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   562篇
  免费   36篇
  国内免费   1篇
耳鼻咽喉   8篇
儿科学   5篇
妇产科学   11篇
基础医学   112篇
口腔科学   1篇
临床医学   58篇
内科学   124篇
皮肤病学   14篇
神经病学   71篇
特种医学   7篇
外科学   38篇
综合类   2篇
预防医学   43篇
眼科学   2篇
药学   34篇
中国医学   1篇
肿瘤学   68篇
  2023年   3篇
  2022年   6篇
  2021年   21篇
  2020年   14篇
  2019年   13篇
  2018年   13篇
  2017年   15篇
  2016年   19篇
  2015年   20篇
  2014年   25篇
  2013年   31篇
  2012年   37篇
  2011年   33篇
  2010年   18篇
  2009年   27篇
  2008年   37篇
  2007年   41篇
  2006年   27篇
  2005年   29篇
  2004年   27篇
  2003年   36篇
  2002年   38篇
  2001年   2篇
  2000年   3篇
  1999年   1篇
  1998年   11篇
  1997年   7篇
  1996年   6篇
  1995年   3篇
  1994年   3篇
  1993年   2篇
  1992年   3篇
  1991年   3篇
  1990年   2篇
  1989年   1篇
  1988年   2篇
  1987年   1篇
  1986年   1篇
  1985年   2篇
  1984年   5篇
  1983年   2篇
  1982年   2篇
  1981年   2篇
  1966年   2篇
  1965年   1篇
  1964年   1篇
  1959年   1篇
排序方式: 共有599条查询结果,搜索用时 19 毫秒
51.
Defibrotide, a polydesoxyribonucleotide derivative with antithrombotic and fibrinolytic activity, capable of inducing the release of PGI2 from vascular endothelia, was proposed as an alternative to standard heparin coverage during blood dialysis for patients at risk of bleeding. The original procedure featured the preliminary washing of the dialysis circuit with heparin, which was then recirculated and eliminated, and the two drugs, heparin and defibrotide, are known to interact with each other. The purpose of this present study was to explore the ex-vivo heparin activity (assessed as anti-Xa activity) in diverse hemodialysis models using defibrotide (800 mg intravenous, in 4 bolus injections) and various dosages of heparin. Anti-Xa activity is negligible in dialysis conducted with defibrotide alone. When the circuit was prewashed with heparin (5000 and 2500 IU), there was evident anti-Xa activity (0.3-0.5 U/ml) in the first 30-60 minutes of dialysis; continuous heparin infusion (500 U/hour) resulted in high anti-Xa activity levels at the end of dialysis. Thus the best hemodialysis procedure for patients at high risk of bleeding should be one utilizing only defibrotide, or defibrotide plus small amounts of calcium heparin infused at the rate of 500 U/hour for not more than two hours.  相似文献   
52.
Plasma B-type natriuretic peptide (BNP) concentration was evaluated in end-stage renal disease patients to verify if measurements before or after the session could furnish different information. BNP levels in plasma from 52 hemodialysis (HD) patients were measured both before and after the first session of the week. Echocardiographic studies were also performed and patients were followed over a period of 28 months. BNP removal from plasma was influenced by equilibrated Kt/V and patient characteristics. Initial plasma BNP concentration was correlated both with cardiac systolic function (LVEF) and mortality rate, independent of blood sample timing (before or after HD). A relative risk of death of 2.67 was found for plasma BNP levels above 335 pg/mL or 232 pg/mL, before and after HD, respectively. Higher BNP levels were observed in patients with higher burden of comorbidity, as measured by the Charlson Comorbidity Index; however, statistical significance was obtained only for BNP measured before HD. In conclusion, measurement of plasma BNP could give a valuable risk stratification of HD patients while cutting costs, by confining echocardiographic studies only to cases with BNP levels above the established cutoff values.  相似文献   
53.
54.
We report here the course and outcomes of 18-month enzyme replacement therapy in two 43 and 41-year-old brothers with Fabry disease. At 18 months of recombinant alpha-galactosidase A (Fabrazyme) infusions, we observed in the older patient: weight gain, decreased proteinuria (from 4 to 1.5 g/d), stabilization of creatinine clearance, much lower frequency and intensity of angina, and in the younger brother: weight gain, stabilization of creatinine clearance and proteinuria, prolongation of PQ interval and improvement of hearing. However, neurologic manifestations deteriorated over treatment period in both patients. No serious infusion-related side effects were observed.  相似文献   
55.
56.
The use of electric fields to alter the conductivity of correlated electron oxides is a powerful tool to probe their fundamental nature as well as for the possibility of developing novel electronic devices. Vanadium dioxide (VO2) is an archetypical correlated electron system that displays a temperature-controlled insulating to metal phase transition near room temperature. Recently, ionic liquid gating, which allows for very high electric fields, has been shown to induce a metallic state to low temperatures in the insulating phase of epitaxially grown thin films of VO2. Surprisingly, the entire film becomes electrically conducting. Here, we show, from in situ synchrotron X-ray diffraction and absorption experiments, that the whole film undergoes giant, structural changes on gating in which the lattice expands by up to ∼3% near room temperature, in contrast to the 10 times smaller (∼0.3%) contraction when the system is thermally metallized. Remarkably, these structural changes are fully reversible on reverse gating. Moreover, we find these structural changes and the concomitant metallization are highly dependent on the VO2 crystal facet, which we relate to the ease of electric-field–induced motion of oxygen ions along chains of edge-sharing VO6 octahedra that exist along the (rutile) c axis.The use of electric fields to influence the transport properties of various materials by electrostatic injection of charge at an interface is the foundation of much of modern day electronics (1). Using a three-terminal field-effect transistor geometry, the magnitude of the electric fields provided by conventional gate dielectrics is limited by their dielectric properties. Much higher electric fields are possible by replacing the conventional gate material with an ionic liquid. Consequently, much higher electrostatically induced charge densities are possible, leading to the control or creation of novel metallic (23) and superconducting phases (47). Materials that are insulating by virtue of strong electron–electron correlations, namely Mott–Hubbard and charge-transfer insulators (8), are anticipated to be particularly sensitive to the injection of small numbers of carriers that could result in their metallization (911). Often these materials exhibit a thermally driven insulator to metal transition: one of these, VO2, exhibits such a transition near room temperature (12, 13) and, for this reason, has been extensively studied (1416). In VO2 the metal to insulator transition (MIT) is accompanied by a structural phase transition (SPT) in which the monoclinic insulating phase transforms to a rutile metallic phase (17). Recently, both Nakano et al. (18) and Jeong et al. (19) showed that ionic liquid (IL) gating of thin films of VO2 results in the suppression of the MIT to temperatures below ∼10 K and, moreover, that the entire film becomes metallic even though gating takes place at the top surface of the film in contact with the IL. However, whereas Nakano et al. (18) claimed the metallization phenomenon was a direct result of electrostatic carrier injection, Jeong et al. (19) presented clear evidence that the metallic state was rather induced by the electric-field–induced migration of oxygen from the film into the IL. An important question is whether the IL gating results in a structural phase transition, as supposed by Nakano et al. (18), or whether the initially insulating film remains in the monoclinic phase and the metallization results rather from the formation of oxygen vacancies (19). Here, we show, using in situ X-ray diffraction and absorption, that IL gating induces massive, reversible structural changes in which the VO2 (001) film expands and contracts along its thickness by up to 3%, but that the film remains in the monoclinic phase. Furthermore, we identify a remarkable dependence of the IL gating phenomenon on the crystal facet of the VO2 films. Whereas the (001) and (101) facets exhibit similar IL gate-induced metallization, almost no effect is observed for films grown with (100) and (110) facets. Because there are open channels in the VO2 crystal structure along the rutile c axis, we associate the IL gating phenomenon with the ease of migration of oxygen along these channels under the influence of electric field. Previously, clear evidence for the formation and refilling of oxygen vacancies under ionic liquid gating of VO2 (001) has been reported (19).The facet-dependent IL-induced metallization and associated structural changes of VO2 were studied using two types of devices shown in Fig. 1 A and B, respectively. We label these devices as type T and X, respectively. In Fig. 1A a typical device type T with a channel area of 200 × 20 µm2 defined by optical lithography is shown (see Methods for details). The channel conductance is measured using the source (S) and drain (D) contacts that are shown in the figure. A drop of the IL (∼100 nl) fully covers the channel and a significant part of the lateral gate electrode (19). Such devices T are suitable for detailed transport studies as a function of temperature and environment. Fig. 1B shows a cell (20) that was specially designed for in situ synchrotron-based X-ray measurements. Device X, used in this cell, is much larger than that of Fig. 1A and indeed is almost as large as the substrate itself (1 × 1 cm2) with S and D gold contacts, defined by shadow masks, along opposite edges of the substrate. The device and the gate electrode, which is formed from a coiled Au wire that surrounds the device, are immersed in ∼2 mL of IL which is introduced through Teflon tubes and which is contained by a 7.5-µm-thick Kapton sheet, sealed with Viton O-rings, that allows for transmission of the incident and diffracted X-ray beams and fluorescent X-rays. The cell is attached to a four-circle X-ray goniometer for the X-ray diffraction studies. Incorporated within the cell is a heater and a Peltier cooler that allows for operation at temperatures ranging from ∼250 K to ∼400 K. Pulsed laser deposition was used to deposit 10-nm-thick VO2 films with four different crystal facets, (001), (101), (100), and (110) on single-crystalline substrates of TiO2 with the same respective crystal orientations, and 20-nm-thick VO2 (001) films on Al2O3(101¯0)(see Methods for details).Open in a separate windowFig. 1.Two different types of ionic liquid devices and facet dependence of gating effect. (A) Optical image of a device with a droplet of HMIM-TFSI with channel size of 200 × 20 µm2. (B) Optical image of an ionic liquid device for in situ X-ray measurement. The entire film surface (10 × 10-mm2 area) is covered with ionic liquid surrounded by a Au wire used as a gate electrode. Source–drain (Top) and gate (Bottom) current versus VG for small size device (C) and large size device (D) fabricated from VO2 films grown on TiO2 substrates of different orientations. Gate voltages were swept at a rate of 3 mV/s (0.3 mV/s for D) and source–drain voltage (VSD) of 100 mV (300 mV for D) is applied.Fig. 1 C and D compares the gate voltage (VG) dependence of the source–drain current ISD and the gate current IG at 270 K for devices X and T. Initially the devices are in the insulating phase but above a certain threshold gate voltage ISD increases substantially. When VG is decreased to zero the devices remain conducting and revert back to their original state only when reverse gated by applying a negative voltage. IG remains below 5 nA for all VG for device T. For device X, IG is much larger but only because of the much larger gated area: the leakage current per unit area of the gated VO2 is similar for both devices (see SI Appendix for a detailed comparison). In the fully gated state, device T is metallic to low temperatures as shown earlier (19) but device X was only measured to 250 K, due to limitations of the X-ray cell, where it remained metallic. An important result is the dramatic dependence of the IL gating on the VO2 crystal facet, as is clearly shown in Fig. 1 C and D.X-ray diffraction θ–2θ curves from a VO2 (001) device X in the insulating phase and the thermally induced metallic phase are shown in Fig. 2A. The unit cell (all peaks are indexed throughout the paper with respect to the rutile unit cell, for simplicity) contracts along the c axis as evidenced by the shift of the VO2 (002) peak to higher as VO2 transforms from the monoclinic to a rutile phase. Fig. 2B shows a sequence of X-ray θ–2θ curves for the device in different gated states as VG was systematically ramped in steps from 0 to +2.2 V to −2.2 V and back to zero. The X-ray data were collected after VG was fixed at each step for 30 min. Fig. 2C shows the corresponding values of ISD for these data. The gate voltage-induced increase in ISD is ∼3 orders of magnitude. The X-ray data show very large shifts in the VO2 (002) peak position which, however, are opposite to that seen for the temperature-driven MIT shown in Fig. 2A. The VO2 (002) peak rather shifts to smaller values. This corresponds, as shown in Fig. 2D, to an expansion of the c-axis parameter in the fully gated metallic state by a factor which is 10 times larger than the contraction in the c-lattice parameter that is observed for the temperature-driven SPT. We note that the threshold voltages at which the lattice changes are observed appear to be smaller than that at which ISD changes. We find that no structural changes are observed when the same experiment is carried out on a 10-nm-thick VO2 film grown with a (110) facet on TiO2(110). X-ray diffraction θ–2θ curves taken during gating at gate voltages of up to 2.8 V show no shift in the VO2 (220) peak position nor any other changes in the X-ray diffraction curves. As shown in Fig. 1D there are similarly no changes in ISD during gating up to VG = 3 V.Open in a separate windowFig. 2.Structural changes of VO2/TiO2(001) by electrolyte gating as a function of gate voltages. (A) XRD patterns of insulating (red) phase at 270 K and metallic phase (black) at 300 K for 10-nm VO2/TiO2(001) with 12-keV photon energy. (B) XRD pattern for in situ X-ray measurements and (C) source–drain current (ISD) versus gate voltage (VG). Both XRD and ISD were measured ∼30 min after VG was applied. (D) c-axis lattice parameter extracted from B versus gate voltage. The error bars are from the nonlinear least-squares fitting algorithm and in many cases are smaller than the symbols.The structural changes that we find on gating VO2 (001) are similar to those found by growing VO2 (001) films of comparable thickness at lower oxygen pressures during pulsed laser deposition. Fig. 3A shows X-ray diffraction θ–2θ curves for a series of five samples, each ∼10 nm thick, prepared at oxygen pressures of 9, 7, 5, 3, and 1 mTorr. As the oxygen pressure is reduced the c-lattice parameter systematically increases, as shown in Fig. 3B. The MIT transition is systematically broadened and suppressed to low temperatures as the oxygen pressure is reduced (19). The c-lattice parameter expansion caused by the introduction of oxygen vacancies by modifying the film growth conditions are similar to those that IL gating induces, and the resulting metallization of the VO2 films is comparable. We presume that the thickness oscillations in the θ–2θ curves from VO2/TiO2(001) in Fig. 2B that disappear on gating result from the loss of coherence in the film structure due to gating and the subsequent formation of oxygen vacancies.Open in a separate windowFig. 3.Crystal structure of oxygen-deficient and gated VO2/TiO2(001). Dependence of (A) XRD θ–2θ curves, and (B) c-lattice parameter and TMIT on oxygen pressure during growth. (C) Reciprocal space maps of VO2(202) peak versus oxygen pressure during film growth and comparison with those for the pristine and gated states of a device formed from a film grown at 9 mtorr. (D) Structure of the monoclinic phase of VO2 looking along the <001> and <110> axes.The crystal facet of the VO2 film is determined by epitaxial growth onto the respective facet of the TiO2 substrate. This also results in clamping of the 10-nm-thick VO2 films to the corresponding TiO2 lattices by coherent strain such that their in-plane unit cell parameters are very similar. This is illustrated in Fig. 3C, which shows reciprocal lattice maps centered near TiO2 (202) in the k = 0 plane for the five films prepared in different oxygen ambients in Fig. 3A. The maps show along the [20l] direction a very sharp, intense TiO2 (202) peak together with a weaker and broader VO2 (202) peak and associated Kiessig fringes (21). The VO2 (202) peak systematically shifts to lower l as the oxygen pressure is reduced. Along the [h02] direction, by contrast, the VO2 and TiO2 peaks have similar narrow widths that indicate in-plane clamping of the VO2 lattice to that of the TiO2 substrate. Thus, during IL gating it is anticipated that only the out-of-plane VO2 lattice parameter can be significantly changed. This is confirmed in the reciprocal space map for a sample that was gated at VG = 3 V for 10 h and the IL removed before the measurement using a laboratory X-ray source.The clamping of the VO2 lattice to that of the TiO2 lattice could offer an explanation for the lack of any significant IL gating response for facets of VO2 for which the c axis lies in plane. It could be that to remove significant amounts of oxygen the lattice needs to expand along the c direction. It is along the rutile c direction that the structure comprises one-dimensional chains of edge-sharing VO6 octahedra that allow for the expansion and contraction of the VO2 lattice during IL gating (Fig. 3D).To inspect the local environment of V we performed in situ X-ray absorption spectroscopy (XAS) at the V K edge using VO2 (001) films grown on Al2O3(101¯0)rather than TiO2 to avoid the otherwise significant fluorescence from Ti in the substrate. The sample was gated and the XAS data were measured at ambient temperature well below the MIT of the pristine film. The X-ray absorption near-edge spectra (XANES) are shown in Fig. 4A for a pristine sample and the same sample after gating (VG = 3 V, 1 h) to a conducting state. The XANES data remain largely unchanged with two exceptions. Firstly, there is a small shift in the position of the inflection point of the V 1s→3d preedge transition and a small decrease in the intensity of the preedge peak that suggest a reduction in the valence state of the V ions [by ∼0.2 electrons per V (22)]. The decrease in the intensity of this preedge feature is also consistent with a decrease in the degree of distortion of the VO6 octahedra (22). Secondly, there is a small shift to lower photon energies in the position of the main V K edge and the white line (V1s → V4p transition) at this edge which is also consistent with a reduced V valence on gating (22). A change in V valence was previously observed in X-ray photoelectron spectroscopy measurements performed on electrolyte-gated VO2 that suggested the formation of oxygen vacancies on gating (19).Open in a separate windowFig. 4.XAS of VO2/Al2O3(101¯0). (A) Vanadium K-edge XANES for a pristine and gated device X. (B and C) χ(R) curves for the data and corresponding fits. The individual contributions to these fits from respective V–O and V–V shells are shown (inverted) in the bottom halves of B and C. The experimental data for the pristine and gated states are shown in blue and red, respectively; the fits to these data are shown in green; the differences between the experimental data and fits are shown in magenta; the contributions from V–O shells are shown in purple; the contributions from V–V shells along the dimer axis are brown and perpendicular to the dimer axis are dark yellow. The brown horizontal lines in B and C are aids to the eye, showing the degree of dimerization, namely, the difference between the intra- and inter-V–V dimer distances. The solid and dashed curves are the moduli and real parts of the Fourier transform of the EXAFS data and the fits, respectively. (D) Table of fitted V–O and V–V bond distances.Much larger changes are observed in the extended X-ray absorption fine structure (EXAFS) (23), χ(k), that is most readily seen in its Fourier transform (FT), namely, χ(R) = FT(k3 χ(k)) (23), that is presented in Fig. 4 B and C. Detailed information on the types of locally ordered neighbor shells and their metrical parameters was obtained by nonlinear least-squares curve fits using calculated amplitudes and phases (24). The k3-weighted data were fit for k varying from 2.6 to 10.3 Å–1 so that shells are distinguishable only if separated by more than ∼0.15 Å. The spectra were well fit (Fig. 4 B and C) with a limited number of shells relative to the crystal structure (SI Appendix). The distances found for the pristine samples are consistent with the monoclinic phase of VO2 below its MIT in which the shorter V–V distances that correspond to those in the (rutile) c direction that is normal to the film plane are split because of the dimerization of the V–V atoms along this direction. In the pristine sample these V–V distances of 2.61 and 3.03 Å differ by 0.42 Å, whereas in the gated sample these distances become 2.94 and 3.16 Å and differ by only 0.22 Å. Note that the almost complete loss of the peak in the χ(R) spectra near R−ϕ = 2.0 Å on gating is not because of a radical change in the V–V chain ordering and departure from the V–V dimerized monoclinic structure, but is because the decreased separation between the short and long dimer V–V pairs causes their individual EXAFS waves to destructively interfere, reducing their combined amplitude in the FT. Thus, a crucial result is that the V–V dimerization, although reduced, is nevertheless retained in the gated state. Moreover, the average V–V distance in the c direction normal to the film planes increases by a much larger amount than is found by diffraction that could indicate rotation of the VO6 octahedra. On the other hand, the V–V distance within the ab plane (∼3.5 Å) changes little on gating (Fig. 4D), suggesting that the structure in the plane is largely unaffected, consistent with the X-ray diffraction data in Fig. 3C.Our in situ X-ray diffraction (XRD) and XAS measurements clearly indicate a giant expansion of the VO2 unit cell that is clearly inconsistent with the formation of the rutile phase that the thermally induced metallic phase exhibits. Moreover, we find these reversible structural changes only in films in which channels formed from chains of edge-sharing VO6 octahedra do not lie in the plane of the films, strongly suggesting that these channels are the paths along which the gate-induced oxygen migration takes place. These gate-induced changes in structure and conductivity are likely to be common to many ionic liquid gated systems, opening the way to a potential future of “liquid electronics.”  相似文献   
57.
AXL is a tyrosine kinase receptor activated by GAS6 and regulates cancer cell proliferation migration and angiogenesis. We studied AXL as new therapeutic target in colorectal cancer (CRC). Expression and activation of AXL and GAS6 were evaluated in a panel of human CRC cell lines. AXL gene silencing or pharmacologic inhibition with foretinib suppressed proliferation, migration and survival in CRC cells. In an orthotopic colon model of human HCT116 CRC cells overexpressing AXL, foretinib treatment caused significant inhibition of tumour growth and peritoneal metastatic spreading. AXL and GAS6 overexpression by immunohistochemistry (IHC) were found in 76,7% and 73.5%, respectively, of 223 human CRC specimens, correlating with less differentiated histological grading. GAS6 overexpression was associated with nodes involvement and tumour stage. AXL gene was found amplified by Fluorescence in situ hybridization (FISH) in 8/146 cases (5,4%) of CRC samples.Taken together, AXL inhibition could represent a novel therapeutic approach in CRC.  相似文献   
58.
59.
60.
Typhoid fever is a public health problem, especially among young children in developing countries. To address this need, a glycoconjugate vaccine Vi-CRM(197), composed of the polysaccharide antigen Vi covalently conjugated to the non-toxic mutant of diphtheria toxin CRM(197), is under development. Here, we assessed the antibody and cellular responses, both local and systemic, following subcutaneous injection of Vi-CRM(197). The glycoconjugate elicited Vi-specific serum IgG titers significantly higher than unconjugated Vi, with prevalence of IgG1 that persisted for at least 60 days after immunization. Vi-specific IgG, but not IgA, were present in intestinal washes. Lymphocytes proliferation after restimulation with Vi-CRM(197) was observed in spleen and mesenteric lymph nodes. These data confirm the immunogenicity of Vi-CRM(197) and demonstrate that the vaccine-specific antibody and cellular immune responses are present also in the intestinal tract, thus strengthening the suitability of Vi-CRM(197) as a promising candidate vaccine against Salmonella Typhi.  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