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   339篇
  免费   27篇
  国内免费   1篇
耳鼻咽喉   1篇
儿科学   1篇
妇产科学   6篇
基础医学   47篇
口腔科学   3篇
临床医学   23篇
内科学   127篇
皮肤病学   2篇
神经病学   32篇
特种医学   21篇
外科学   49篇
综合类   3篇
预防医学   12篇
眼科学   3篇
药学   16篇
肿瘤学   21篇
  2024年   1篇
  2023年   14篇
  2022年   18篇
  2021年   15篇
  2020年   11篇
  2019年   13篇
  2018年   21篇
  2017年   10篇
  2016年   13篇
  2015年   12篇
  2014年   16篇
  2013年   21篇
  2012年   32篇
  2011年   38篇
  2010年   13篇
  2009年   7篇
  2008年   12篇
  2007年   13篇
  2006年   14篇
  2005年   15篇
  2004年   15篇
  2003年   13篇
  2002年   8篇
  2001年   4篇
  2000年   3篇
  1999年   2篇
  1998年   1篇
  1997年   7篇
  1993年   1篇
  1991年   1篇
  1987年   1篇
  1969年   1篇
  1967年   1篇
排序方式: 共有367条查询结果,搜索用时 31 毫秒
41.
42.
43.

Objectives

The aim of this study was to determine whether high-risk patients with left main coronary artery disease (LMCAD) and prior cerebrovascular disease (CEVD) preferentially benefit from revascularization by percutaneous coronary intervention (PCI) compared with coronary artery bypass grafting (CABG).

Background

Patients with known CEVD requiring revascularization are often referred to PCI rather than CABG. There is a paucity of data regarding the impact of CEVD in patients with LMCAD undergoing revascularization.

Methods

In the EXCEL (Evaluation of XIENCE Versus Coronary Artery Bypass Surgery for Effectiveness of Left Main Revascularization) trial, patients with LMCAD and low or intermediate SYNTAX (Synergy Between PCI with Taxus and Cardiac Surgery) scores were randomized to PCI with everolimus-eluting stents versus CABG. The effects of prior CEVD, defined as prior stroke, transient ischemic attack, or carotid artery disease, on 30-day and 3-year event rates were assessed.

Results

Prior CEVD was present in 233 of 1,898 patients (12.3%). These patients were older and had higher rates of comorbidities, including hypertension, diabetes, peripheral vascular disease, anemia, chronic kidney disease, and prior PCI, compared with those without prior CEVD. Patients with prior CEVD had higher rates of stroke at 30 days (2.2% vs. 0.8%; p = 0.05) and 3 years (6.4% vs. 2.2%; p = 0.0003) and higher 3-year rates of the primary endpoint of all-cause death, stroke, or myocardial infarction (25.0% vs. 13.6%; p < 0.0001). The relative effects of PCI versus CABG on the 30-day and 3-year rates of stroke (pinteraction = 0.65 and 0.16, respectively) and the 3-year rates of the primary composite endpoint (pinteraction = 0.14) were consistent in patients with and those without prior CEVD.

Conclusions

Patients with LMCAD and prior CEVD compared with those without CEVD have higher rates of stroke and reduced event-free survival after revascularization. Data from the EXCEL trial do not a priori support a preferential role of PCI over CABG in patients with known CEVD.  相似文献   
44.

Aims

Troponin levels are commonly elevated among patients hospitalized for heart failure (HF), but the prevalence and prognostic significance of early post‐discharge troponin elevation are unclear. This study sought to describe the frequency and prognostic value of pre‐discharge and post‐discharge troponin elevation, including persistent troponin elevation from the inpatient to outpatient settings.

Methods and results

The ASTRONAUT trial (NCT00894387; http://www.clinicaltrials.gov ) enrolled hospitalized HF patients with ejection fraction ≤40% and measured troponin I prior to discharge (i.e. study baseline) and at 1‐month follow‐up in a core laboratory (elevation defined as >0.04 ng/mL). This analysis included 1469 (91.0%) patients with pre‐discharge troponin data. Overall, 41.5% and 29.9% of patients had elevated pre‐discharge [median: 0.09 ng/mL; interquartile range (IQR): 0.06–0.19 ng/mL] and 1‐month (median: 0.09 ng/mL; IQR: 0.06–0.15 ng/mL) troponin levels, respectively. Among patients with pre‐discharge troponin elevation, 60.4% had persistent elevation at 1 month. After adjustment, pre‐discharge troponin elevation was not associated with 12‐month clinical outcomes. In contrast, 1‐month troponin elevation was independently predictive of increased all‐cause mortality [hazard ratio (HR) 1.59, 95% confidence interval (CI) 1.18–2.13] and cardiovascular mortality or HF hospitalization (HR 1.28, 95% CI 1.03–1.58) at 12 months. Associations between 1‐month troponin elevation and outcomes were similar among patients with newly elevated (i.e. normal pre‐discharge) and persistently elevated levels (interaction P ≥ 0.16). The prognostic value of 1‐month troponin elevation for 12‐month mortality was driven by a pronounced association among patients with coronary artery disease (interaction P = 0.009).

Conclusions

In this hospitalized HF population, troponin I elevation was common during index hospitalization and at 1‐month follow‐up. Elevated troponin I level at 1 month, but not pre‐discharge, was independently predictive of increased clinical events at 12 months. Early post‐discharge troponin I measurement may offer a practical means of risk stratification and should be investigated as a therapeutic target.
  相似文献   
45.
Low doses of the humanized anti‐CD20 monoclonal antibody, veltuzumab, were evaluated in 41 patients with immune thrombocytopenia (ITP), including 9 with ITP ≤1 year duration previously treated with steroids and/or immunoglobulins, and 32 with ITP >1 year and additional prior therapies. They received two doses of 80–320 mg veltuzumab 2 weeks apart, initially by intravenous (IV) infusion (N = 7), or later by subcutaneous (SC) injections (N = 34), with only one Grade 3 infusion reaction and no other safety issues. Thirty‐eight response‐assessable patients had 21 (55%) objective responses (platelet count ≥30 × 109/l and ≥2 × baseline), including 11 (29%) complete responses (CRs) (platelet count ≥100 × 109/l). Responses (including CRs) occurred with both IV and SC administration, at all veltuzumab dose levels, and regardless of ITP duration. Responders with ITP ≤1 year had a longer median time to relapse (14·4 months) than those with ITP >1 year (5·8 months). Three patients have maintained a response for up to 4·3 years. SC injections resulted in delayed and lower peak serum levels of veltuzumab, but B‐cell depletion occurred after first administration even at the lowest doses. Eight patients, including 6 responders, developed anti‐veltuzumab antibodies following treatment (human anti‐veltuzumab antibody, 19·5%). Low‐dose SC veltuzumab appears convenient, well‐tolerated, and with promising clinical activity in relapsed ITP.( Clinicaltrials.gov identifier: NCT00547066.)  相似文献   
46.
The attractive/repulsive relationship between superconductivity and magnetic ordering has fascinated the condensed matter physics community for a century. In the early days, magnetic impurities doped into a superconductor were found to quickly suppress superconductivity. Later, a variety of systems, such as cuprates, heavy fermions, and Fe pnictides, showed superconductivity in a narrow region near the border to antiferromagnetism (AFM) as a function of pressure or doping. However, the coexistence of superconductivity and ferromagnetic (FM) or AFM ordering is found in a few compounds [RRh4B4 (R = Nd, Sm, Tm, Er), R′Mo6X8 (R′ = Tb, Dy, Er, Ho, and X = S, Se), UMGe (M = Ge, Rh, Co), CeCoIn5, EuFe2(As1−xPx)2, etc.], providing evidence for their compatibility. Here, we present a third situation, where superconductivity coexists with FM and near the border of AFM in Fe1−xPdxTe. The doping of Pd for Fe gradually suppresses the first-order AFM ordering at temperature TN/S, and turns into short-range AFM correlation with a characteristic peak in magnetic susceptibility at TN. Superconductivity sets in when TN reaches zero. However, there is a gigantic ferromagnetic dome imposed in the superconducting-AFM (short-range) cross-over regime. Such a system is ideal for studying the interplay between superconductivity and two types of magnetic (FM and AFM) interactions.Since the first discovery of superconductivity (SC) a century ago, the effects of magnetic impurities and the possibility of magnetic ordering in superconductors has been a central topic of condensed matter physics. Due to strong spin scattering (1, 2), it has generally been believed that the conduction electrons cannot be both magnetically ordered and superconducting. Even though it is thought that Cooper pairs in cuprates, heavy fermions, and Fe-based compounds are mediated by spin fluctuations (35), SC generally occurs after suppressing the magnetic ordering either through chemical doping or the application of hydrostatic pressure (610). However, there is growing evidence for the coexistence of superconductivity with either ferromagnetic (FM) (1120) or antiferromagnetic (AFM) ordering (2124). With the decrease of temperature (T), some of these systems show magnetic ordering before the superconducting transition (Tc) (1417, 20), some are ordered in a reversed sequence (1113, 18, 19, 22), some have the two orderings occur concomitantly (22, 25), and some show reentrant superconductivity (partially) overlapping with a magnetically ordered phase (1113, 26). Despite extensive investigations of interaction between SC and magnetic moments, there is so far no unified theory for the coexistence of SC and magnetism. With the lack of theoretic guidance, the existing experimental findings lead to two schools of thought: one is that both orders result from the same conduction electrons as evidenced by their synchronized magnetic and superconducting orders (22), and the other is that there are two separate sets of electrons responsible for magnetic ordering and superconductivity, respectively (19, 21, 25). What remains incomprehensible is the case where superconductivity and magnetic ordering coexist but are in competition with each other, as seen in Fe-based systems (23, 24, 27).The discovery of superconductivity in Fe-based compounds has sparked enormous interest in the scientific community. Although Fe is the most well known ferromagnet, all parent compounds of Fe-based superconductors exhibit AFM ordering. Though superconductivity is induced after suppressing the AFM ordering, it can coexist with either remaining AFM ordering (23, 24, 27) or new FM ordering (18, 19), and this provides an ideal platform for studying the interplay between superconductivity and magnetism. Among the Fe-based superconductors, the chalcogenide FeTe1−xSex is unique in several aspects: (i) it is the only compound composed of slabs of Fe(Te/Se)4 stacked together without an interlayer spacer; (ii) it becomes superconducting via isovalent doping of Se for Te, with the highest Tc occurring at the 50% doping level; and (iii) the parent compound Fe1+yTe shows nonmetallic electrical conduction and forms (π, π)-type AFM ordering with a large magnetic moment (28, 29), in contrast to the (π, 0)-type ordering with a small magnetic moment seen in Fe pnictides. The unusual AFM order in Fe1+yTe cannot be explained by a simple Fermi-surface nesting picture, thus leading to arguments for a correlated local-moment scenario (27). Because of these differences, the application of hydrostatic pressure or partial doping on the Fe site, such as with Co or Ni, does not generate the generic phase diagram as seen for other Fe-based superconductors (3033).To gain insight into the relationship between magnetism and superconductivity, we choose to substitute Pd for Fe in FeTe. This decision is motivated by the following: (i) Pd is known to exhibit FM instability (34, 35) even though it is paramagnetic in bulk, and (ii) PdTe is a superconductor with Tc ∼4.5 K (3642). The partial replacement of Fe by Pd will allow us to understand the roles of Fe, Pd, and Te in both magnetism and superconductivity. Based on electrical transport, and magnetic and thermodynamic property measurements, we show that the ground state of Fe1−xPdxTe varies from AFM to FM ordering, to a superconducting state; this is one of rare cases where superconductivity flirts with both AFM and FM.Single crystals of Fe1−xPdxTe were grown via high-temperature melting, with the procedure described in SI Text. The crystal structure and the phase purity were measured by both powder and single-crystal X-ray diffraction. Fig. 1 shows powder X-ray diffraction patterns (Fig. 1A) and lattice parameters obtained from single-crystal X-ray refinement (Fig. 1B) for different doping levels; it confirms that the undoped FeTe forms a tetragonal structure, belonging to the P4/nmm space group. At room temperature, the lattice parameters are a = 3.8202 Å and c = 6.2686 Å. Similar to previous observations (43), our single-crystal refinement result indicates that there are ∼9% extra Fe atoms (T2 site of Fig. 1B) incorporated at interstitial sites of the Te layers. As soon as Pd is introduced into the system, the interstitial sites of Fe1−xPdxTe are occupied by Pd atoms (i.e., T2 = Pd when x > 0); though it remains in the same tetragonal space group, its lattice parameters a and c both increase with increasing Pd doping concentration for xxs ∼0.6 (Fig. 1B). This finding indicates that Pd doping creates negative pressure, which is surprising because Pd2+ (∼0.80 Å) has a nearly identical ionic radius as that of Fe2+ (∼0.77 Å). The crystal structure of Fe1−xPdxTe remains tetragonal up to xs (∼0.6). Above xs, Fe1−xPdxTe crystallizes in a hexagonal structure (space group P63/mmc), and its lattice parameters also increase with increasing x (Fig. 1B); this results in an increase of the unit cell volume with increasing x in the entire doping range (Fig. 1B). In the hexagonal structure, the system no longer incorporates any interstitial sites (Tables S1S4). Though the hexagonal structure at x > xs can be regarded as a deformed FeTe structure, the local environment of Fe/Pd is transformed from a tetrahedron (x < xs) to an octahedron (x > xs) (44).Open in a separate windowFig. 1.Fe1−xPdxTe: (A) Powder X-ray diffraction patterns and (B) its unit cell parameters at room temperature obtained from single-crystal X-ray refinement.To correlate the structural information with physical properties, we show, in Fig. 2, the doping dependence of structural, electrical, and magnetic properties of Fe1−xPdxTe, with specifics presented later. The undoped (x = 0) Fe1.09Te undergoes both an AFM ordering and a structural transition from paramagnetic (PM) tetragonal (high temperatures) to an AFM-ordered monoclinic (low temperatures) phase at TN/S ∼70 K, with the first-order characteristic. Upon Pd doping, TN/S is suppressed (solid circles), reaching zero at x = xN/S ∼0.15 (dash line). In this doping region, the electrical resistivity (ρ) changes from nonmetallic (NM) character (dρ/dT < 0) above TN/S to metallic (M) behavior (dρ/dT > 0) below TN/S. When the characteristic feature for the first-order transition is absent at xN/S, the magnetic susceptibility of Fe1−xPdxTe exhibits a peak at TN. The value of TN decreases with increasing x (Fig. 2, open circles), which eventually reaches zero at xc ∼0.88. The magnetic susceptibility peak is a signature for short-range (SR) AFM correlation developed below TN, with no obvious fingerprint in the electrical resistivity. Above xc, superconductivity appears (Fig. 2, solid diamonds). The transition temperature Tc increases with increasing x. Remarkably, the system shows ferromagnetism (Fig. 2, solid squares) in the doping range of 0.78 ≤ x ≤ 0.98 with the maximum TFM (∼190 K) at xc, where AFM and SC vanish. In addition to the variable magnetic phases, there is doping-induced structural phase transition, being tetragonal at x < xs ∼0.6 (Fig. 2, dark blue) and hexagonal at x > xs (light blue). Such a structural transition is responsible for a change of electronic structure, manifested by NM behavior at x < xs and a metallic character at x > xs.Open in a separate windowFig. 2.Phase diagram of Fe1−xPdxTe. AFM, antiferromagnetic; AFM-M, antiferromagnetically ordered metallic state; FM, ferromagnetically ordered state; M, metallic; NM, nonmetal; PM, paramagnetic; PM-M, paramagnetic metal; SC, superconductivity; SR, short-range; SR-AFM, short-range AFM correlation.We now present the detailed experimental results in support of the phase diagram. Fig. 3A shows the temperature dependence of ρ of Fe1−xPdxTe for 0 ≤ x ≤ 0.5 (the experimental technique is described in SI Text). Similar to previous observations (45), the electrical resistivity of undoped (x = 0) Fe1.09Te increases with decreasing temperature at high temperatures, but sharply decreases below TN/S ∼70 K; this is caused by the first-order coupled structural and AFM phase transition (45, 46). Below TN/S, the structure of Fe1.09Te becomes monoclinic. Upon Pd doping, TN/S is suppressed, and the first-order transition is no longer observed at x = 0.2. The smooth and monotonic temperature dependence of ρ indicates the absence of both structural and AFM transitions at 0.1 < x < 0.2. Though it sustains a nonmetallic temperature dependence (dρ/dT < 0) down to 2 K, the magnitude of the electrical resistivity deceases with increasing x in the entire doping range above TN/S (Fig. 3 A and C). For 0.5 < x ≤ 1.0, the electrical resistivity of Fe1−xPdxTe continues to decrease with increasing x, as shown in Fig. 3B. More interesting is that ρ shows metallic behavior (dρ/dT > 0) in the entire temperature range measured for 0.5 < x ≤ 1.0. We believe that the cross-over from nonmetallic to metallic character is associated with the structural change at xs. Furthermore, there is a sharp drop of resistivity at low temperatures for 0.88 < x ≤ 1.0, as shown in Fig. 3D. This drop is due to the emergence of superconductivity because the magnetic susceptibility shows diamagnetism as well in the temperature range (Fig. 3H). Note that the superconducting transition temperature Tc increases with increasing x.Open in a separate windowFig. 3.(A and B) Temperature dependence of the electrical resistivity (ρ) of Pd1−xFexTe at the indicated compositions. (C and D) Temperature dependence of ρ near the first-order AFM/structural transition in the low-doping region, and the superconducting transition in the high-doping region plotted as ρ/ρ(5 K) vs. T, respectively. (E–G) Temperature dependence of the magnetic susceptibility (χ) of Fe1−xPdxTe measured at a magnetic field of 1 T. G Inset shows χ vs. x = 0.99 and 1.0; H is the χ(T) measured at 20 oersted (Oe) for 0.97 ≤ x ≤ 1.0.With Pd doping, the magnetic properties of Fe1−xPdxTe also change. Fig. 3 E–H shows the temperature dependence of the magnetic susceptibility, χ, at indicated values x. For 0 ≤ x < 0.2, χ initially increases with decreasing temperature, and then drops steeply at TS/N (Fig. 3E); this confirms the first-order nature of the structural/AFM transition at 0 ≤ x < 0.2. Such a feature is absent for xxN/S ∼0.15. Instead, there is a characteristic peak in magnetic susceptibility at TN (Fig. 3F). Because there is no anomaly in electrical resistivity (Fig. 3 A–C) or specific heat, the susceptibility peak cannot be due to a true phase transition, but rather marks the onset of SR AFM correlation as seen in other materials (47). With increasing x, TN moves to lower temperatures, and the magnitude of χ decreases as well (Fig. 3F). Though the characteristic peak is expected to completely vanish around x = xc ∼0.88, the magnetic susceptibility reveals a dramatic increase below another characteristic temperature TFM at x ≥ 0.78, as shown in Fig. 3G, which indicates the onset of ferromagnetic ordering at TFM. With increasing x, TFM initially increases then decreases after reaching the maximum (∼190 K) at x = xc ∼0.88. Interestingly, the low-temperature susceptibility peak is still present for samples with x = 0.78, 0.80, 0.83, and 0.88, which suggests the coexistence of FM ordering and AFM interactions in this doping region. Above xc, χ(T) continues to show FM behavior at high temperatures, but becomes negative at low temperatures (below Tc) under low magnetic fields, as depicted in Fig. 3H. The negative χ below Tc indicates the emergence of superconductivity, because their electrical resistivities also drop sharply (Fig. 3D). Whereas TFM decreases, Tc increases with increasing x (above xc). As shown Fig. 3G Inset, χ(T) exhibits paramagnetic behavior above Tc for x = 0.99 and 1.0, indicating the complete suppression of FM. However, Tc continues to increase with x (Fig. 3H).Clearly, there is a coexistence of AFM and FM ordering at 0.78 ≤ x ≤ 0.88, and of FM ordering and SC at 0.88 ≤ x ≤ 0.98 at low temperatures for Fe1−xPdxTe, which is further supported by the following results. First, the high-temperature magnetic susceptibility data allows us to extract both Curie–Weiss temperature CW) and effective magnetic moment (μeff) by fitting our experimental data to the Curie–Weiss law , where NA is the Avogadro’s constant and kB is Boltzmann’s constant. The x dependences of θCW and μeff are shown in Fig. 4 A and B, respectively. Note that θCW < 0 for x ≤ 0.8, indicating that the dominant magnetic interaction is AFM in this doping region. The fact that | θCW | >> TN/S implies 2D AFM ordering for x < 0.2, whereas a small θCW value at 0.2 ≤ x ≤ 0.8 is consistent with our interpretation that there is short-range AFM correlation. And θCW > 0 for 0.8 ≤ x < 0.98, which confirms the ferromagnetic interaction in this doping range. Although θCW is comparable to the corresponding transition temperature TFM, the effective magnetic moment is small, as shown in Fig. 4B, which implies that the FM ordering results mainly from Fe, even though Pd has to be the vehicle of coupling between Fe atoms (see discussion below). According to the single-crystal X-ray refinement results (Tables S1S3), the actual Fe concentration is very close to the nominal value for 0.8 ≤ x ≤ 0.98. We thus estimate the effective magnetic moment of Fe using the nominal Fe concentrations; this gives μeff = 1.60μB/Fe (x = 0.80), 2.42μB/Fe (x = 0.83), 3.79μB/Fe (x = 0.88), 4.21μB/Fe (x = 0.92), 3.53μB/Fe (x = 0.95), and 2.52μB/Fe (x = 0.98). Note that the highest magnetic moment for x = 0.92 is close to the Fe moment in Fe1.09Te ( 4.65μB/Fe), as plotted in Fig. 4B. In the latter case, the actual ordered magnetic moment is ∼3.31μB/Fe (28), smaller than that obtained from the Curie–Weiss constant; this is explained as due to the itinerant nature of the electrons in FeTe (28). For Fe1−xPdxTe, we may estimate the ordered magnetic moment from the magnetization. Fig. 4C shows the field dependence (H) of magnetization (M) at T = 3 K for x between 0.80 and 0.98. Note that M(H) reveals a well-defined hysteresis loop, confirming the nature of FM ordering for 0.8 ≤ x ≤ 0.98. Though M(H) is not saturated up to 6 T, we may estimate the lower bound of the ordered magnetic moment using M(H = 6 T) data shown in Fig. 4B. At T = 3 K, μ ∼0.33μB/Fe (x = 0.80), 0.48μB/Fe (x = 0.83), 1.49μB/Fe (x = 0.88), 1.85μB/Fe (x = 0.92), 1.67μB/Fe (x = 0.95), and 0.22μB/Fe (x = 0.98). In addition to the fact that M (H = 6 T) < Msat, the small ordered magnetic moment could result from (i) the itinerant nature of the electrons and (ii) the reduced concentration of Fe in the highly Pd doped region. Nevertheless, the specific heat reveals anomalies in the compounds with high TFM (Fig. 4D), indicating a true phase transition at TFM.Open in a separate windowFig. 4.Fe1-xPdxTe (A) Doping dependence of Curie–Weiss temperature. (B) Doping dependence of the effective magnetic moment and magnetization at 6 T. (C) Magnetization vs. magnetic field at 3 K for x = 0.8, 0.88, 0.92, 0.95, and 0.98. (D) Specific heat vs. temperature for x = 0.92 and 0.88.From the results presented here, it is most likely that the FM dome is intrinsic, due to the ordering of the Fe magnetic moments. Within the dome, our single-crystal X-ray refinement indicates that the structure remains the same as that of pure PdTe with actual Fe concentration close to the nominal value, and there are no interstitial sites of either Pd or Fe, which strongly suggests that the FM ordering is not due to the formation of Fe clusters. In early studies, FM was detected in alloys of Fe in Pd (i.e., Pd1−xFex) at compositions down to 0.08% Fe (48), because magnetic moments on the Fe sites polarize the surrounding Pd matrix (4851). It seems that Fe doping in PdTe compound gives rise to a similar effect, i.e., Pd matrix mediates the magnetic interactions between Fe atoms.The central question is, how can both FM ordering and superconductivity coexist in the region of 0.88 ≤ x ≤ 0.98? Having excluded the possibility of structural phase separation, one may consider the scenario of electronic phase separation with either macroscopic coexistence of singlet SC and FM in different regions or microscopic coexistence of triplet SC and FM (11, 21, 52). If Fe1−xPdxTe were a triplet superconductor, a small amount of impurity or disorder would completely suppress superconductivity, as seen in Sr2RuO4 (53). Though we observe anomalous specific heat below Tc of PdTe (42), the insensitivity to Fe doping does not support triplet SC. For the former case, further experimental investigations are necessary, such as combined scanning electron microscopy and tunneling microscopy to directly probe both SC and FM regions. Theoretically, almost all Pd d orbitals in PdTe are occupied due to the covalency of the Pd–Te bonding, according to recent first-principle calculations (44), and this leads to a strong suppression of the local magnetic moment (44). As shown in Fig. 3G (Inset), the magnetic susceptibility of PdTe (x = 1) indeed exhibits paramagnetic behavior above Tc. Upon substitution of Pd by Fe, Tc decreases nearly linearly with increasing Fe concentration (1 − x; Fig. 2). As Fe concentration reaches ∼3% (1 − x = 0.03), both SC and FM ordering coexist. The fact that FM ordering does not immediately kill superconductivity is likely due to the unique electronic structure of PdTe. First-principles calculations indicate that the states in the proximity of the Fermi level consist mainly of Te p electrons, which are weakly hybridized with Pd eg orbitals (44); this is in dramatic contrast to the electronic structure of FeTe, in which the states near the Fermi level derived from Fe with direct Fe–Fe interactions, and the Te p states lie well below the Fermi level and hybridized weakly with the Fe d states (54). For PdTe, the partial doping of Fe may lead to an even weaker hybridization between Te p and Pd eg orbitals, due to the shift of Fermi level. Our experimental observation is in support of this scenario: with increasing Fe concentration, superconductivity is gradually suppressed to zero, at which the FM ordering reaches the maximum (Figs. 2 and and4).4). After Tc reaches zero at xc, the disappearance of the superconducting energy gap allows the rearrangement of electronic structure, which is in favor of antiferromagnetic interaction; the latter apparently competes with the existing FM interaction, thus leading to the complete suppression of FM at x < 0.78.In summary, the substitution of Fe by Pd results in an extremely rich phase diagram for Fe1−xPdxTe. Powder and single-crystal X-ray diffraction measurements both indicate that there is doping-induced structural transition from a tetragonal phase at x < xs ∼0.6 to a hexagonal phase at x > xs. Correspondingly, the electrical resistivity changes from nonmetallic character at x < xs to metallic behavior at x > xs. Magnetically, the system undergoes a first-order AFM transition at TN/S for xxN/S ∼0.15, with the structure transition occurring at the same temperature from tetragonal to monoclinic. When TN/S approaches zero at xN/S, there is no longer a temperature-induced structural transition, whereas short-range AFM correlation exists for xN/S < x ≤ 0.88. Superconductivity sets in at xxc ∼0.88. In addition, a gigantic FM dome with 0.78 ≤ x ≤ 0.98 is centered at the critical concentration xc, where both AFM and SC vanish. The coexistence of FM ordering with either SR AFM or SC results most likely from the unique electronic structure of Fe1−xPdxTe. With the low Pd concentration, both electrical conduction and magnetism of Fe1−xPdxTe are determined by Fe/Pd d electrons, because they are near the Fermi level. Due to the ferromagnetic instability and the extended orbitals of 4d electrons of Pd, both AFM interactions and electrical resistivity in Fe1−xPdxTe decrease with increasing x. Before the complete suppression of SR AFM, FM ordering occurs due to the Fe magnetic moment polarization in the Pd matrix. However, the hexagonal structure at x > xs with Pd/Fe in an octahedral environment leads to more localized d electrons from Pd/Fe but itinerant p electrons from Te. It seems that the weak hybridization between pd electrons allows for the magnetic ordering of Fe/Pd moments and superconducting ordering of p conduction electrons. Nevertheless, these two orderings compete with each other. Hence, the physics of Fe1−xPdxTe in the hexagonal phase is in contrast to known Fe-based superconductors, in which physical properties are mainly determined by Fe d states. Fe1−xPdxTe provides another and rare example for studying the interplay between SC, FM, and AFM ordering, where two separate sets of electrons are responsible for FM ordering and superconductivity, respectively.  相似文献   
47.
BACKGROUND: Measures of aortic stiffness--aortic pulse wave velocity (PWV) and augmentation index (AIx)--have been shown to be powerful predictors of survival in adult haemodialysis (HD) patients. Very few data have been reported regarding arterial stiffness in paediatric renal populations. METHODS: PWV and aortic AIx were determined from contour analysis of arterial waveforms recorded by applanation tonometry using a SphygmoCor device in 14 children on HD (age = 14.1 years) and in 15 age, height matched children controls. RESULTS: Pre-HD AIx (29.7 +/- 15.4%) and PWV (6.6 +/- 1.0 m/s) were significantly higher compared with children controls (8.3 +/- 8.0% and 5.4 +/- 0.6 m/s, respectively, P < 0.0001). The only significant difference between normal and HD children was BP level: 103/61 vs 114/72 mmHg, P < 0.05. In children of HD patients, a multiple linear regression model including BP, age, height, weight, Ca and P levels as independent variables accounted for 57% of the variability in AIx. Dialysis had no impact on AIx (post-HD: 28.5 +/- 12.7%) or on PWV (post-HD: 6.7 +/- 0.8 m/s). CONCLUSIONS: We show, in this first-ever report of increased arterial stiffness in children on dialysis, that end-stage renal disease is associated with abnormalities in arterial wall elastic properties, comparable with adult levels, even in childhood. Most importantly, the absence of a discernible amelioration with dialysis implies that purely structural and not functional alterations lie behind the increased arterial stiffness.  相似文献   
48.
Ecotoxicological screening of dust sampled throughout a Kenyan tannery was conducted using a luminescence (lux)-based bacterial biosensor for both solid and liquid assays. This was complemented by chemical analysis in an attempt to identify possible causative toxic components. The biosensor results showed a highly significant (p < 0.001) difference in both solid and liquid phase toxicity in samples collected from various identified sampling points in the tannery. A positive correlation was observed between results of the solid and liquid phase techniques, for most of the sampling points indicating that the toxic contaminants were bioavailable both in the solid and liquid state. However, the results generally indicated toxicity associated with liquid phase except certain areas in solid phase such as chemical handling, buffing area and weighing. The most toxic tannery area identified was the weighing area (p < 0.001), showing the lowest bioluminescence for both the solid (0.38 +/- 2.21) and liquid phases (0.01 +/- 0.001). Chromium was the metal present in the highest concentration indicating levels higher than the stipulated regulatory requirement of 0.5 mg Cr/m3 for total Cr (highest Cr concentration was at chemical handling at 209.24 mg l(-1)) in all dust samples. The weighing area had the highest Ni concentration (1.87 mg l(-1)) and the chemical handling area showed the highest Zn concentration (31.9 mg l(-1)). These results raise environmental health concerns, as occupational exposure to dust samples from this site has been shown to give rise to elevated concentrations (above the stipulated levels) of chromium in blood, urine and some body tissues, with inhalation being the main route. Health and Safety Executive (HSE), UK, and American Conference of Governmental Industrial Hygienist (ACGIH) and National Institute for Occupational Safety and Health (NIOSH), USA stipulates an occupational exposure limit of 0.5 mg Cr/m3 (8 h TWA) for total chromium. However, schedule 1 of Controls of substances hazardous to health (COSHH) regulations developed by HSE, indicate 0.05 mg m3 (8 h TWA reference periods) to be the limit for Cr (VI) exposure. The exposure limit for individual (e.g., Cr, Zn, Ni etc.) contaminants (homogeneity) was not exceeded, but potential impact of heterogeneity (multi-element synergistic effect) on toxicity requires application of the precautionary principle.  相似文献   
49.
50.
We consider a system made up of different physical, chemical, or biological species undergoing replication, transformation, and disappearance processes, as well as slow diffusive motion. We show that for systems with net growth the balance between kinetics and the diffusion process may lead to fast, enhanced hydrodynamic transport. Solitary waves in the system, if they exist, stabilize the enhanced transport, leading to constant transport speeds. We apply our theory to the problem of determining the original mutation position from the current geographic distribution of a given mutation. We show that our theory is in good agreement with a simulation study of the mutation problem presented in the literature. It is possible to evaluate migratory trajectories from measured data related to the current distribution of mutations in human populations.  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