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   1994篇
  免费   115篇
  国内免费   10篇
耳鼻咽喉   9篇
儿科学   34篇
妇产科学   28篇
基础医学   362篇
口腔科学   33篇
临床医学   190篇
内科学   414篇
皮肤病学   34篇
神经病学   211篇
特种医学   132篇
外科学   186篇
综合类   12篇
一般理论   2篇
预防医学   170篇
眼科学   30篇
药学   132篇
中国医学   9篇
肿瘤学   131篇
  2024年   2篇
  2023年   20篇
  2022年   44篇
  2021年   66篇
  2020年   35篇
  2019年   53篇
  2018年   71篇
  2017年   50篇
  2016年   56篇
  2015年   63篇
  2014年   93篇
  2013年   133篇
  2012年   187篇
  2011年   175篇
  2010年   94篇
  2009年   77篇
  2008年   147篇
  2007年   143篇
  2006年   130篇
  2005年   125篇
  2004年   100篇
  2003年   88篇
  2002年   78篇
  2001年   12篇
  2000年   2篇
  1999年   7篇
  1998年   13篇
  1997年   9篇
  1996年   9篇
  1995年   4篇
  1994年   5篇
  1993年   6篇
  1992年   2篇
  1991年   2篇
  1989年   1篇
  1988年   1篇
  1987年   1篇
  1986年   2篇
  1985年   1篇
  1984年   2篇
  1983年   2篇
  1982年   2篇
  1981年   1篇
  1980年   1篇
  1974年   2篇
  1965年   1篇
  1958年   1篇
排序方式: 共有2119条查询结果,搜索用时 31 毫秒
21.
22.

Introduction

Non-tuberculous mycobacteria (NTM) are ubiquitous organisms associated with various infections. The aim of the study was to determine the most relevant clinical characteristics of NTM during the 7-year period.

Methodology

A retrospective study of NTM infections was conducted between January 2009 and December 2016. The American Thoracic Society/Infectious Disease Society of America criteria were used to define cases of pulmonary or an extrapulmonary site.

Results

A total of 85 patients were included in the study. Pulmonary cases predominated 83/85 (98%), while extrapulmonary NTM were present in 2/95 (2%) patients. Overall, ten different NTM species were isolated. The most common organisms were slow-growing mycobacteria (SGM) presented in 70/85 (82.35%) patients. Isolated SGM strains were Mycobacterium avium complex (MAC) in 25/85 (29.41%) patients, M. xenopi in 20/85 (23.53%) patients, M. kansasii in 15/85 (17.65%) patients and M. peregrinum and M. gordonae in 5/85 (5.88%) patients each. Isolated rapid-growing mycobacteria (RGM) strains were M. abscessus in 8/85 (9.41%) patients, M. fortuitum in 4/85 (4.71%) patients and M. chelonae in 3/85 (3.53%) patients. Almost all patients (98%; 83/85) had comorbidities. Among 75 (88.24%) patients who completed follow-up, 59 (69.41%), 10 (11.76%) and 6 (7%), were cured, experienced relapse and died, respectively.

Conclusion

In the present study, pulmonary NTM infections were more frequent compared to extrapulmonary disease forms. SGM were most common isolates with MAC pulmonary disease the most frequently found. Comorbidities have an important role in NTM occurrence. Further investigation should focus on an NTM drug susceptibility testing.
  相似文献   
23.
With laboratory and numerical work, we demonstrate that one of the main diffusion coefficients and the smaller eigenvalue of the Fick diffusion matrix are invariant to the number of methylene groups of the alcohol in ternary mixtures composed of an aromatic (benzene), a ketone (acetone) and one of three different alcohols (methanol, ethanol or 2-propanol). A critical analysis of the relationship between the kinetic and thermodynamic contributions to the diffusion coefficients allows us to explain this intriguing behaviour of this class of mixture. These findings are reflected by the diffusive behaviour of the according binary subsystems. Our approach provides a promising systematic framework for future investigations into the important and challenging problem of transport diffusion in multicomponent liquids.

The Fick diffusion coefficient matrix of three ternary mixtures composed of an aromatic (benzene), a ketone (acetone) and one of three different alcohols (methanol, ethanol or 2-propanol) is investigated with laboratory and numerical work.

Multicomponent diffusion plays a crucial role in various natural and industrial processes involving mass transfer.1–3 Liquids appearing in nature and technical applications are essentially multicomponent. However, only data on binary diffusion coefficients are relatively abundant because the diffusion behavior of ternary and higher mixtures is much more complex.4,5 Describing the isothermal–isobaric diffusion of a ternary mixture by Fick’s law requires four different diffusion coefficients that are composition dependent. The presence of cross diffusion coefficients aggravates the interpretation and data processing in experimental work, resulting in large uncertainties.6,7 Thus, efforts are being made to develop new methods for analysis of multicomponent diffusion explicitly addressing various degrees of complexity.8–10 Predictive equations for multicomponent diffusion of liquids mostly rely on extensions of the Darken relation,11–13 which is only valid for ideal mixtures.14 The underlying physical phenomena in non-ideal mixtures are not well understood and the lack of experimental data impedes the development and verification of new predictive equations.The objective of this study was not only to measure and predict the Fick diffusion coefficient matrix for a series of ternary liquid mixtures, rather, the emphasis lied on understanding common features and whether they can be related to the behavior of the pure components and binary subsystems. Three ternary mixtures that are composed of organic compounds were selected, i.e. an aromatic, a ketone and an alcohol. Throughout, the first two components were benzene (1) and acetone (2) and the third component was one of the alcohols, methanol, ethanol or 2-propanol. For each mixture, nine state points along a composition path with a constant content of benzene, x1 = 0.33 mol mol−1, were studied under ambient conditions (298.15 K and 0.1 MPa). Seven of the state points were ternary mixtures and two were binary subsystems. To obtain reliable results for the Fick diffusion coefficient matrix, two complementary approaches were used, i.e. experiments and predictive molecular simulations. This combination allows for a critical analysis and leads to a deeper understanding of the underlying phenomena.14,15The Taylor dispersion technique was utilized for the experiments.16,17 In this method, a small quantity of mixture with a slightly different composition is injected into a laminar stream. It disperses due to convection and diffusion while flowing through a capillary tube and the refractive index is measured at its end to sample the concentration distribution. We have used the same apparatus as in previous works.6,7 The Fick diffusion matrix is obtained by fitting working equations to the measured signal, i.e. the Taylor peak. The mathematical model of the Taylor dispersion technique was originally developed on the basis of Fick’s law in the volume reference frame. In a ternary mixture, two molar fluxes Jvi relative to a volume averaged velocity are related to gradients of molar concentration ∇Ci with four diffusion coefficients Dvij. Alternatively, fluxes expressed in the molar reference frame Ji are relative to a molar averaged velocity and the mole fraction gradients ∇xi act as a driving force1with molar density ρ. The fluxes of all three components are constrained by ΣJi = 0. The main diffusion coefficients D11 and D22 relate the flux of one component to its own mole fraction gradient and the cross diffusion coefficients D12 and D21 describe the coupling of the flux of one component with the gradient of the other. The third component does not appear in eqn (1) explicitly, but in general it affects all four diffusion coefficients. The transformation of experimental data from the volume to the molar reference frame (Dvij to Dij) could be done here on the basis of the pure component volumes (see the ESI).Equilibrium molecular dynamics (MD) simulations were employed in this work, allowing for examination at the microscopic scale. The underlying molecular models were rigid, non-polarizable force fields of united atom type, consisting of a varying number of Lennard–Jones, point charge, dipole and quadrupole sites (see the ESI). Note that the force field parameters were adjusted to pure fluid properties only so that all simulation results for the mixtures are strictly predictive. Diffusion coefficients were sampled with the Green–Kubo formalism, based on integrated correlation functions of net velocities of the contained species.11,15 Thereby, phenomenological coefficients Δij were obtained, associating the diffusive fluxes with the chemical potential gradients ∇μi2with gas constant R and temperature T. Fluxes Ji correspond to the molar reference frame as in eqn (1).The diffusion coefficients from experiment and simulation are related to different driving forces so that the chemical potential gradients have to be transformed to the mole fraction gradients for their comparison.18 This transformation is contained in the thermodynamic factor matrix Γ3with the activity coefficient of species i being γi, which expresses the non-ideality of a mixture with respect to the composition. This relationship shows that the Fick diffusion coefficients are actually the product of two contributions, a kinetic Δij and a thermodynamic Γij. The separate observation of these two contributions promotes understanding of the underlying physical phenomena. In the present study, the thermodynamic factor was calculated using the Wilson excess Gibbs energy (gE) model, using parameters fitted to experimental vapor–liquid equilibrium data of the binary subsystems (see the ESI). This combination of MD simulation results with a gE model was successfully used in previous work to predict Fick diffusion coefficients, including several binary subsystems of the ternary mixtures studied here.19The four elements of the Fick diffusion coefficient matrix were determined for the three ternary mixtures, benzene + acetone + methanol/ethanol/2-propanol, for nine different compositions, each at ambient temperature and pressure.Results for the first main element of the diffusion matrix D11, which relates the flux of benzene to its own mole fraction gradient, are shown in Fig. 1(a). The experimental data agree quantitatively with the molecular simulation data. D11 increases with the acetone content in the ternary mixture. Since mixtures with a constant mole fraction of benzene (x1 = 0.33 mol mol−1) were studied throughout, the left edge of Fig. 1(a) corresponds to the binary limit of benzene + alcohol, while the right edge corresponds to that of benzene + acetone. Analysis of the ternary diffusive fluxes implies the following asymptotic behavior of the diffusion coefficients towards the binary limits:7 (i) at the infinite dilution limit, x2 → 0, the ternary coefficient D11 tends to the binary Fick diffusion coefficient of benzene + alcohol; (ii) at the other limit, x3 → 0, D11D12 = D22D21Dbin (benzene + acetone) should hold. The present experimental and simulation results for D11 are consistent with these asymptotic limits.Open in a separate windowFig. 1Top: The main Fick diffusion coefficient (molar reference frame) of benzene D11 in the three ternary mixtures benzene (1) + acetone (2) + alcohol (3) at a constant benzene mole fraction x1 = 0.33 mol mol−1 from experiment (triangles) and MD simulation combined with the Wilson gE model (circles). Both data sets were sampled at the same compositions, but are slightly shifted in the plot for visibility reasons. The symbols at the edges of this plot are the binary diffusion coefficients of benzene + alcohol (x2 → 0) and of benzene + acetone (x3 → 0). Bottom: The binary Fick diffusion coefficient of the subsystems benzene + alcohol and benzene + acetone. Most of the binary experimental data were taken from the literature.20–27An inspection of Fig. 1(a) provides an unexpected finding: the main element D11 is almost identical for all three mixtures along the examined composition path, i.e. it is independent of the contained type of alcohol. To explain this intriguing behavior of D11, the properties of the pure components are considered first (see
M (g mol−1) ρ (mol l−1) ρ m (g l−1) D 0 10−9 (m2 s−1)
Benzene78.1111.147 (2)870.6 (1)2.226 (4)
Acetone58.0813.536 (3)786.2 (2)4.538 (8)
Methanol32.0424.541 (6)786.3 (2)2.449 (6)
Ethanol46.0717.132 (4)789.3 (2)0.974 (3)
2-Propanol60.1012.803 (1)769.5 (1)0.604 (7)
Open in a separate windowAll five components are liquid under ambient conditions so that their self-diffusion coefficients are of the same order of magnitude. Molar masses M and molar densities ρ, indicating the differences in mass and size of the molecules, give an introductory idea of their diffusion behavior. Benzene molecules are heavier and larger than acetone molecules, resulting in a self-diffusion coefficient D0 that is only about half that of acetone. The three alcohols are characterized by increasing mass and size in the order: methanol, ethanol, 2-propanol. Although the methanol molecules are the smallest, due to hydrogen bonding interactions, their self-diffusion coefficient is similar to that of benzene. Molecules associated by hydrogen bonds propagate as an assembly, which significantly slows down their mobility. This is not only the case for methanol, but also for ethanol and 2-propanol. Because these molecules are larger, the bonded clusters are also larger and thus even slower. This hydrogen bonding behavior of the alcohol molecules causes micro-heterogeneity and cluster formation in mixtures with other fluids,28,29 which influences their kinetic and thermodynamic behavior.Next, the binary subsystems of the ternary systems with different types of alcohol were examined. The Fick diffusion coefficient of the three binary benzene + alcohol mixtures and that of benzene + acetone is shown in Fig. 1(b). The benzene mole fraction, x1 = 0.33 mol mol−1, which was constant along the ternary composition path, is marked in the plot by a dashed vertical line. The binary Fick diffusion coefficient of all three benzene + alcohol mixtures has almost the same value in the concentration range around equimolar composition. However, at both infinite dilution limits (x1 → 0 and x1 → 1), the benzene + methanol system has a higher Fick diffusion coefficient than the benzene + ethanol or benzene + 2-propanol systems. The open question is why the Fick diffusion coefficients are similar in a wide composition range. Although these binary data are available in the literature, to the best of our knowledge, they have never been discussed from this point of view.In a binary mixture there is only a single Fick diffusion coefficient and eqn (3) reduces toD = ĐΓ,4where Đ is the Maxwell–Stefan (MS) diffusion coefficient. Đ represents the kinetic contribution to the diffusion behavior, which was sampled here using MD simulations from net velocity correlation functions, while Γ corresponds to the thermodynamic non-ideality, which was calculated using the Wilson gE model. Both contributions are separately shown in Fig. 2(a) and (c) for the three binary benzene + alcohol mixtures. The largest kinetic contribution, i.e. the MS diffusion coefficient, appears for benzene + methanol, followed by benzene + ethanol, which is also slightly larger than that of benzene + 2-propanol (see Fig. 2(a)). The same order was observed for the self-diffusion coefficient of the pure alcohols, which also decreases from methanol over ethanol to 2-propanol. The non-ideal composition dependence of the MS diffusion coefficient is a consequence of the hydrogen bonding behavior of the alcohols. The formation of clusters causes a correlated propagation of molecules. This leads to significant contributions of velocity correlations between unlike molecules,12,14,30 which are considered MS diffusion coefficient sampling (see the ESI). The thermodynamic factor exhibits the converse order: benzene mixed with methanol is the most non-ideal with the smallest thermodynamic factor, followed by ethanol and 2-propanol. Multiplying these two contributions leads to a similar Fick diffusion coefficient over a wide composition range of the three binary mixtures.Open in a separate windowFig. 2(a) The Maxwell–Stefan diffusion coefficient Đ of the three binary mixtures benzene + alcohol, (b) the phenomenological coefficient Δ11 of the ternary mixtures from MD simulation, (c) the thermodynamic factor Γ of the three binary mixtures and (d) the thermodynamic factor Γ11 of the ternary mixtures from the Wilson gE model.Building on this understanding, we further demonstrate that a similar interplay between kinetic and thermodynamic contributions is responsible for the independence of D11 of the alcohol type for ternary mixtures of benzene and acetone with methanol, ethanol or 2-propanol. It follows from eqn (3) that D11 = Δ11Γ11 + Δ12Γ21. Molecular simulation data show that the first term dominates the sum, while the second term is negligibly small. The kinetic Δ11 and thermodynamic Γ11 contributions of the first term are depicted in Fig. 2(b) and (d). Indeed, as in the binary case, methanol exhibits the highest kinetic and the lowest thermodynamic contribution, providing that the product Δ11Γ11 is the same for all considered types of alcohol. It can thus be concluded that the interplay between kinetics and thermodynamics leads to similar binary and ternary diffusion coefficients for mixtures of benzene and acetone with methanol, ethanol or 2-propanol. To examine the clustering behavior of the alcohols in the ternary mixtures, hydrogen bonding statistics were sampled using molecular simulations on the basis of geometric criteria31 (see Fig. 3). Most of the alcohol molecules are bonded to dimers and trimers within the ternary mixtures. The fractions of monomers, dimers, trimers and tetramers are almost identical for all three alcohols.Open in a separate windowFig. 3Hydrogen bonding statistics obtained from MD simulation in the three ternary mixtures benzene + acetone + alcohol, i.e. methanol (red), ethanol (blue) and 2-propanol (green), at a constant benzene mole fraction x1 = 0.33 mol mol−1.An important remaining question is whether the quantitative similarity of the binary and ternary diffusion coefficients can also relate to the second main Fick diffusion coefficient of the studied ternary mixtures. The diffusion coefficient D22, characterizing the diffusive flux of acetone under its own mole fraction gradient, is shown in Fig. 4(a). The presence of benzene affects D22, resulting in a less steep increase of that coefficient with higher acetone content. On average, D22 is 1.5 to 2 times larger than D11, which is in agreement with the twice as large self-diffusion coefficient of acetone compared to that of benzene. D22 is fairly similar for ethanol and 2-propanol and noticeably higher for methanol. The binary diffusion coefficient of acetone + alcohol, shown in Fig. 3(b), resembles the behavior of D22 in the ternary mixtures. As in the preceding discussion of D11 and the corresponding binary subsystems, we decomposed the diffusion coefficient D22 = Δ21Γ12 + Δ22Γ22 into its kinetic and thermodynamic contributions. Molecular simulation data show that the cross term Δ21Γ12 is again negligibly small. The kinetic contributions for the ternary Δ22 as well as for the binary Đ (acetone + alcohol) mixtures are identical in the case of ethanol and 2-propanol, but much larger in the case of methanol. However, here the thermodynamic contributions for mixtures with methanol (Γ22 and Γ) cannot compensate for the large kinetic values. Separate analysis of kinetics and thermodynamics is a novel way for understanding diffusion.Open in a separate windowFig. 4Top: The main Fick diffusion coefficient (molar reference frame) of acetone D22 in the three ternary mixtures benzene (1) + acetone (2) +alcohol (3) at a constant benzene mole fraction x1 = 0.33 mol mol−1 from experiment (triangles) and MD simulation combined with the Wilson gE model (circles). Both data sets were sampled at the same compositions, but are slightly shifted in the plot for visibility reasons. Bottom: The binary Fick diffusion coefficient of the subsystems acetone + alcohol and acetone + benzene. Most of the binary experimental data were taken from the literature.20,24,25We may thus draw the conclusion that for the liquid ternary mixtures benzene + acetone + alcohol, the qualitative behavior of the main coefficients D11 and D22 can directly be related to the binary subsystems, including the influence of contained alcohols on the composition dependent diffusion coefficients.An important feature of ternary diffusion are the cross effects that cannot be related to binary behavior. As is often the case, the two cross coefficients of the studied ternary mixtures are significantly smaller than the main ones. The cross coefficient of benzene D12, relating the flux of benzene to the mole fraction gradient of acetone, has mostly small negative values for all three ternary mixtures, except for small alcohol concentrations in the mixture with methanol, where it is positive. The second cross coefficient of acetone D21 must be zero at the limit x2 → 0, which is confirmed by the trend of the data. At the other limit x3 → 0, the coefficients are positive and increasing towards the limit of diluted alcohol, with the highest values in the mixture with methanol and the lowest for 2-propanol.In contrast to the individual elements of the Fick diffusion coefficient matrix, the eigenvalues of the matrix do not depend on the reference frame or on the order of components. Furthermore, a constraint imposed by the second law of thermodynamics is that the eigenvalues of the Fick diffusion coefficient matrix must be real and positive for a thermodynamically stable mixture. The eigenvalues of the diffusion matrix obtained by experiment and simulation fulfil these specifications. They show the same variation with composition and dependence on the type of alcohol, which was already observed for the main elements of the diffusion matrix (see Fig. 5). The larger eigenvalue D1 increases with acetone content and shows higher values in the ternary mixture with methanol, while it is slightly lower for ethanol and 2-propanol. This correlates with the behavior of D22. The smaller eigenvalue D2, like the main coefficient D11, is independent of the type of alcohol for the three studied ternary mixtures.Open in a separate windowFig. 5Eigenvalues of the Fick diffusion coefficient matrix of the three ternary mixtures benzene (1) + acetone (2) + alcohol (3) at a constant benzene mole fraction x1 = 0.33 mol mol−1 from experiment (triangles) and MD simulation combined with the Wilson gE model (circles).Fick diffusion coefficients of three different ternary mixtures, i.e. benzene + acetone + methanol/ethanol/2-propanol, were analyzed. Two complementary approaches were utilized to obtain reliable data, experiments and molecular simulation. We identified an important feature of this class of mixture (an aromatic, a ketone and an alcohol): namely that one of the main diffusion coefficients D11, where D11 < D22, and the smaller eigenvalue D2 are independent of the alcohol type along the studied composition path. This insight was reflected in another finding that the Fick diffusion coefficient of the binary benzene + alcohol subsystems also does not depend on the alcohol type. The underlying mechanism of this unusual behavior was explained by separately considering the kinetic and thermodynamic contributions to the diffusion coefficients. The results presented here provide a promising framework for future systematic investigations into the important and challenging problem of diffusion in multicomponent liquid mixtures. In order to provide a more substantial understanding of phenomena occurring in multicomponent mixtures, the present study can be continued and extended by replacing one main component of the ternary mixture, e.g. benzene, with another aromatic substance, e.g. toluene.  相似文献   
24.
Peripheral bronchial identification on chest CT using unsupervised machine learning     
Daniel A. Moses  Laughlin Dawes  Claude Sammut  Tatjana Zrimec 《International journal of computer assisted radiology and surgery》2018,13(9):1379-1395

Purpose

To automatically identify small- to medium-diameter bronchial segments distributed throughout the lungs.

Methods

We segment the peripheral pulmonary vascular tree and construct cross-sectional images perpendicular to the lung vasculature. The bronchi running with pulmonary arteries appear as concentric rings, and potential center points that lie within the bronchi are identified by looking for circles (using the circular Hough transform) and rings (using a novel variable ring filter). The number of candidate bronchial center points are further reduced by using agglomerative hierarchical clustering applied to the points represented with 18 features pertaining to their 3D position, orientation and appearance of the surrounding cross-sectional image. Resulting clusters corresponded to bronchial segments. Parameters of the algorithm are varied and applied to two experimental data sets to find the best values for bronchial identification. The optimized algorithm was then applied to a further 21 CT studies obtained using two different CT vendors.

Results

The parameters that result in the most number of true positive bronchial center points with > 95% precision are a tolerance of 0.15 for the hierarchical clustering algorithm and a threshold of 75 HU with 10 spokes for the ring filter. Overall, the performance on all 21 test data sets from CT scans from both vendors demonstrates a mean number of 563 bronchial points detected per CT study, with a mean precision of 96%. The detected points across this group of test data sets are relatively uniformly distributed spatially with respect to spherical coordinates with the origin at the center of the test imaging data sets.

Conclusion

We have constructed a robust algorithm for automatic detection of small- to medium-diameter bronchial segments throughout the lungs using a combination of knowledge-based approaches and unsupervised machine learning. It appears robust over two different CT vendors with similar acquisition parameters.
  相似文献   
25.
Proinflammatory cytokines inhibit the expression and function of human type I 5'-deiodinase in HepG2 hepatocarcinoma cells     
Jakobs TC  Mentrup B  Schmutzler C  Dreher I  Köhrle J 《European journal of endocrinology / European Federation of Endocrine Societies》2002,146(4):559-566
  相似文献   
26.
Age-associated accumulation of CMV-specific CD8+ T cells expressing the inhibitory killer cell lectin-like receptor G1 (KLRG1)   总被引:3,自引:0,他引:3  
Ouyang Q  Wagner WM  Voehringer D  Wikby A  Klatt T  Walter S  Müller CA  Pircher H  Pawelec G 《Experimental gerontology》2003,38(8):911-920
Large clonal expansions of peripheral CD8+ T cells carrying receptors for single epitopes of CMV and EBV are common in the elderly and may be associated with an immune risk phenotype predicting mortality. Here we show that the frequency of CD8+ T cells expressing the inhibitory killer cell lectin-like receptor G1 (KLRG1), a marker of cells unable to undergo further clonal expansion, was markedly elevated in CD8+ T cells from old donors. Moreover, tetramer staining revealed that the elevated frequency of CMV-specific CD8+ T cells in the elderly was due to an accumulation of cells bearing this dominant negative receptor. The fraction of CMV-specific T cells able to secrete interferon-gamma after specific antigenic stimulation was significantly lower in the elderly than in the young, although the total number of functional cells was comparable. Therefore, the majority of the clonally expanded virus-specific CD8+ cells in the elderly was dysfunctional. Thus, T cell responses are altered in the aged by an accumulation of replicatively senescent dysfunctional T cells carrying receptors for persistent herpes viruses. The presence of clonal expansions of such virus-specific cells may shrink the available repertoire for other antigens and contribute to the increased incidence of infectious disease in the elderly.  相似文献   
27.
Myeloablative intensified conditioning regimen with in vivo T-cell depletion (ATG) followed by allografting in patients with advanced multiple myeloma. A phase I/II study of the German Study-group Multiple Myeloma (DSMM)     
Kröger N  Einsele H  Wolff D  Casper J  Freund M  Derigs G  Wandt H  Schäfer-Eckart K  Wittkowsky G  Schmitz N  Krüger W  Zabelina T  Renges H  Ayuk F  Krüll A  Zander A;German Study-group Multiple Myeloma 《Bone marrow transplantation》2003,31(11):973-979
We investigated toxicity and efficacy of in vivo T-cell depletion with anti-thymocyte globulin (ATG) as part of an intensified myeloablative conditioning regimen followed by allogeneic stem cell transplantation in patients with advanced multiple myeloma. The conditioning regimen consisted of modified total body irradiation, busulfan and cyclophosphamide (n=15) or in the case of prior dose-limiting radiotherapy of busulfan and cyclophosphamide (n=3). The median age was 44 years (range, 29-53) and the median time from diagnosis to transplant was 12 months (range, 6-144). Grade II-IV acute graft-versus-host disease (GvHD) occurred in six patients (35%). Severe grade III/IV GvHD developed in one patient (6%). Three patients died of therapy-related causes (17%). A complete remission (CR) with negative immunofixation after allogeneic transplantation was seen in eight of the evaluable patients (53%). After a median follow-up of 41 months (range, 8-84), the estimated overall survival at 6 years for all patients is 77% (CI 95%: 58-96%). The estimated progression-free survival at 6 years for all patients is 31% (CI 95%: 2-59%) and 46% (CI 95%: 9-83%) for patients with CR. In vivo T-cell depletion with ATG resulted in a low rate of severe GvHD with low treatment-related mortality, and a substantial number of long-term survivors.  相似文献   
28.
18F‐Fluorodeoxyglucose (FDG)‐PET features of focal nodular hyperplasia (FNH) of the liver     
Amir Kurtaran  Alexander Becherer  Franz Pfeffel  Christian Müller  Tatjana Traub  Jrn Schmaljohann  Klaus Kaserer  Markus Raderer  Wolfgang Schima  Robert Dudczak  Kurt Kletter  Irene Virgolini 《Liver international》2000,20(6):487-490
Abstract: Aim: The aim of this paper is to describe the imaging pattern of focal nodular hyperplasia (FNH) by 18F‐fluorodeoxyglucose (18F‐FDG) positron emission tomography (PET). Methods: Eight consecutive asymptomatic patients with histologic proof of FNH underwent 18F‐FDG PET imaging. The lesions were found incidentally. The 18F‐FDG PET imaging was performed with a dedicated PET tomograph after intravenous injection of 300–370 MBq 18F‐FDG. The 18F‐FDG accumulation in the lesions was (semi)quantified by calculating the standardized uptake value (SUV) and SUV has been corrected for the lean body mass (LBM). Eight patients with liver metastases spread from melanoma (n=2) and colorectal carcinoma (n=6) served as controls. The size of the FNH lesions and of the control group ranged from 2.0 to 8.5 cm (mean 4.83 cm±2.37) and from 1.5 to 6 cm (mean 3.28±1.52), respectively. Results: While in malignant liver lesions the accumulation of 18F‐FDG was significantly increased, all FNH lesions showed normal or even decreased accumulation of 18F‐FDG. In FNH lesions, SUV ranged between 1.5 and 2.6 (mean 2.12±0.38), whereas all liver metastases showed an increased SUV ranging between 6.20 and 16.00 (mean 10.07±3.79). The SUV corrected for LMB (SUVLBM) was similar to the SUV and ranged between 0.9 and 2.2 (mean 1.81±0.41) for FNH and between 5.9 and 16.3 (mean 9.15±4.03), respectively. Conclusion: In contrast to liver metastases, there is no increased glucose metabolism in FNH in vivo. The imaging feature of FNH by 18F‐FDG‐PET imaging is not specific for FNH; however, it may be helpful to differentiate FNH from liver metastases in cancer patients if radiological methods are not diagnostic.  相似文献   
29.
Interstitial Capillary in Normal and in Transplanted Kidneys: An Ultrastructural Study     
Anastazija Hvala  Dušan Ferluga  Tomaž Rott  Tatjana Kobenter  Mira Koselj-Kajtna  Staša Kaplan-Pavlovčič 《Ultrastructural pathology》2013,37(4):295-299
Knowledge about the normal structure and pathology of interstitial capillary is limited. Splitting and multilayering of the basal membrane (BM), as a marker of chronic rejection, has been published in association with transplant glomerulopathy. The authors investigated the ultrastructural features of the interstitial capillary basal membrane in normal (15 biopsies) and in transplanted kidneys (27 biopsies from 21 patients), expressing transplant glomerulopathy (8 biopsies from 6 patients), acute tubulo-interstitial rejection (9 biopsies from 6 patients), and recurrent or de novo glomerulonephritis (10 biopsies from 8 patients). All biopsies were fixed in 1%OsO 4, embedded in Epon, and examined by electron microscope. Measurements of the interstitial capillary BM were made. The BM of interstitial capillary of intact kidney was a homogenous continuous structure, 88 nm in width on average. Thickening with diffuse multilayering of BM was most intensive in patients with transplant glomerulopathy, and much less intensive in patients with acute tubulointerstitial rejection and in patients with recurrent or de novo glomerulonephritis. These findings may provide the first information about the morphology of the normal basal lamina of interstitial capillary and support the diagnostic value of interstitial capillary changes in chronic rejection.  相似文献   
30.
On hemispheric specialisation and visual field effects in the perception of print: A comment on Jordan,Patching, and Thomas     
Tatjana A. Nazir 《Cognitive neuropsychology》2013,30(1):73-80
Introduction Due to structural characteristics of the visual pathways, stimuli that are presented in the right half of the visual field (RVF) are initially projected to the left cerebral hemisphere, while those presented in the left half of the visual field (LVF) are projected to the right cerebral hemisphere. This anatomical feature has frequently been taken to support the notion that the well-documented RVF advantage in recognising printed words is a reflection of functional differences between the two hemispheres; notably that of the dominance of the left hemisphere for processing language. Word stimuli that are sent straight to the left hemisphere are believed to profit from more efficient processing than those sent initially to the right hemisphere, because the latter stimuli must follow a longer and more noisy pathway before reaching the language-dominant hemisphere. In the work by Jordan, Patching, and Thomas (2003) the above notion is further developed to speculate that the point of entry of visual information into the cortex may determine the procedure that will underlie the ensuing word recognition process: "... the left hemisphere can process words by mapping orthographic information in parallel onto lexical entries whereas the right hemisphere has a more rudimentary process, that can only map orthographic information sequentially" (p. 50).  相似文献   
[首页] « 上一页 [1] [2] 3 [4] [5] [6] [7] [8] [9] [10] [11] 下一页 » 末  页»
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