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   3609篇
  免费   275篇
  国内免费   26篇
耳鼻咽喉   25篇
儿科学   86篇
妇产科学   77篇
基础医学   684篇
口腔科学   92篇
临床医学   225篇
内科学   897篇
皮肤病学   98篇
神经病学   302篇
特种医学   74篇
外科学   442篇
综合类   25篇
一般理论   3篇
预防医学   212篇
眼科学   88篇
药学   283篇
中国医学   9篇
肿瘤学   288篇
  2024年   6篇
  2023年   45篇
  2022年   122篇
  2021年   173篇
  2020年   102篇
  2019年   119篇
  2018年   143篇
  2017年   103篇
  2016年   102篇
  2015年   141篇
  2014年   150篇
  2013年   170篇
  2012年   265篇
  2011年   254篇
  2010年   160篇
  2009年   126篇
  2008年   244篇
  2007年   214篇
  2006年   216篇
  2005年   225篇
  2004年   190篇
  2003年   174篇
  2002年   143篇
  2001年   30篇
  2000年   24篇
  1999年   36篇
  1998年   34篇
  1997年   30篇
  1996年   17篇
  1995年   17篇
  1994年   11篇
  1993年   8篇
  1992年   8篇
  1991年   9篇
  1990年   9篇
  1989年   14篇
  1987年   9篇
  1986年   5篇
  1985年   6篇
  1984年   9篇
  1983年   5篇
  1982年   3篇
  1981年   4篇
  1980年   4篇
  1979年   3篇
  1978年   4篇
  1973年   2篇
  1971年   4篇
  1970年   6篇
  1967年   3篇
排序方式: 共有3910条查询结果,搜索用时 31 毫秒
41.
42.
43.
44.
45.
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   
46.
Clinical Implementation and Validation of Automated Human Genome Variation Society (HGVS) Nomenclature System for Next-Generation Sequencing–Based Assays for Cancer     
Keith M. Callenberg  Lucas Santana-Santos  Liang Chen  Wayne L. Ernst  Michelle B. De Moura  Yuri E. Nikiforov  Marina N. Nikiforova  Somak Roy 《The Journal of molecular diagnostics : JMD》2018,20(5):628-634
  相似文献   
47.
Cross‐cultural examination of beliefs about the causes of bulimia nervosa among Australian and Japanese females          下载免费PDF全文
Rachel Dryer PhD  Yuri Uesaka PhD  Emmanuel Manalo PhD  Graham Tyson PhD 《The International journal of eating disorders》2015,48(2):176-186
  相似文献   
48.
The length‐dependent activation of contraction is equally impaired in impuberal male and female rats in monocrotaline‐induced right ventricular failure          下载免费PDF全文
Oleg Lookin  Alexander Balakin  Daniil Kuznetsov  Yuri Protsenko 《Clinical and experimental pharmacology & physiology》2015,42(11):1198-1206
The length‐dependent activation of contraction is attenuated in the failing myocardium of adult male rats. This pathological change is not seen in adult female rats, possibly because of a protective effect of sex hormones. The present study evaluated length‐dependent changes in isometric twitch, Ca2+ transient (CaT) and action potential (AP) in the right ventricular myocardium of impuberal healthy male and female rats (control) and in rats treated with a single injection of 50 mg/kg monocrotaline (MCT). Compared with sex‐matched control rats, MCT‐treated male and female rats exhibited increased right ventricular weight (134% and 142% of control, respectively), decreased left ventricular weight (72% and 79%), twitch attenuation (48.8 ± 2.7% and 57.5 ± 1.2%) and prolongation (125 ± 3% and 127 ± 2%), CaT attenuation (37.8 ± 0.4% and 39.1 ± 1.1%) and prolongation (114 ± 1% and 116 ± 1%) and AP prolongation at 90% repolarization (195 ± 2% and 203 ± 1%). The MCT‐treated male rats exhibited a 50% lower integral magnitude and an approximately 25% larger time‐to‐peak ‘bump’ compared with control male rats. These parameters in MCT‐treated female rats tended to show similar changes to those seen in the control female rats, with no significant difference between the two groups. In all groups, integral magnitude and time‐to‐peak ‘bump’ increased with length. In conclusion, the length‐dependent activation of contraction was equally blunted in the failing right ventricular myocardium of impuberal male and female rats. This was related to changes in CaT and AP, which were similar between male and female rats. Therefore, puberty is necessary for manifestation of the protective effects of sex hormones on this remodelling.  相似文献   
49.
High accumulation of arsenic in the esophagus of mice after exposure to arsenite     
Daigo Sumi  Miyu Tsurumoto  Yuri Yoshino  Masahisa Inoue  Takehiko Yokobori  Hiroyuki Kuwano  Seiichiro Himeno 《Archives of toxicology》2015,89(10):1751-1758
  相似文献   
50.
Ventilatory efficiency during ramp exercise in relation to age and sex in a healthy Japanese population     
Kohei Ashikaga  Haruki Itoh  Tomoko Maeda  Hidetaka Itoh  Yuri Ichikawa  Shiori Tanaka  Ryuichi Ajisaka  Akira Koike  Shigeru Makita  Kazuto Omiya  Yuko Kato  Hitoshi Adachi  Masatoshi Nagayama  Akihiko Tajima  Naomi Harada  Yoshihiro J Akashi 《Journal of cardiology》2021,77(1):57-64
BackgroundThe current understanding of ventilator efficiency variables during ramp exercise testing in the normal Japanese population is insufficient, and the responses of tidal volume (VT) and minute ventilation (V?E) to the ramp exercise test in the normal Japanese population are not known.MethodsA total of 529 healthy Japanese subjects aged 20–78 years underwent cardiopulmonary exercise testing using a cycle ergometer with ramp protocols. VT and V?E at rest, at anaerobic threshold, and at peak exercise were determined. The slope of V?E versus carbon dioxide (V?CO2) (V?E vs. V?CO2 slope), minimum V?E/V?CO2, and oxygen uptake efficiency slope (OUES) were determined.ResultsFor males and females in their 20 s, peak VT (VTpeak) was 2192 ± 376 and 1509 ± 260 mL (p < 0.001), peak V?E (V?Epeak) was 80.6 ± 18.7 and 57.7 ± 13.9 L/min (sex differences p < 0.001), the V?E vs. V?CO2 slope was 24.4 ± 3.2 and 25.7 ± 3.2 (p = 0.035), the minimum V?E/V?CO2 was 24.2 ± 2.3 and 27.0 ± 2.8 (p < 0.001), and the OUES was 2452 ± 519 and 1991 ± 315 (p < 0.001), respectively. VTpeak and V?Epeak decreased with age and increased with weight and height. The V?E vs. V?CO2 slope and minimum V?E/V?CO2 increased with age, while conversely, the OUES decreased with age.ConclusionsWe have established the normal range of VT and V?E responses, the V?E vs. V?CO2 slope, the minimum V?E/V?CO2, and the OUES for a healthy Japanese population. Some of these parameters were influenced by weight, height, sex, and age. These results provide useful reference values for interpreting the results of cardiopulmonary exercise testing in cardiac patients.  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