quick process. An analogous scenario is that the spin of holes relaxes in hundreds of femtoseconds or fewer in GaAs as a result of band mixing and spin–orbit coupling. The electron spin flip could lead to hole spin flip via strong exchange interaction accompanying intervalley scattering, which is realized by the virtual annihilation of a bright exciton in the K valley and then generation in the K′ valley or vice versa (22). This non-single-particle spin relaxation leads to valley depolarization instead of the decrease of luminescence intensity that results from coupling with dark excitons. Generally, the exciton-binding energy decreases with the relaxation of spatial confinement. However, first principle calculation shows that monolayer and bilayer WS2 share the similar band dispersion and effective masses around K valley in their Brillouin zone as a result of spin–valley coupling (7). It implies that the binding energy of excitons around K valley in bilayer WS2 is similar to or slightly less than that in monolayer WS2. As the exchange interaction is roughly proportional to the square of exciton binding energy, the spin-flip rate and consequently intervalley scattering via exciton exchange interactions is presumably comparable or smaller to some extent in bilayer WS2 (Supporting Information). Nevertheless, this is unlikely the major cause of the anomalously robust valley polarization in bilayer WS2.Another possibility includes extra spin-conserving channels via intermediate intervalley-interlayer scatterings in bilayer WS2, which are absent in monolayers (23). The extra spin-conserving channel may compete with the spin-flip process and reduce the relative weight of spin-flip intervalley scattering to some extent. However, the mechanism and the strength are unclear so far. Overall, the robust circular polarization in bilayers likely results from combined effects of the shorter exciton lifetime, smaller exciton-binding energy, extra spin-conserving channels, and the coupling of spin, layer, and valley degrees of freedom, indicating the relatively weak intervalley scattering in bilayer system. Further quantitative study is necessary to elaborate the mechanism.We also investigated the PL from bilayer WS2 under a linearly polarized excitation. A linearly polarized light could be treated as a coherent superposition of two opposite-helicity circularly polarized lights with a certain phase difference. The phase difference determines the polarization direction. In semiconductors, a photon excites an electron–hole pair with the transfer of energy, momentum, and phase information. The hot carriers energetically relax to the band edge in a quick process around 10−1 ∼ 101 ps through runs of inelastic and elastic scatterings, e.g., by acoustic phonons. During the quick relaxation process, generally the phase information randomizes and herein coherence fades. In monolayer TMDCs, the main channel for carrier relaxation is through intravalley scatterings including Coulomb interactions with electron (hole) and inelastic interactions with phonons, which are valley independent and preserve the relative phase between K and K′ valleys (24). In bilayer WS2, the suppression of intervalley scattering consequently leads to the suppression of inhomogeneous broadening in carrier’s phase term. Subsequently, the valley coherence demonstrated in monolayer WSe2 (24) is expected to be enhanced in bilayers (13). The valley coherence in monolayer and bilayer WS2 could be monitored by the polarization of PL under linearly polarized excitations.Fig. 5A shows the linearly polarized components of PL under a linearly polarized excitation of 2.088 eV at 10 K. The emission from indirect band gap is unpolarized and A exciton displays a pronounced linear polarization following the excitation. The degree of linear polarization P=I()I()I()+I() is around 80%, where I()(I(⊥)) is the intensity of PL with parallel (perpendicular) polarization with respect to the excitation polarization. In contrast, the linear polarization is much weaker in monolayer samples (4% under the same experimental conditions, as shown in Fig. 5B). As presented in Fig. 5C, the polarization of A exciton is independent of crystal orientation and exactly follows the polarization of excitations. The degree of the linear polarization in bilayer WS2 slightly decreases with the increased temperature and drops from 80% at 10 K to 50% at room temperature (Fig. 5D). This is the paradigm of the robust valley coherency in bilayer WS2.Open in a separate windowFig. 5.Linearly polarized excitations on monolayer and bilayer WS2. (A) Linear-polarization-resolved luminescence spectra of bilayer WS2 under near-resonant linearly polarized excitation (2.088 eV) at 10 K. Red (black) presents the spectrum with parallel (cross) polarization with respect to the linear polarization of excitation source. A linear polarization of 80% is observed for exciton A, and the indirect gap transition (I) is unpolarized. (B) Linear-polarization-resolved luminescence spectra of monolayer WS2 under near-resonant linearly polarized excitation (2.088 eV) at 10 K. Red (black) denotes the spectrum with the parallel (cross) polarization with respect to the linear polarization of excitation source. The linear polarization for exciton A in monolayer WS2 is much weaker, with a maximum value of 4%. (C) Polar plot for intensity of the exciton A in bilayer WS2 (black) as a function of the detection angle at 10 K. Red curve is a fit-following cos2(θ). (D) The degree of linear polarization of exciton A in bilayer WS2 (black) as a function of temperature. The curve (red) is a fit following a Boltzmann distribution where the intervalley scattering by phonons is assumed. (E) Electric doping dependence of the linear polarization of exciton A in bilayer WS2 at 10 K.The linear polarization of both exciton and trion in bilayer, contrasting to the circular polarization, which is insensitive to the electric field in the range, shows a weak electric gating dependence as shown in Fig. 5E. The PL linear polarization, presenting valley coherence, decreases as the Fermi level shifts to the conduction band. It does not directly affect intervalley scattering within individual layers and makes negligible change in circular dichroism. Nevertheless, the electric field between the layers induces a layer polarization and slightly shifts the band alignments between the layers by different amounts in conduction and valence bands (13, 25), although the shift is indistinguishable in the present PL spectra due to the broad spectral width. The layer polarization and the shift of band alignments may induce a relative phase difference between two layers and therefore affect the PL linear polarization via interference. Further study is needed to fully understand the mechanism.In summary, we demonstrated anomalously robust valley polarization and valley polarization coherence in bilayer WS2. The valley polarization and valley coherence in bilayer WS2 are the direct consequences of giant spin–orbit coupling and spin valley coupling in WS2. The depolarization and decoherence processes are greatly suppressed in bilayer, although the mechanism is ambiguous. The robust valley polarization and valley coherence make bilayer WS2 an intriguing platform for spin and valley physics.  相似文献   
37.
Since the pioneering works of Carr-Purcell and Meiboom-Gill [Carr HY, Purcell EM (1954) Phys Rev 94:630; Meiboom S, Gill D (1985) Rev Sci Instrum 29:688], trains of π-pulses have featured amongst the main tools of quantum control. Echo trains find widespread use in nuclear magnetic resonance spectroscopy (NMR) and imaging (MRI), thanks to their ability to free the evolution of a spin-1/2 from several sources of decoherence. Spin echoes have also been researched in dynamic decoupling scenarios, for prolonging the lifetimes of quantum states or coherences. Inspired by this search we introduce a family of spin-echo sequences, which can still detect site-specific interactions like the chemical shift. This is achieved thanks to the presence of weak environmental fluctuations of common occurrence in high-field NMR--such as homonuclear spin-spin couplings or chemical/biochemical exchanges. Both intuitive and rigorous derivations of the resulting "selective dynamical recoupling" sequences are provided. Applications of these novel experiments are given for a variety of NMR scenarios including determinations of shift effects under inhomogeneities overwhelming individual chemical identities, and model-free characterizations of chemically exchanging partners.  相似文献   
38.
Among the biological phenomena that fall within the emerging field of "quantum biology" is the suggestion that magnetically sensitive chemical reactions are responsible for the magnetic compass of migratory birds. It has been proposed that transient radical pairs are formed by photo-induced electron transfer reactions in cryptochrome proteins and that their coherent spin dynamics are influenced by the geomagnetic field leading to changes in the quantum yield of the signaling state of the protein. Despite a variety of supporting evidence, it is still not clear whether cryptochromes have the properties required to respond to magnetic interactions orders of magnitude weaker than the thermal energy, k(B)T. Here we demonstrate that the kinetics and quantum yields of photo-induced flavin-tryptophan radical pairs in cryptochrome are indeed magnetically sensitive. The mechanistic origin of the magnetic field effect is clarified, its dependence on the strength of the magnetic field measured, and the rates of relevant spin-dependent, spin-independent, and spin-decoherence processes determined. We argue that cryptochrome is fit for purpose as a chemical magnetoreceptor.  相似文献   
39.
In conventional spectroscopy, transitions between electronic levels are governed by the electric dipole selection rule because electric quadrupole, magnetic dipole, and coupled electric dipole-magnetic dipole transitions are forbidden in a far field. We demonstrated that by using nanostructured electromagnetic fields, the selection rules of absorption spectroscopy could be fundamentally manipulated. We also show that forbidden transitions between discrete quantum levels in a semiconductor nanorod structure are allowed within the near-field of a noble metal nanoparticle. Atomistic simulations analyzed by an effective mass model reveal the breakdown of the dipolar selection rules where quadrupole and octupole transitions are allowed. Our demonstration could be generalized to the use of nanostructured near-fields for enhancing light-matter interactions that are typically weak or forbidden.  相似文献   
40.
Water oxidation is a linchpin in solar fuels formation, and catalysis by single-site ruthenium complexes has generated significant interest in this area. Combining several theoretical tools, we have studied the entire catalytic cycle of water oxidation for a single-site catalyst starting with [RuII(tpy)(bpm)(OH2)]2+ (i.e., [RuII-OH2]2+; tpy is 2,2∶6,2′′-terpyridine and bpm is 2,2′-bypyrimidine) as a representative example of a new class of single-site catalysts. The redox potentials and pKa calculations for the first two proton-coupled electron transfers (PCETs) from [RuII-OH2]2+ to [RuIV = O]2+ and the following electron-transfer process to [RuV = O]3+ suggest that these processes can proceed readily in acidic or weakly basic conditions. The subsequent water splitting process involves two water molecules, [RuV = O]3+ to generate [RuIII-OOH]2+, and H3O+ with a low activation barrier (∼10 kcal/mol). After the key O---O bond forming step in the single-site Ru catalysis, another PECT process oxidizes [RuIII-OOH]2+ to [RuIV-OO]2+ when the pH is lower than 3.7. Two possible forms of [RuIV-OO]2+, open and closed, can exist and interconvert with a low activation barrier (< 7 kcal/mol) due to strong spin-orbital coupling effects. In Pathway 1 at pH = 1.0, oxygen release is rate-limiting with an activation barrier ∼12 kcal/mol while the electron-transfer step from [RuIV-OO]2+ to [RuV - OO]3+ becomes rate-determining at pH = 0 (Pathway 2) with Ce(IV) as oxidant. The results of these theoretical studies with atomistic details have revealed subtle details of reaction mechanisms at several stages during the catalytic cycle. This understanding is helpful in the design of new catalysts for water oxidation.  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