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   6215篇
  免费   386篇
  国内免费   56篇
耳鼻咽喉   61篇
儿科学   159篇
妇产科学   94篇
基础医学   1073篇
口腔科学   348篇
临床医学   467篇
内科学   1431篇
皮肤病学   189篇
神经病学   548篇
特种医学   116篇
外科学   560篇
综合类   46篇
一般理论   2篇
预防医学   516篇
眼科学   113篇
药学   459篇
中国医学   50篇
肿瘤学   425篇
  2024年   8篇
  2023年   114篇
  2022年   225篇
  2021年   353篇
  2020年   204篇
  2019年   280篇
  2018年   272篇
  2017年   187篇
  2016年   258篇
  2015年   249篇
  2014年   287篇
  2013年   355篇
  2012年   544篇
  2011年   573篇
  2010年   279篇
  2009年   213篇
  2008年   370篇
  2007年   343篇
  2006年   308篇
  2005年   315篇
  2004年   250篇
  2003年   243篇
  2002年   191篇
  2001年   23篇
  2000年   20篇
  1999年   22篇
  1998年   48篇
  1997年   25篇
  1996年   14篇
  1995年   11篇
  1994年   11篇
  1993年   10篇
  1992年   7篇
  1991年   7篇
  1990年   3篇
  1989年   1篇
  1988年   4篇
  1987年   4篇
  1986年   1篇
  1985年   2篇
  1984年   2篇
  1983年   4篇
  1982年   4篇
  1981年   5篇
  1980年   2篇
  1979年   2篇
  1977年   1篇
  1975年   2篇
  1974年   1篇
排序方式: 共有6657条查询结果,搜索用时 31 毫秒
71.
Metabolic Brain Disease - Anxiety Disorders and Posttraumatic Stress Disorders (PTSD) associated with type-1 diabetes mellitus (T1DM) are increasingly common comorbidities and the treatment is...  相似文献   
72.
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   
73.
Polymer-Mediated Inhibition of Pro-invasive Nucleic Acid DAMPs and Microvesicles Limits Pancreatic Cancer Metastasis     
Ibtehaj Naqvi  Ruwan Gunaratne  Jessica E. McDade  Angelo Moreno  Rachel E. Rempel  Douglas C. Rouse  Silvia Gabriela Herrera  David S. Pisetsky  Jaewoo Lee  Rebekah R. White  Bruce A. Sullenger 《Molecular therapy》2018,26(4):1020-1031
  相似文献   
74.
Differential Uptake of Endosulfan in the South American Toad Under Sublethal Exposure     
Gabriela Svartz  Damián Marino  Alicia Ronco  Cristina S. Pérez Coll 《Archives of environmental contamination and toxicology》2015,69(1):104-111
  相似文献   
75.
HIV Infection Among People Who Inject Drugs in the United States: Geographically Explained Variance Across Racial and Ethnic Groups     
Sabriya L. Linton  Hannah L.?F. Cooper  Mary E. Kelley  Conny C. Karnes  Zev Ross  Mary E. Wolfe  Don Des Jarlais  Salaam Semaan  Barbara Tempalski  Elizabeth DiNenno  Teresa Finlayson  Catlainn Sionean  Cyprian Wejnert  Gabriela Paz-Bailey  for the National HIV Behavioral Surveillance Study Group 《American journal of public health》2015,105(12):2457-2465
Objectives. We explored how variance in HIV infection is distributed across multiple geographical scales among people who inject drugs (PWID) in the United States, overall and within racial/ethnic groups.Methods. People who inject drugs (n = 9077) were recruited via respondent-driven sampling from 19 metropolitan statistical areas (MSAs) for the Centers for Disease Control and Prevention’s 2009 National HIV Behavioral Surveillance system. We used multilevel modeling to determine the percentage of variance in HIV infection explained by zip codes, counties, and MSAs where PWID lived, overall and for specific racial/ethnic groups.Results. Collectively, zip codes, counties, and MSAs explained 29% of variance in HIV infection. Within specific racial/ethnic groups, all 3 scales explained variance in HIV infection among non-Hispanic/Latino White PWID (4.3%, 0.2%, and 7.5%, respectively), MSAs explained variance among Hispanic/Latino PWID (10.1%), and counties explained variance among non-Hispanic/Latino Black PWID (6.9%).Conclusions. Exposure to potential determinants of HIV infection at zip codes, counties, and MSAs may vary for different racial/ethnic groups of PWID, and may reveal opportunities to identify and ameliorate intraracial inequities in exposure to determinants of HIV infection at these geographical scales.Since the mid-1990s, there has been an increase in studies evaluating whether features of the social, economic, physical, and political environment (i.e., place characteristics) affect health. This focus on place characteristics is evident in the development of theories conceptualizing place characteristics as health determinants,1–3 in the use of geospatial and systematic social observation methods to measure place characteristics,4–10 in the application of multilevel modeling to assess the potential impacts of place characteristics,11–18 and in the recognition that interventions should not solely encourage individual behavior change but also modify environmental features.3,16,19Literature emerging from this field of research demonstrates that place characteristics operationalized at different geographical scales influence psychosocial processes and individual behaviors that increase vulnerability to several health outcomes. With rare exception,20–24 however, studies of place and health typically assess the potential influence of place characteristics at a single geographical scale and do not simultaneously evaluate characteristics of other geographical scales. For example, several studies, including our own,25,26 sample participants from a single metropolitan statistical area (MSA) to assess the relationships of census tract characteristics to health, without sampling participants from multiple MSAs to simultaneously assess the relationships of tract-, county-, and MSA-level characteristics to health.25–32 The decision to focus on characteristics of a single geographical scale may arise because of data availability, cost constraints, or feasibility.Studies of place and health that focus on a single geographical scale, however, may misspecify relationships and hinder the exploration of causal pathways in 2 ways. First, studies that focus on features measured at a single geographical scale may overlook potential health determinants that are operationalized at other geographical scales. For instance, research assessing the relationships of features of neighborhoods (e.g., economic deprivation, racial/ethnic composition, policing practices, and “crackdowns”) cannot determine the influence of policies, laws, and governmental expenditures that are operationalized at county, MSA, and state levels, and shape neighborhood environments. Second, studies of features of a single geographical scale cannot determine whether relationships between characteristics operating at one geographical scale are confounded, mediated, or modified by characteristics of other geographic scales.3,16,33 The possibility that at least 1 of these mechanisms can occur has been demonstrated in research conducted by Warner and Gomez, which suggests that, among Black women diagnosed with breast cancer, residing in census blocks with high concentrations of Black residents is more protective against mortality in more racially segregated metropolitan areas than less racially segregated metropolitan areas.34In addition, research assessing the association of place-based factors with health outcomes rarely highlights the extent to which variance in health outcomes is explained by place and place-based factors. Determining whether health outcomes vary geographically can generate hypotheses about inequities in exposure to potential place-based determinants of health, and thereby inform how interventions and social policies are developed and spatially concentrated.35The present study illustrates the generative possibilities of extending research beyond a single geographical scale by achieving 2 primary aims. The study’s first aim is to determine the share of total variance in HIV infection that is apportioned to zip codes, counties, and MSAs among people who inject drugs (PWID). In the United States, PWID account for 22% of people living with HIV,36 and a growing body of literature demonstrates that features of neighborhoods such as census-tract racial composition and block-level social or physical disorder are associated with HIV-related outcomes among PWID,37,38 as are features of MSAs, including drug-related law enforcement, income inequality, residential segregation, and health service access.39–41 Revealing the geographical scale to which variance in HIV infection is apportioned among PWID can stimulate hypotheses about inequities in exposure to place-based determinants of HIV and inform the development and tailoring of place-based interventions. For example, finding high MSA-level variance in HIV infection may support analyses of whether MSA-level variations in health care service access predict variance in HIV serostatus and, if they do, support interventions to increase health care access in low-access MSAs. In contrast, if little to no variance in HIV infection among PWID is apportioned to MSAs, PWID may encounter a relatively uniform exposure to health care service access.Previous studies have found that variance in some health outcomes vary across racial/ethnic groups.42,43 The second aim of this study therefore tests the hypothesis that variance in HIV infection will differ within each of 3 racial/ethnic groups of PWID: non-Hispanic/Latino Whites, non-Hispanic/Latino Blacks, and Hispanics/Latinos.  相似文献   
76.
Analysis of pediatric adverse reactions to transfusions          下载免费PDF全文
Sarah Vossoughi  Gabriela Perez  Barbee I. Whitaker  Mark K. Fung  Brie Stotler 《Transfusion》2018,58(1):60-69
  相似文献   
77.
Kinematic Analysis of a Drinking Task in Chronic Hemiparetic Patients Using Features Analysis and Statistical Parametric Mapping     
Gabriela Lopes Santos  Thiago Luiz Russo  Angela Nieuwenhuys  Davide Monari  Kaat Desloovere 《Archives of physical medicine and rehabilitation》2018,99(3):501-511.e4

Objective

To compare sitting posture and movement strategies between chronic hemiparetic and healthy subjects while performing a drinking task, using statistical parametric mapping (SPM) and feature analysis.

Design

Cross-sectional study.

Setting

A university physical therapy department.

Participants

Participants (N=26) consisted of chronic hemiparetic (n=13) and healthy individuals (n=13) matched for sex and age.

Interventions

Not applicable.

Main Outcome Measures

The drinking task was divided into phases: reaching, transporting the glass to mouth, transporting the glass to table, and returning to initial position. An SPM 2-sample t test was used to compare the entire kinematic waveforms of different joint angles (trunk, scapulothoracic, humerothoracic, elbow). Joint angles at the beginning and end of the motion, movement time, peak velocity timing, trajectory deviation, normalized integrated jerk, and range of motion were extracted from the motion data. Group differences for these parameters were analyzed using independent t tests.

Results

At the static posture and beginning of the reaching phase, patients showed a shoulder position more deviated from the midline and externally rotated with increased scapula protraction, medial rotation, anterior tilting, trunk anterior flexion and inclination to the paretic side. Altered spatiotemporal variables throughout the task were found in all phases, except for the returning phase. Patients returned to a similar posture as the task onset, except for the scapula, which was normalized after the reaching phase.

Conclusions

Chronic hemiparetic subjects showed more deviations in the proximal joints during seated posture and reaching. However, the scapular movement drew nearer to the healthy individuals' patterns after the first phase, showing an interesting point to consider in rehabilitation programs.  相似文献   
78.
Intrarater and Inter-rater Reliability of Maximal Voluntary Neck Muscle Strength Assessment Using a Handheld Dynamometer in Women With Headache and Healthy Women     
Ana Paula de Oliveira Carnevalli  Débora Bevilaqua-Grossi  Ana Izabela Sobral Oliveira  Gabriela Ferreira Carvalho  César Fernández-De-Las-Peñas  Lidiane Lima Florencio 《Journal of manipulative and physiological therapeutics》2018,41(7):621-627

Objective

This study aimed to determine the inter-rater and intrarater reliability, agreement, and minimal detectable change (MDC) of the neck muscle strength test using a handheld dynamometer in healthy women and women with headaches.

Methods

Neck muscle strength in maximal voluntary contraction was measured using the Lafayette Manual Muscle Testing attached to a nonelastic belt in 25 women with migraines and in 25 healthy women. Three repetitions of flexion, extension, and lateral flexion were performed. The tests were performed by 2 examiners on the same day, with a 10-minute interval, and by 1 examiner, with a 1-week interval. The reliability was verified by the intraclass correlation coefficient, the agreement determined by standard error measurement, and the MDC calculated.

Results

The protocol exhibited moderate to excellent intrarater and inter-rater reliabilities in both groups (intraclass correlation coefficientrange, 0.53-0.90). The standard error measurement ranged from 0.43 to 1.81 and the MDC from 1.49 up to 4.61.

Conclusion

Quantification of neck muscle strength using the handheld dynamometer with an attached nonelastic belt exhibited moderate to excellent intra- and inter-rater reliability in women with and without migraines. Moreover, the standard error measurement and MDC were proven to be useful in the interpretation of data and in guiding clinical decisions.  相似文献   
79.
Involvement of nitric oxide in improving stress-induced behavioural alteration by glatiramer acetate treatment in female BALB/c mice     
Cecilia Gabriela Pascuan  Elias Hugo Simon  Ana María Genaro  María Laura Palumbo 《Psychopharmacology》2015,232(9):1595-1605
  相似文献   
80.
Lithium increases leukocyte mitochondrial complex I activity in bipolar disorder during depressive episodes     
Rafael T. de Sousa  Emilio L. Streck  Marcus V. Zanetti  Gabriela K. Ferreira  Breno S. Diniz  Andre R. Brunoni  Geraldo F. Busatto  Wagner F. Gattaz  Rodrigo Machado-Vieira 《Psychopharmacology》2015,232(1):245-250
  相似文献   
[首页] « 上一页 [3] [4] [5] [6] [7] 8 [9] [10] [11] [12] [13] 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