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   24572篇
  免费   2061篇
  国内免费   1745篇
耳鼻咽喉   195篇
儿科学   387篇
妇产科学   232篇
基础医学   2773篇
口腔科学   370篇
临床医学   3500篇
内科学   3228篇
皮肤病学   300篇
神经病学   1105篇
特种医学   816篇
外国民族医学   15篇
外科学   1927篇
综合类   4765篇
现状与发展   1篇
一般理论   2篇
预防医学   2458篇
眼科学   580篇
药学   2528篇
  24篇
中国医学   1465篇
肿瘤学   1707篇
  2024年   46篇
  2023年   286篇
  2022年   674篇
  2021年   973篇
  2020年   760篇
  2019年   606篇
  2018年   632篇
  2017年   667篇
  2016年   608篇
  2015年   1020篇
  2014年   1193篇
  2013年   1290篇
  2012年   1992篇
  2011年   2337篇
  2010年   1655篇
  2009年   1445篇
  2008年   1798篇
  2007年   1765篇
  2006年   1574篇
  2005年   1559篇
  2004年   919篇
  2003年   712篇
  2002年   562篇
  2001年   472篇
  2000年   523篇
  1999年   416篇
  1998年   280篇
  1997年   236篇
  1996年   180篇
  1995年   183篇
  1994年   166篇
  1993年   134篇
  1992年   127篇
  1991年   124篇
  1990年   89篇
  1989年   83篇
  1988年   66篇
  1987年   64篇
  1986年   51篇
  1985年   40篇
  1984年   15篇
  1983年   12篇
  1982年   9篇
  1981年   8篇
  1980年   3篇
  1979年   9篇
  1974年   2篇
  1973年   4篇
  1972年   2篇
  1965年   1篇
排序方式: 共有10000条查询结果,搜索用时 46 毫秒
991.
目的 分析Twin-block矫治器治疗生长发育期骨性Ⅱ类错儿童的临床疗效,探讨其机制。方法 选择生长发育期骨性Ⅱ类患者20例(男9例,女11例),进行Twin-block功能矫治。所有患者治疗前后拍摄X线头颅侧位片,应用Pancherz分析法测量患者治疗前后硬组织指标变化情况,同时,测量治疗前后的上气道线距变化,采用配对t检验检测治疗前后各指标差异。结果 Twin-block治疗前后,ss/OLP、pg/OLP明显增大, Co/OLP和前牙覆盖明显减小, Go-Me长度明显增加,磨牙关系明显改善,有统计学意义(P<0.05),其余各参数无统计学差异。气道测量值Adl-PNS、Ad2-PNS、Mcnamara线、U-MPW、TB-TPPW治疗后均有增大,差异有统计学意义(P<0.05)。结论 Twin-block功能矫治器能有效地改善骨性Ⅱ类患者的咬关系和侧貌,可以有效增加上气道各段前后径长度,这些变化主要由髁突前移导致的下颌骨生长发育所致。  相似文献   
992.
993.
Lung cancer development is associated with extensive pulmonary inflammation. In addition, the linkage between chronic obstructive pulmonary disease (COPD) and lung cancer has been demonstrated in population-based studies. IL-17–producing CD4 helper T cells (Th17 cells) play a critical role in promoting chronic tissue inflammation. Although Th17 cells are found in human COPD and lung cancer, their role is not understood. We have thus used a mouse model of lung cancer, in which an oncogenic form of K-ras (K-rasG12D), frequently found in human lung cancer, is restrictedly expressed in lung epithelial cells [via Clara cell secretory protein (CCSPcre)]. In this model, Th17 and Treg but not Th1 cells were found enriched at the tumor tissues. When CCSPcre/K-rasG12D mice were weekly challenged with a lysate of nontypeable Haemophilus influenza (NTHi), which induces COPD-type inflammation and accelerates the tumor growth, they showed greatly enhanced Th17 cell infiltration in the lung tissues. Lack of IL-17, but not IL-17F, resulted in a significant reduction in lung tumor numbers in CCSPcre/K-rasG12D mice and also those treated with NTHi. Absence of IL-17 not only resulted in reduction of tumor cell proliferation and angiogenesis, but also decreased the expression of proinflammatory mediators and reduced recruitment of myeloid cells. Depletion of Gr-1+CD11b+ myeloid cells in CCSPcre/K-rasG12D mice suppressed tumor growth in lung, indicating Gr-1+CD11b+ myeloid cells recruited by IL-17 play a protumor role. Taken together, our data demonstrate a critical role for Th17 cell-mediated inflammation in lung tumorigenesis and suggest a novel way for prevention and treatment of this disease.Inflammation plays an important role in tumor development (1, 2). Although targeting inflammation and tumor microenvironment has been considered as a new direction of cancer therapy, the mechanisms underlying cancer-associated inflammation have not been well understood. Lung cancer is a leading cause of death in the world. Accumulating evidence has shown that inflammation is associated with pathogenesis of lung cancer, especially those induced by cigarette smoke (3). The primary risk factor among smokers to develop lung cancer is the presence of chronic obstructive pulmonary disease (COPD) (4), which is characterized by chronic pulmonary inflammation, airway remodeling and destruction of lung parenchyma. Human lung cancers are inflicted with alterations in various subsets of lymphocytes and myeloid cells (5, 6), reminiscent of immune activation during chronic inflammation. Several studies have shown NFκB signaling as a mechanistic link between inflammation and lung cancer using a mouse model of lung adenocarcinoma (7, 8). However, the specific inflammatory cell types or molecules potentiating lung cancer are not understood clearly.We and others have identified a novel subset of CD4 helper T cells that produce IL-17 and are referred as Th17 cells (9, 10). Th17 cells have been associated with inflammatory diseases such as rheumatoid arthritis, asthma, lupus, and allograft rejection. An important function of IL-17 is to promote tissue inflammation through the up-regulation of proinflammatory cytokines and chemokines (11). Consistently, we have shown that transgenic overexpression of IL-17 in the lungs resulted in chemokine up-regulation and tissue infiltration by leukocytes, although mice treated with neutralizing IL-17–specific antibody were also found to be resistant to the induction of experimental autoimmune encephalomyelitis (9). These and other studies collectively demonstrated that IL-17 and Th17 cells play nonredundant function in promoting inflammation.Increased frequencies of IL-17 and Th17 cells have been reported in patients with different types of tumors (12), including lung adenocarcinoma (13). The density of intratumoral IL-17–positive cells in primary human nonsmall cell lung cancer was inversely correlated with patient outcome and correlated with smoking status of the patients (14). Th17 cells specific for a common tumor antigen were found in lung cancer patients as part of their spontaneous immune response to the autologous tumor (15). However, the function of Th17 cells and IL-17 in the development of lung cancer remains to be shown. Animal model studies have revealed contrasting roles of IL-17 in various tumors (16). Tumor-promoting effect of IL-17 was shown in some models such as colon cancer (1720), whereas in others, IL-17 supported anti-tumor immunity, including in B16 melanoma model (2124). Thus, the role of IL-17 could be complex and tumor-specific.To properly evaluate the role of IL-17 in inflammation-associated lung cancer, we used a model of oncogenic K-ras mutation expressed only in the lung. Mice expressing K-ras mutation in Clara cells (CCSPcre/K-rasG12D mice) spontaneously develop lung adenocarcinoma (25). In addition, we induced COPD-type lung inflammation by challenging mice with lysates of nontypeable Haemophilus influenza (NTHi). Inflammation driven by NTHi can promote tumor growth in CCSPcre/K-rasG12D mice (25). These experiments collectively indicate a tumorigenic role of IL-17–mediated inflammation in the development of lung cancer.  相似文献   
994.
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   
995.
Adenine is commonly used to establish the animal models for chronic kidney injury and its renal interstitial fibrosis. As an endogenous substance, adenine-induced kidney damage has not yet been fully studied and elucidated, except for inflammatory reaction. Here we analyzed the proteomics of kidney of rats after adenine overloading using LS-MS/MS assay, and observed the role of anemoside B4 (B4). The results showed that adenine could down-regulate 285 proteins and up-regulate 164 proteins in rat kidney tissue compared with the normal group. Down-regulated proteins mainly affected related pathways, such as energy metabolism, while up-regulated proteins affected inflammatory response pathways and metabolic pathways. B4 could significantly reverse the down-regulation of about 40 proteins, which were involved in mitochondria, redox processes, extracellular exosomes, acetylation and other signaling pathways. Simultaneously, B4 could inhibit the up-regulation of five proteins caused by adenine, which were involved in cell cycle, oocyte meiosis, PI3K-Akt and other signaling pathways. Further experimental results of mRNA expression using real-time PCR assay supported the proteomic analysis. Therefore, we proposed that the damage of rat kidney caused by adenine was more complicated, not only with an inflammatory reaction, but also with extensive effects to various metabolic processes in the body. This work provided a valuable clue for comprehensive understanding of adenine-induced renal damage.  相似文献   
996.
We aimed to explore whether diurnal blood pressure (BP) peak characteristics have a significant influence on the association between left ventricular damage with the two BP components (morning BP vs. afternoon peak BP) in untreated hypertensives. This cross‐sectional study included 1084 hypertensives who underwent echocardiography and 24‐h ambulatory BP monitoring. Participants were stratified according to the relationship between morning systolic BP (MSBP; average SBP within 2 h of waking up) and afternoon peak systolic BP (ASBP; average SBP between 16:00 and 18:00). Afternoon and morning hypertension was defined as ≥ 135/85 mm Hg. The morning and afternoon peak BPs occurred at around 7:00 and 17:00, respectively. In general hypertensives, morning BP and afternoon peak BP are significantly different in absolute values (for binary SBP, McNemar''s χ2 = 6.42; p = .014). ASBP was more pronounced than MSBP in 602 patients (55.5%), in whom 24‐h SBP showed higher consistency with ASBP than with MSBP (Kappa value: 0.767 vs 0.646, both p < .01). In subjects with ASBP ≥ MSBP, ASBP was associated with left ventricular hypertrophy independent of MSBP (logistic regression analysis odds ratio: 1.046, p < .01), and left ventricular mass index was more strongly correlated with ASBP than with MSBP (multiple regression coefficient β: 0.453, p < .01), in which the relationships held true independently of 24‐h SBP. The opposite results were obtained in subjects with MSBP > ASBP. Peak BP‐guided monitoring may serve as an effective approach to out‐of‐office hypertension monitoring and control, providing the best consistency with 24‐h average SBP and highest discrimination performance for target organ damage, independently of 24‐h SBP.  相似文献   
997.
High‐normal albuminuria is related to the morbidity and mortality of cardiovascular disease. Arterial stiffness has been regarded as a predictor of cardiovascular disease. However, the relationship between high‐normal albuminuria and arterial stiffness is uncertain in Chinese population. A total of 1343 Chinese participants (aged 58.9 ± 12.1 years, 63.53% male) were included in this study. High‐normal albuminuria was defined as urinary albumin‐to‐creatinine ratio (UACR) above the median within normal albuminuria. Based on the level of UACR, all participants were divided into low‐normal albuminuria group (UACR < 6.36 mg/g, n = 580), high‐normal albuminuria group (6.36 mg/g ≤ UACR < 30 mg/g, n = 581), microalbuminuria (30 mg/g ≤ UACR < 300 mg/g, n = 162), and macroalbuminuria (UACR ≥ 300 mg/g, n = 20). Arterial stiffness was assessed by measuring carotid‐femoral pulse wave velocity (cfPWV). With the increment of UACR, the level of cfPWV was increased gradually (P < .001). Stepwise multiple regression analysis showed that systolic blood pressure, age, serum creatinine, heart rate, logarithmic (LG)‐transformed UACR, and fasting plasma glucose were independently associated with cfPWV in all subjects (P < .001). LG‐UACR was found to be related to cfPWV in high‐normal albuminuria and macroalbuminuria subjects. After further stratification in the high‐normal albuminuria subjects, their relation remained in male, elderly over 65 years old, or normotensives. In summary, UACR is associated with arterial stiffness in subjects with proteinuria excretion in high normal level. High‐normal albuminuria might be an early indicator of arterial stiffness, especially in male, elderly, or normotensives in Chinese population. Furthermore, age and blood pressure are still observed to be the most important risk factor of arterial stiffness.  相似文献   
998.
Recent clinical guidelines recommend lower blood pressure (BP) goals for most patients, and recent trends have favored use of automated unattended BP measurements in the office setting to minimize observer error and white‐coat effects. Patients attending a routinely scheduled CVD clinic visit were prospectively randomized to BP measured using an attended, followed by an unattended method, or vice versa, after a controlled rest period. All study BP measurements were obtained in triplicate using the automated Omron HEM‐907XL BP monitor, and averaged. The outcome was difference in SBP. Routinely measured clinic BP from the same visit was extracted from the medical record, and compared with attended and unattended BP. A total of 102 patients were randomized, and mean age was 63 years, 52% female and 75% Caucasian. Attended and unattended SBP was 125.4 ± 20.4 and 122.6 ± 21.0 mm Hg, mean ± SD, respectively. Routine clinic SBP was 130.6 ± 23.6 mm Hg. Attended SBP was 2.7 mm Hg higher than the unattended measurement (95% CI 1.3‐4.1; P = .0002). Routine clinic SBP was 5.2 mm Hg higher than attended SBP (95% CI 2.4‐8.0; P = .0003) and 8.0 mm Hg higher than unattended SBP (95% CI 5.4‐10.5; P < .0001). Attended measurement of BP is significantly higher than unattended measurement and the difference is physiologically meaningful, even in a CVD cohort with generally well‐controlled hypertension. Furthermore, routine clinic SBP substantially overestimates both attended and unattended automated SBP, with important implications for treatment decisions like dose and/or drug escalation.  相似文献   
999.
探讨快速康复外科(fast track surgery,FTS)在产科手术中的安全性及可行性。收集择期剖宫产孕妇493例,随机分为快速康复组150例,对照组343例(包括止痛药组39例,止痛药+杜密克组29例,镇痛泵组71例,未用止痛药及杜密克镇痛泵组204例),比较术后排气时间、排尿时间、排便时间、并发症、子宫复旧、疼痛评分和住院时间。快速康复组较对照组术后疼痛轻、子宫复旧好、胃肠功能及膀胱功能恢复快、术后住院时间短(P<0.05或P<0.01),术后并发症及Hb下降值无明显差异(P>0.05);在对照组中,止痛药+杜密克组排气排便更快(P<0.05)。FTS方案可减轻患者术后疼痛、加快术后胃肠道功能及子宫复旧,降低术后住院时间、便秘发生率,同时并不增加尿潴留、腹胀发生率以及出血量。  相似文献   
1000.
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