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   212篇
  免费   21篇
  国内免费   4篇
儿科学   2篇
妇产科学   1篇
基础医学   6篇
临床医学   15篇
内科学   13篇
皮肤病学   8篇
神经病学   2篇
特种医学   1篇
综合类   56篇
预防医学   16篇
药学   18篇
中国医学   97篇
肿瘤学   2篇
  2024年   1篇
  2023年   4篇
  2022年   2篇
  2021年   8篇
  2020年   3篇
  2019年   7篇
  2018年   8篇
  2017年   10篇
  2016年   5篇
  2015年   11篇
  2014年   21篇
  2013年   18篇
  2012年   34篇
  2011年   27篇
  2010年   16篇
  2009年   13篇
  2008年   8篇
  2007年   2篇
  2006年   8篇
  2005年   5篇
  2004年   3篇
  2003年   2篇
  2002年   6篇
  2000年   1篇
  1999年   3篇
  1996年   1篇
  1995年   1篇
  1994年   1篇
  1993年   2篇
  1991年   2篇
  1990年   2篇
  1987年   1篇
  1969年   1篇
排序方式: 共有237条查询结果,搜索用时 62 毫秒
81.
Asian summer monsoon (ASM) variability and its long-term ecological and societal impacts extending back to Neolithic times are poorly understood due to a lack of high-resolution climate proxy data. Here, we present a precisely dated and well-calibrated tree-ring stable isotope chronology from the Tibetan Plateau with 1- to 5-y resolution that reflects high- to low-frequency ASM variability from 4680 BCE to 2011 CE. Superimposed on a persistent drying trend since the mid-Holocene, a rapid decrease in moisture availability between ∼2000 and ∼1500 BCE caused a dry hydroclimatic regime from ∼1675 to ∼1185 BCE, with mean precipitation estimated at 42 ± 4% and 5 ± 2% lower than during the mid-Holocene and the instrumental period, respectively. This second-millennium–BCE megadrought marks the mid-to late Holocene transition, during which regional forests declined and enhanced aeolian activity affected northern Chinese ecosystems. We argue that this abrupt aridification starting ∼2000 BCE contributed to the shift of Neolithic cultures in northern China and likely triggered human migration and societal transformation.

Climatic change and variability can have large and long-lasting consequences for ecosystems and human societies (17). Despite a complex interplay of environmental and nonenvironmental factors, favorable (e.g., warm and wet) climatic conditions have been globally linked to the rise of civilizations, whereas unfavorable conditions have been associated with social instability, human migration, and the more-frequent transformation of civilizations (819). The paucity of high-resolution climate proxy archives that extend prior to the CE, however, prevents a detailed analysis of the linkages between climate variability and potential societal responses for this early period. This is particularly the case for the vast region influenced by the Asian summer monsoon (ASM), for which a good coverage of archaeological data exists that potentially can be used to link climate variability with societal change far back in time.Here, we present an exactly calendar-year dated (by dendrochronological cross-dating) tree-ring–based stable oxygen isotope chronology (the Delingha [DLH] δ18O chronology, Figs. 1 and and2)2) covering ∼6,700 y from 4680 BCE to 2011 CE, which represents the longest existing precisely dated isotope chronology in Asia. In this chronology, we combined stable isotope series from 53 living and relict trees from the DLH region on the northeastern Tibetan Plateau (TP) (Fig. 1), based on a total of 9,526 isotope measurements (SI Appendix, Materials and Methods). The agreement in point-to-point variability between individual tree-ring samples (Fig. 2 A and C) demonstrates the reliability of this composite mean isotope chronology.Open in a separate windowFig. 1.Locations of Holocene paleoclimate records included in this study. The arrows depict the ASM and the Westerlies. The blue dashed line indicates the approximate present-day northern extent of the ASM region based on the observed mean 2 mm/d summer isohyet after ref. 52. The blue triangles represent stalagmite records, the purple dots indicate loess-paleosol profiles, the red asterisks indicate lake sediment records, and the green crosses indicate tree-ring chronologies (including DLH). See SI Appendix, Table S6 for details about each paleoclimate record.Open in a separate windowFig. 2.The DLH tree-ring δ18O chronology. (A) Visualization of all 44 δ18O measurement series. (B) DLH δ18O chronology (navy blue line), third-order polynomial fitting of this chronology (thick black line), and July solar insolation between 30°N and 60°N (red line). The gray shading indicates the 95% CI of the composite δ18O chronology. For better comparison, the y-axis of the δ18O chronology was reversed. (C) Sample depth (with the black line indicating the number of trees in the pooled series) of the DLH δ18O chronology and Rbar (gray line) and EPS (purple line) of the δ18O dataset, calculated over a 250-y window in steps of 1 y. The Rbar time series was smoothed with a 100-y Gaussian-weighted filter. The annual values with EPS ≥ 0.85 accounts for 80.2% during 3250 BCE to 2011 CE, whereas 91.2% of values have EPS ≥ 0.25 and 37.7% are ≥ 0.50 before 3250 BCE.The DLH region is situated at the present-day northwestern fringe of the ASM region (Fig. 1), and our tree-ring record sensitively reflects temporal changes in ASM intensity (SI Appendix, Figs. S16 and S17). Due to the current arid conditions (mean annual precipitation of 170.4 mm, about 85% of which falls in summer [May to September]), tree growth in this region is strongly controlled by precipitation (20). Via soil moisture, precipitation variability controls δ18O ratios in tree-ring cellulose, which is confirmed by the fact that 49% of the variance in annual instrumental precipitation data (prior August to current July; 1956 to 2011) is accounted for by the DLH δ18O chronology. This strong relationship, confirmed by leave-one-out cross-validation (Fig. 3A), allows us to reconstruct regional hydroclimate variability with an unprecedented detail with a 5-y minimum resolution over the past ∼6,700 y (Fig. 3 BD).Open in a separate windowFig. 3.Annual (prior August to current July) tree-ring δ18O precipitation reconstruction ranging from 4680 BCE to 2011 CE. (A) Comparison between reconstructed (red) and instrumental (blue) precipitation (1956 to 2011 CE). The horizontal dashed line indicates the annual mean precipitation (170.4 mm) over the instrumental period (1956 to 2011 CE). (B) Reconstructed precipitation (blue) and 95% CIs (light blue shading). The sky-blue step lines represent regime shifts, and the associated shading indicates 95% CIs for each subperiod (SI Appendix, Materials and Methods). Significant changes in temporal trends (yellow line, with magenta circles indicating trend change point years with P < 0.05: 544 CE, 709 BCE, 1501 BCE, and 2000 BCE; SI Appendix, Materials and Methods). The red horizontal line is the reconstructed mean precipitation of the entire period (4680 BCE to 2011 CE). (C) Extreme dry and wet annual events 4680 BCE to 2011 CE. The events were identified in the precipitation reconstruction as those years in which the precipitation exceeded the 10th and 90th percentiles of the whole period and expressed as percent anomalies from the instrumental period mean. (D) The 100-y running SD of the reconstructed mean annual precipitation. (E) Prehistoric cultural responses to rapid climatic change on the northeastern TP and in northern China (47, 53). The dots of different colors indicate calibrated accelerator mass spectrometry dates of charred grains and bones unearthed from Neolithic and Bronze sites on the northeastern TP, while the pink step line represents temporal variations of number of dated sites every 300 y. The purple step line denotes variations of war frequency over time in east Qinghai Province during the past two millennia (32, 33).Our precipitation reconstruction shows a pronounced multimillennial drying trend (Figs. 3B and and4A).4A). This trend is in agreement with proxy evidence of lower temporal resolution from stalagmite δ18O records from eastern China (2123), pollen-based precipitation reconstructions from eastern China (24), and other moisture-sensitive proxy archives (Figs. 1 and 4 B and C, and SI Appendix, Figs. S12–S15). However, our DLH reconstruction quantifies long- and short-term climatic events at a much higher temporal resolution and with precise dating accuracy, offering a unique benchmark record to synchronize Chinese archaeological evidence and anchor a range of contemporary paleoenvironmental data. It also benefits from a robust calibration between the climate proxy and instrumental climatic data, and an in-depth comparison with model simulations.Open in a separate windowFig. 4.Comparison of the DLH tree-ring δ18O precipitation reconstruction with other paleoclimatic records spanning the Holocene. (A) Anomaly percentage of the DLH precipitation reconstruction calculated over the period 4680 BCE to 1950 CE (this study). (B) Pollen-based annual precipitation anomaly percentage in Gonghai Lake calculated over the common period 4680 BCE to 1950 CE (24). (C) Normalized stalagmite composite δ18O record from eastern China. The y-axis of the composite δ18O record was reversed for better comparison. Each stalagmite δ18O record was first normalized over the common period 4700 BCE to 1300 CE using the equation (abm) / bs, where a is the original value, and bm and bs are the mean and SD of the common period, respectively. See SI Appendix, Table S6 (site no.: 1 to 6) for details about each stalagmite record employed in the calculation. (D) Variation in location of the ITCZ reflected by Cariaco Basin Ti concentrations (26). All horizontal lines represent the long-term average calculated over the common period 4680 BCE to 1950 CE. The long-term precipitation average values are 200 and 511 mm, respectively for panels (A and B). For panels (AD), all series were first interpolated annually by using a piecewise linear interpolation method, and then each series (thin line) was smoothed by a 100-point low-pass filter (heavy line) to highlight the centennial scale variability.A long-term aridification trend since the mid-Holocene is evident, which closely matches a corresponding negative trend in summer solar insolation from 30 to 60°N (Figs. 2B andand3B).3B). Thus, we hypothesize that summer insolation has been a primary driver of long-term aridification at the northern limits of the ASM zone of China since the mid-Holocene. Decreasing summer insolation may have considerably reduced the thermal contrast between the Asian continent and the surrounding oceans, thereby leading to a displacement of the Intertropical Convergence Zone (ITCZ) and a weakening of the ASM circulation resulting in reduced precipitation in the ASM marginal areas.The long-term aridification that characterizes our DHL reconstruction and other proxy evidence (SI Appendix, Fig. S15), accompanied by the cooling trend through the middle to late Holocene, is confirmed by the CCSM3 climate model (SI Appendix, Materials and Methods) that simulates decreasing temperature and precipitation trends in northern China (25). Our precipitation reconstruction is positively correlated with centennial-scale China-wide temperature variability over the most recent two millennia (SI Appendix, Fig. S18), suggesting that future large-scale warming might be associated with even greater moisture supply in this region. Model simulations also suggest that the long-term moisture variations in the marginal monsoon region are closely linked to shifts in the mean position of the ITCZ, as also indicated by titanium concentration trends from the Cariaco Basin in the Caribbean Sea (26) (Fig. 4D).In addition to temporal ASM variability, the mean DLH δ18O value can also reflect changes in spatial ASM extent. We compared the mean δ18O value of our DLH chronology with another Qilian juniper isotope chronology from the Animaqing Mountains located 300 km to the southeast of our study site at a similar elevation. For the recent period (1930 to 2011 CE), δ18O in Animaqing amounts to 30.78 ± 1.33‰ (27), which is significantly lower than at DLH (32.84 ± 1.07‰). However, the mean value in the earliest part of our DLH δ18O chronology (4680 to 3000 BCE; 29.80 ± 1.12‰) is closer to the present-day Animaqing values, indicating that humid present-day climate conditions in the Animaqing Mountains may be used as a modern analog for mid-Holocene climate in the DLH region. Given this, we infer that during the mid-Holocene, the ASM limit extended at least 300 km further northwest compared to its present-day limit.An assumed northward shift of the ASM boundary during the mid-Holocene is supported by additional regional paleoclimatic evidence of lower temporal resolution. A 300- to 400-km northwestward migration of the ASM rain belt during the early and mid-Holocene has been suggested from a lake size record from northeastern China (28) and from plant biomass data in loess sections across the Loess Plateau (29). A climate reconstruction combining vegetation type and sedimentary facies in aeolian deposits (30) further suggests that deserts in northern China retreated by ∼200 km to the northwest during the mid-Holocene (4800 ± 300 BCE).Our high-resolution precipitation reconstruction provides absolute estimates for precipitation differences between the mid-Holocene and present-day conditions. We estimate mean annual precipitation during the mid-Holocene (here, 4680 to 3000 BCE) as 279 ± 10 mm, which exceeds the average levels of the entire reconstruction period (4680 BCE to 2011 CE; 200 ± 9 mm) and of the instrumental period (1956 to 2011 CE; 170.4 mm) by 40 (∼38 to 41% at 95% confidence) and 63% (∼57 to 69% at 95% confidence), respectively (Figs. 3B and and4A4A).Our precipitation reconstruction also reveals centennial-scale variability that differs substantially from a ∼20-y–resolution pollen-based annual precipitation record (24) (Fig. 4 A and B). In comparison with this pollen-based reconstruction, which shows precipitation variations in the range of ±25% of the long-term average, the DLH δ18O reconstruction displays a much larger centennial-scale variability, ranging from −50 to 50%.Using a sequential Student’s t test approach, we identified several major, clearly dateable centennial-scale regime shifts (Fig. 3B and SI Appendix, Fig. S10 and Table S7) in our DLH record (31) (SI Appendix, Materials and Methods). We detected the strongest shifts toward dry conditions around 3350, 2815, 2095, 1675, and 70 BCE and 346 CE (SI Appendix, Table S7). Regime shifts toward wetter conditions were typically less dramatic, and occurred in 2565, 1185 BCE, and 760 CE (SI Appendix, Table S5). The precise dating of these regime shifts allows us to determine the duration and magnitude of past dry epochs.The most severe and long-lasting dry period prior to the CE occurred c. 1675 to 1185 BCE (Fig. 3B and SI Appendix, Table S7), representing a remarkable megadrought (mainly represented on a millennial scale with three obvious centennial droughts superimposed, SI Appendix, Fig. S11) with an estimated mean annual precipitation of 42 ± 4 and 5 ± 2% less than the average over the mid-Holocene (4680 to 3000 BCE) and the instrumental period (1956 to 2011 CE), respectively. Trend-point analysis (SI Appendix, Fig. S10) confirms that this 1675 to 1185 BCE megadrought marks a low in the long-term general drying trend in the DLH reconstruction, which intensified between ∼2000 and ∼1500 BCE (Fig. 3B). This period of rapidly decreasing moisture availability starting ∼2000 BCE and culminating ∼1500 BCE thus arguably marks the transition from the mid- to the late Holocene Asian moisture regime.Another period of long-lasting extremely dry conditions occurred c. 346 to 763 CE (Fig. 3B and SI Appendix, Table S7). This extremely dry period, when war frequency reached a maximum in east Qinghai Province due to conflicts between different local regimes and decreased rapidly afterward (32, 33) (Fig. 3E), was also recorded in other hydroclimatic proxies in China (20) and partly overlaps with the “Late Antique Little Ice Age” (LALIA) (2). The correspondence of social unrest and drought indicates a likely impact of climate deterioration on society at that time. At a hemispheric scale, Zhang et al. (34) argued that climate change may have imposed a spatially wider-ranging effect on human civilization.The LALIA megadrought represents the culmination of the millennial-scale drying trend in the DLH reconstruction, which reversed around ∼544 CE (indicated by trend-point analysis; P < 0.05; SI Appendix, Fig. S10 and Fig. 3B). As a result of this hydroclimatic trend reversal, precipitation and insolation trends started to diverge by the middle of the first millennium CE, when solar insolation continued to decrease, whereas precipitation did not (Figs. 2B and and3B3B).Our mid-Holocene–length hydroclimate reconstruction thus records multiple distinct climate regime shifts. However, it does not support a significant transition in the hydroclimate of our study region around ∼2200 BCE during the so-called “4.2-ka event” (35), nor the notion that this rapid climate deterioration and associated global-scale megadroughts should be regarded as a generalized climatic transition from the mid- to late Holocene (36).At high temporal resolution, our DLH reconstruction shows that moisture conditions alternated between extremely wet and dry periods at interannual, decadal, and multidecadal timescales (Fig. 3B and SI Appendix, Table S8). For example, mean annual precipitation extremes of opposite signs can occur within a few decades (e.g., 309 mm in 1990 BCE compared with 47 mm in 1950 BCE and 313 mm in 1715 BCE compared with 95 mm in 1675 BCE). In the most recent 50 y (1956 to 2011), precipitation has increased in our study region and had previously been found to be the wettest period of the past 3,500 y (20). However, our DHL precipitation reconstruction indicates that this wet recent period is not unprecedented in historical times (Fig. 3B). The discrepancy between the two studies can likely be attributed to the strength of the precipitation signal in the two tree-ring parameters (tree-ring width in ref. 20 versus δ18O in this study), the extension of the DLH δ18O chronology into the wetter mid-Holocene, and concerns about whether the detrended tree-ring width record (20) is able to capture climate variability on millennial timescales (SI Appendix, Fig. S12).Wet extremes occurred with the highest intensity and frequency prior to 2800 BCE (Fig. 3C and SI Appendix, Tables S3 and S8). In line with the long-term aridification trend, the frequency and magnitude of wet extremes in our record decreased over the following two millennia. In contrast, the frequency of dry extremes increased and peaked around 660 CE, with potentially harmful impacts on contemporary human societies.Precipitation variability has changed considerably over time, as shown by a 100-y running SD plot (Fig. 3D). Over the entire record, the mean SD is 42 mm, but extended periods of low SD occurred from 4680 to 3200 BCE, 2500 to 2000 BCE, and 1000 to 1500 CE. The first of these is particularly notable because of the sudden transition toward a period with particularly high variability around 3200 BCE.The humid climate during the mid-Holocene and the subsequent aridification had major impacts on the ecological environment in China. Pollen records from northern China testify to a broad-scale transition from forest to steppe vegetation in the climate-sensitive ASM margin around ∼1600 BCE (37) (SI Appendix, Fig. S19). In the more humid eastern TP, a phase of major deterioration of Picea forests occurred after 1600 BCE. Woody debris in Qinghai Lake sediments verify that spruce (Picea crassifolia Kom.) forests had already developed in the region 7700 to 2200 BCE and subsequently disappeared (38). Combining these results with our ASM reconstruction, we propose that wetter conditions during the mid-Holocene played a major role in establishing a denser regional forest cover. The subsequent abrupt aridification (reaching a very dry regime by ∼1675 BCE) initiated a broad-scale forest decline in northern China, finally resulting in the disappearance of spruce forests in the Qinghai Lake basin. The mid- to late Holocene aridification trend is also reflected by enhanced aeolian activity (39).Our DLH precipitation reconstruction supports assessments of the societal responses to rapid climatic change in China. The wet and climatically rather stable mid-Holocene (Fig. 3 B and D) likely contributed to facilitate the expansion of the Yangshao culture across China (Fig. 3E). The prosperity of the Majiayao (3300 to 2000 BCE) and Qijia cultures (2300 to 1600 BCE) in the Gansu-Qinghai region (4043) may also be associated with contemporary favorable regional climate conditions. In the northern and southern Loess Plateau, two large-scale Neolithic urban centers, Shimao (2300 to 1800 BCE) and Taosi (2300 to 1900 BCE), flourished (44, 45). Both centers were abandoned after 1800 BCE, perhaps partly as a result of the rapid regime shift from a wet to a dry climate in the second-millennium BCE (considering the radiocarbon dating uncertainty of the archaeological material).This second-millennium–BCE megadrought may also have had a major impact on human civilizations in the semiarid and arid regions of northern China, where water availability is a major constraint for human subsistence. A sudden drop in the number of archaeological sites on the northeastern TP occurred between 2000 and 1400 BCE, as shown by calibrated accelerator mass spectrometry radiocarbon dates of charred grains and bones (Fig. 3E). The Qijia culture began to disintegrate around 1600 BCE and evolved into multiple cultures (e.g., Kayue, Xindian, and Nuomuhong) (Fig. 3E). Such dry and cold climate along with increased climate variability (Fig. 3D), coupled with innovations in agriculture, could have contributed to the process and led to a change in a subsistence strategy from millet farming to combined barley and wheat farming in the Gansu-Qinghai region (46). Substituting millet production with barley that is better adapted to the cooler and drier conditions likely limited the risk of crop failure and enabled humans to cultivate at TP altitudes above 3,000 m above sea level (43, 46, 47). After ∼1500 BCE, barley spread southwards into the southeastern TP and replaced millet that could not adapt to cooler and drier conditions of the late Holocene (48). Meanwhile, in the western Loess Plateau, human subsistence went through a major transition from long-established rain-fed agriculture to mobile pastoralism after ∼1600 BCE (42, 49), which is consistent with the c. 1675 to 1190 BCE megadrought recorded in our precipitation reconstruction.The effects of the second-millennium–BCE megadrought become apparent in a comprehensive review of archaeological evidence across China, including 51,074 sites covering most parts of China and spanning the early Neolithic to early Iron Age (c. 8000 to 500 BCE) (50, 51). Herein, a steady increase in the number of archaeological sites can be detected from 5800 to 1750 BCE (50), implying continuous cultural development in large areas of China. The absence of evidence for irrigation-based farming indicates that rain-fed agriculture was sufficient to sustain Neolithic and early Chalcolithic communities (52). The abrupt aridification around 1675 BCE corresponded to a sudden reduction in the number of archaeological sites, as well as a contraction in the areal distribution of sites across all of China (SI Appendix, Fig. S20). The number of archaeological sites around the middle and lower reaches of the Yellow River decreased substantially, marking the almost-complete abandonment of the Guanzhong Basin (51), while the highest number of sites during this period can be found in northeastern China (50, 51). Therefore, it seems that the aridification around 2000 to 1500 BCE could be, at least partly, responsible for a large human migration phase in northern China. At the same time (2000 to 1600 BCE), the earliest documented Chinese kingdoms associated with the Xia dynasty emerged, which were later replaced by the Shang dynasty (∼1600 to 1000 BCE) (53). In view of all the evidence stated above, we propose that the second-millennium–BCE megadrought might have accelerated the disintegration of these historical civilizations.In conclusion, we present a precisely dated benchmark timeseries representing multiscale variability in ASM intensity and extent over the past 6,700 y. We show that solar insolation is responsible for driving most of the multimillennial variation in ASM intensity. We identified two severe and long-lasting dry periods, 1675 to 1185 BCE and 346 to 763CE, that both correspond to periods of regional societal turbulence. We propose that rapidly decreasing moisture availability starting ∼2000 BCE marks the transition from mid- to late Holocene and resulted in unfavorable environmental conditions, ultimately exerting severe pressures on natural forest vegetation, crop production, and societal development in northern China. These cultures collapsed one by one, initiated around ∼2000 BCE by the aridification of the local climate. In this context, some of the extreme drought events recorded by our reconstruction might have accelerated the disintegration of ancient civilizations. The complexity of their social structure, associated with differing adaptation abilities and strategies to resist adverse climatic stress, can explain regional differences in timing of their disintegration.  相似文献   
82.
Following low incidence of respiratory syncytial virus (RSV) infections in 2020 during the COVID-19 pandemic, we noted a resurgence in hospitalised children in spring/summer 2021 following relaxation of public health measures. We compared this outbreak to previous autumn/winter seasons. We found higher weekly case numbers and incidence rates, more cases from urban neighbourhoods with lower socioeconomic status, and similar clinical presentation and severity. Public health implications include the re-evaluation of palivizumab administration and the need for surge capacity planning.  相似文献   
83.
2008年全国研究生暑期学校(护理学)教学效果评价   总被引:1,自引:0,他引:1  
目的:探讨首届全国护理学研究生暑期学校的教学效果。方法:采用自编的“护理学暑校教学评价表”对参加首届护理学暑期学校的128名护理专业研究生和青年教师进行问卷调查。结果:学员对各指标的满意度均在85%以上。结论:首届护理学暑期学校的举办收到了良好的效果,为今后护理学暑期学校的举办以及我国护理研究生教学管理提供了参考。  相似文献   
84.
目的:探讨中药联合针灸治疗小儿夏季腹泻的临床疗效.方法:将解放街道社区卫生服务中心门诊收治的夏季腹泻患儿105例随机分为观察组55例及对照组50例,观察组应用中医的治疗原则联合中药和针灸进行辨证施治;对照组予以西医治疗.两组均在治疗7d后予以疗效评价.结果:观察组治愈50例,有效4例,无效1例,总有效率为98.2%;对照治愈38例,有效6例,无效6例,总有效率为88%.两组总有效率比较,差异具有统计学意义(P<0.05).结论:中药联合针灸治疗小儿夏季腹泻疗效肯定,副作用小,相比西医治疗具有明显优势.  相似文献   
85.
冬病夏治在中医学中独成体系且治法众多,外治法有穴位贴敷的"三伏贴"、择时施针用灸的"三伏针"或"三伏灸",内治法有内服汤药、丸散膏丹和饮食调养的"三伏补",该法实用有效,操作简单方便,易于患者接受,便于临床推广。  相似文献   
86.
郭明升  李薇 《临床荟萃》2014,29(12):1342
目的 探讨冬、夏季节椎-基底动脉系统短暂性脑缺血发作(VBAS-TIA)患者临床特点的差别.方法 225例VBAS-TIA患者按照发病时间分为冬季组(134例)和夏季组(91例),比较组间的性别、年龄和伴发高血压、糖尿病、高脂血症、心脏病的例数;比较组间的血压、左心室射血分数、心输出量、颈部血管血流速度以及颈动脉粥样硬化(AS)、重度狭窄、椎动脉迂曲/纤细的例数.结果 与冬季组比较,夏季组患者的年龄较轻、男性和伴发高血压、高脂血症的例数均较少(P<0.05);伴发糖尿病、心脏病的例数均差异无统计学意义(P>0.05).与冬季组比较,夏季组患者的血压、颈部血管血流速度以及AS、重度狭窄、椎动脉迂曲/纤细的例数均较少(P<0.01);左心室射血分数、心输出量的比较均差异无统计学意义(P>0.05).结论 季节因素可能影响着不同时期VBAS-TIA的临床特点.  相似文献   
87.
目的:分析冬病夏治穴位贴敷联合中医辨证治疗儿童咳嗽变异性哮喘(Cough Variant Asthma,CVA)的临床疗效。方法:选取2014年8月至2016年10月CVA患儿90例,按照随机数表法将所有的入选患儿分为观察和对照组,各45例,对照组给予中医辨证的常规治疗,观察组在中医辨证基础上给予冬病夏治穴位贴敷治疗,观察比较2组的治疗前后血清免疫球蛋白水平情况、临床症状改善情况、疗效和观察指标情况以及患儿家属对治疗的满意度。结果:2组患儿治疗前Lg A、Lg M、Lg G、Lg E指标,差异无统计学意义(P0.05),治疗后,观察组患儿的Lg A、Lg M、Lg G指标均高于对照组,Lg E指标低于对照组,差异有统计学意义(P0.05);治疗后,观察组出现阵发呛咳、痰黏难咳、手足心热、夜卧不安、大便秘结症状的例数均少于对照组,差异有统计学意义(P0.05);观察组总有效率(86.67%)高于对照组(68.89%),差异有统计学意义(P0.05);观察组每年复发次数、复发咳嗽症状持续时间、每年因此病住院人数和每年感冒次数均少于对照组,差异有统计学意义(P0.05);观察组患儿家属对治疗的满意度(91.11%)高于对照组(75.56%),差异有统计学意义(P0.05)。结论 :对CVA儿童进行冬病夏治穴位贴敷联合中医辨证治疗,安全可靠,治疗效果好,患儿家属满意度高。  相似文献   
88.
目的:分析总结在稳定期慢性阻塞性肺疾病(COPD)治疗中,中医三伏灸、穴位注射配合肺功能康复治疗的应用效果。方法:选取2015年6月至2017年6月罗定市中医院收治的稳定期慢性阻塞性肺患者120例作为研究对象,随机分为对照组和观察组,每组60例,其中对照组患者给予对症处理、家庭氧疗以及健康指导等常规治疗,观察组患者在上述治疗的基础上外加中医冬病夏治(用本院自制灸热贴按疗程规范进行三伏灸)、穴位注射药物(黄芪注射液)治疗,比较2组患者治疗前后的免疫功能、肺功能指标以及每年的发病次数和医疗花费情况,分析2组患者治疗后的症状改善以及临床疗效。结果:2组患者治疗前的T淋巴细胞亚群、免疫球蛋白以及肺功能指标变化不显著(P0.05),具有可比性;治疗后2组患者的CD4~+、CD3~+以及CD4~+/CD8~+比值均有不同程度的升高,CD8~+降低,但观察组患者T淋巴细胞亚群变化更为显著(P0.05);治疗后,观察组患者的血清Ig M、Ig G、Ig A显著高于对照组,而COPD发病次数和患者的医疗费用均明显低于对照组,差异有统计学意义(P0.05);2组患者治疗前后的FEV1、FEV1%以及FVC%差异不显著(P0.05)。对照组患者的治疗有效率为53.33%,观察组的治疗有效率为76.78%,观察组治疗效果显著优于对照组(P0.05)。结论:中医三伏灸、穴位注射配合肺功能康复治疗稳定期疗法可以显著增强患者机体免疫力,降低发病次数,减少医疗开支,减轻家庭经济负担,对肺功能指标改善较弱,但可以显著减轻咳、痰、喘、气促等临床症状,增强治疗有效率。  相似文献   
89.
刘红梅  杨帆  王竹  李晓娟 《河南中医》2021,41(2):310-314
穴位注射治疗支气管哮喘,适用于该病临床各期,可有效缩短哮喘控制时间,减少哮喘复发次数,取穴一般以近部取穴为主,同时与远部取穴、辨证取穴相结合。支气管哮喘发作多具有节律性和周期性,所以穴位注射可因时令施治,冬病夏治穴位注射治疗支气管哮喘可在防治过程中起到重要作用。目前,穴位注射所用药物种类很多,西药有卡介菌多糖核酸、曲安奈德、核酪注射液、维生素类等;中药有喘可治、黄芪注射液、复方当归注射液等;也有采用自体血穴位注射的方法,以上治疗均能在不同程度上控制哮喘症状,改善肺功能。  相似文献   
90.
Despite calls to increase the number and diversity of doctorally prepared nurses, recent data indicate a severe shortage of PhD‐prepared nurses, especially those of racial/ethnic minority backgrounds. This is concerning, given that evidence indicates that racial/ethnic minority PhD‐prepared nurses are well‐positioned to address health disparities, by attending to the needs/concerns of medically underrepresented groups. The purpose of this article is to describe and provide a preliminary evaluation of a summer research program for minority nursing students. Online surveys were administered to assess for student satisfaction, knowledge gains, attitudes toward research, and intentions to pursue a PhD among minority undergraduate nursing students (N = 6) participating in the 10‐week program. Favorable trends were observed related to satisfaction, knowledge gains, and attitudes toward research. Fifty percent of the sample intended to pursue a PhD immediately after the program, compared to none before the program, and this result was maintained at 1‐year post‐program. The summer research program appears to be a promising strategy for increasing the number/diversity of PhD‐prepared nurses. More research on the implementation of programs exposing minority nursing students to health disparities research is needed to strengthen evidence that similar programs can serve to increase the pipeline of diverse doctorally prepared nurses.  相似文献   
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号