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   5116篇
  免费   430篇
  国内免费   11篇
耳鼻咽喉   31篇
儿科学   179篇
妇产科学   126篇
基础医学   689篇
口腔科学   94篇
临床医学   456篇
内科学   1303篇
皮肤病学   113篇
神经病学   464篇
特种医学   233篇
外科学   725篇
综合类   56篇
预防医学   379篇
眼科学   116篇
药学   298篇
中国医学   14篇
肿瘤学   281篇
  2023年   35篇
  2022年   68篇
  2021年   146篇
  2020年   109篇
  2019年   152篇
  2018年   170篇
  2017年   149篇
  2016年   144篇
  2015年   146篇
  2014年   178篇
  2013年   224篇
  2012年   312篇
  2011年   276篇
  2010年   161篇
  2009年   167篇
  2008年   202篇
  2007年   245篇
  2006年   226篇
  2005年   242篇
  2004年   171篇
  2003年   164篇
  2002年   119篇
  2001年   124篇
  2000年   112篇
  1999年   113篇
  1998年   63篇
  1997年   50篇
  1996年   50篇
  1995年   40篇
  1994年   37篇
  1993年   40篇
  1992年   120篇
  1991年   109篇
  1990年   105篇
  1989年   100篇
  1988年   84篇
  1987年   66篇
  1986年   69篇
  1985年   52篇
  1984年   38篇
  1983年   40篇
  1982年   26篇
  1980年   17篇
  1979年   30篇
  1978年   28篇
  1977年   19篇
  1974年   19篇
  1972年   16篇
  1969年   18篇
  1966年   15篇
排序方式: 共有5557条查询结果,搜索用时 15 毫秒
111.
Background: Macrophages account for 5% to 30% of the inflammatory infiltrate in periodontitis and are activated by the classic and alternative pathways. These pathways are identified by indirect markers, among which interferon (IFN)‐γ and interleukin‐6 (IL)‐6 of the classic pathway and IL‐4 of the alternative pathway have been studied widely. Recently, factor XIII‐A (FXIII‐A) was reported to be a good marker of alternative pathway activation. The aim of this study is to determine the macrophage activation pathways involved in chronic periodontitis (CP) by the detection of the indirect markers IFN‐γ, IL‐6, FXIII‐A, and IL‐4. Methods: Biopsies were taken from patients with CP (n = 10) and healthy individuals (n = 10) for analysis of IFN‐γ, IL‐6, IL‐4, and FXIII‐A by Western blot (WB), immunohistochemistry (IHC), and enzyme‐linked immunosorbent assay (ELISA). The same biopsies of healthy and diseased gingival tissue were used, and the expressions of these markers were compared between healthy individuals and those with CP. Results: The presence of macrophages was detected by CD68+ immunohistochemistry and their IFN‐γ, IL‐6, IL‐4, and FXIII‐A markers by WB, IHC, and ELISA in all samples of healthy and diseased tissue. IL‐6, IL‐4, and FXIII‐A were significantly higher in patients with CP, whereas FXIII‐A was higher in healthy individuals. Conclusion: The presence of IFN‐γ, IL‐6, IL‐4, and FXIII‐A in healthy individuals and in patients with CP suggests that macrophages may be activated by both classic and alternative pathways in health and in periodontal disease.  相似文献   
112.
113.
Objective. To test the hypothesis that local proliferation contributes significantly to the hyperplasia of rheumatoid synovium. Methods. Immunohistologic and chemical staining was used to identify 3 markers of cell proliferation: proliferating cell nuclear antigen, c-myc proto-oncogene, and nucleolar organizer regions. Synovium from 21 patients with rheumatoid arthritis, 34 with degenerative joint disease, and 7 with joint trauma was examined. Results. All 3 markers indicated substantial, active proliferation of synovial lining cells in synovium with hyperplasia. Proliferating cells showed type I procollagen immunoreactivity but were negative for CD68, a monocyte/macrophage marker. Proliferation was greater in rheumatoid arthritis than in the other conditions evaluated. Conclusion. In situ proliferation of fibroblast-like synoviocytes in the synovium lining contributes considerably to the increase in cell numbers in rheumatoid synovium.  相似文献   
114.
115.
116.
Objective: To evaluate the effects of an internet-platform exergame cycling programme on cardiovascular fitness of youth with cerebral palsy (CP). Methods: In this pilot prospective case series, eight youth with bilateral spastic CP, Gross Motor Functional Classification System (GMFCS) level III, completed a six-week exergame programme. Outcomes were obtained at baseline and post-intervention. The primary outcome measure was the GMFCS III-specific shuttle run test (SRT-III). Secondary outcomes included health-related quality of life (HQL) as measured by the KIDSCREEN-52 questionnaire, six-minute walk test, Wingate arm cranking test and anthropomorphic measurements. Results: There were significant improvements in the SRT-III (t?=??2.5, p?=?0.04, d?=?0.88) post-intervention. There were no significant changes in secondary outcomes. Conclusion: An exergame cycling programme may lead to improvement in cardiovascular fitness in youth with CP. This study was limited by small sample size and lack of a comparison group. Future research is warranted.  相似文献   
117.
Scalable, solvent-free synthesis of 3,5-isoxazoles under ball-milling conditions has been developed. The proposed methodology allows the synthesis of 3,5-isoxazoles in moderate to excellent yields from terminal alkynes and hydroxyimidoyl chlorides, using a recyclable Cu/Al2O3 nanocomposite catalyst. Furthermore, the proposed conditions are reproducible to a 1.0-gram scale without further milling time variations.

A practical and scalable mechanochemical 1,3-dipolar cycloaddition between hydroxyimidoyl chlorides and terminal alkynes catalyzed by Cu/Al2O3 allows a quick access to 3,5-isoxazole derivatives.

The addition of oxygen or nitrogen-containing heterocycles in drug candidates has become a common feature of the recently approved drugs by the FDA.1,2 In particular, isoxazoles are common molecular scaffolds employed in medicinal chemistry due to the non-covalent interactions such as hydrogen bonding (through the N) and π–π stacking (by the unsaturated 5-membered ring).3–6 Within the isoxazole family, 3,5-isoxazoles (1) are regularly utilized as pharmacophores in medicinal chemistry.2,5,6 Selected examples including muscimol (GABAa agonist), isocarboxazib (antidepressant), isoxicam (anti-inflammatory), berzosertib (ATR kinase inhibitor), and sulfamethoxazole (antibiotic) are highlighted in Fig. 1.7–10Open in a separate windowFig. 1Examples of isoxazoles with pharmacological activity.Various methodologies to synthesize 3,5-isoxazoles have been developed over the years.7,9,11–13 Specifically, 1,3-dipolar cycloaddition between terminal alkynes (2) and nitrile oxides (4) formed in situ by deprotonation of hydroxyimidoyl chlorides (3) is a standard route to access 3,5-isoxazoles (1) (Fig. 2).7,9,14 Recent reports have sought to mitigate the environmental impact of this reaction by performing 1,3-dipolar cycloaddition under solvent-free conditions, using green solvents such as water or ionic liquids, under metal-free conditions, or using mild oxidants.14–28 However, these methodologies have a low atom economy, have a higher hazardous waste production, and are less energy efficient. Therefore, developing a greener methodology that enables rapid and efficient access to these scaffolds is highly desirable.Open in a separate windowFig. 21,3-Dipolar cycloaddition of terminal alkynes and nitrile oxides.Mechanochemistry has been recognized as an environmentally friendly technique as reactions can be performed under solvent-free conditions. Additionally, in some instances, work-up and purification are simplified or absent from procedures, and the process consumes less energy than other solution-based techniques.29–33 The use of mechanochemical techniques to synthesize isoxazoles is limited. Sherin et al. reported a synthesis of 3,5-isoxazoles (7) by grinding in a mortar and pestle curcumin derivatives (5), hydroxylamine (6), and sub-stoichiometric amounts of acetic acid to form the 3,5-isoxazole (7) in short times and excellent yields (Fig. 3a).34 Likewise, Xu et al. studied the synthesis of trisubstituted isoxazoles (10) via 1,3-dipolar cycloaddition of N-hydroxybenzimidoyl chlorides (8) and N-substituted β-enamino carbonyl (9) compounds by ball-milling (Fig. 3b) in high yields, short reaction times, in the absence of catalyst and liquid additives.35 To our knowledge, mechanochemical synthesis of 3,5-isoxazoles (1) from terminal alkynes (2) and hydroxyimidoyl chloride (3) has not been reported (Fig. 3c). The proposed methodology employs a planetary ball-milling technique that provides a route to access in large scale, short reaction times, and high atom economy the corresponding 3,5-isoxazoles (1). Additionally, it utilizes synthetically accessible or commercially available motifs such as terminal alkynes (2) and hydroxyimidoyl chlorides (3) that are recurrent or easily installed in many substrates. Herein, we report a mechanochemical 1,3-dipolar cycloaddition using the planetary ball-mill to synthesize a wide range 3,5-isoxazoles from a broad library of alkynes and (E,Z)-N-hydroxy-4-nitrobenzimidoyl chloride (3a), ethyl (E,Z)-2-chloro-2-(hydroxyimino)acetate (3b), hydroxycarbonimidic dibromide (3c), or (E,Z)-N-hydroxy-4-methoxybenzimidoyl chloride (3d) in moderate to excellent yields, in short reaction time, and with less waste production than in solution based reactions (Fig. 3c).Open in a separate windowFig. 3Previously reported synthesis of isoxazoles.We began our investigation by performing an optimisation of the 1,3-dipolar cycloaddition reaction between alkyne 2a and hydroxyimidoyl chlorides 3a by milling the selected substrates in a stainless-steel (SS) jar in the planetary ball-mill to obtain 3,5-isoxazole 1a (36,37Optimization of reaction conditionsa
EntryChanges from optimized conditionsYieldb (%) of 1a
1None72
2Milling for 10 min, 7 SS balls59
3Milling for 15 min 7 SS balls64
4Milling for 30 min58
5Milling for 40 min60
6Using 1.0 equiv. of 3a65
7Using 2.0 equiv. of 3a57
8Using K2CO371
9Using Cs2CO371
10Using CaCO344
11Using Ag2CO318
12Using NEt3N.R.
Open in a separate windowaReaction Conditions: 0.166 mmol of 2a, 0.250 mmol of 3a, 0.332 mmol of Na2CO3, SS beaker (50 mL capacity), 8 × SS milling balls (10 mm diameter), 20 min milling, 60 Hz.b 1H-NMR yields were measured using 1,3,5-trimethoxybenzene as an internal standard.Having optimized the milling time, we next attempted to improve the yield by varying the equivalents of hydroxyimidoyl chlorides 3a since reaction stoichiometry has been shown to impact the product formed during mechanochemical reactions.38,39 3,5-isoxazole 1a was obtained in lower yields when using equimolar amounts alkyne 2a to hydroxyimidoyl chlorides 3a (entry 6, 40–45 Likewise, increasing the equivalents of 18a from 1.0 to 2.0 equivalents lowered the yield of the reaction (entry 7, 46,47 Using triethylamine (NEt3) proved impractical as the addition of NEt3 to hydroxyimidoyl chlorides was highly exothermic in the absence of solvent (entry 12, 35,48 Therefore, a milling time optimization for other alkyne and hydroxyimidoyl chloride combinations revealed that the most optimal milling time was determined to be between 10 and 30 minutes (see ESI for milling time optimizations).As shown in Fig. 4, stannanyl isoxazole 1a and 1b, silyl isoxazole 1c, and phenyl isoxazole 1d were synthesized with satisfactory yields under the proposed conditions. To explain these results, we suggest an electronic argument. The electron-withdrawing character of the metal substituents, stannyl or silyl of alkyne 2a and 2b, respectively, accelerates the reaction by deactivating the alkyne moiety.49–51 It is observed that alkyne 2a bearing the alkylstannane substituent has a more pronounced effect than the alkyne with the silyl substituent (2b). Therefore, alkyne 2a was the most reactive as it reacted with hydroxyimidoyl chlorides 3a and 3b to synthesize 3,5-isoxazole 1a and 1b respectively, in short times and excellent yields (Fig. 4). On the other hand, ethynyltrimethylsilane (2b) was less reactive as it could only react with a more labile hydroxyimidoyl chlorides 3b to form 3,5-isoxazole 1c (Fig. 4). Comparably, we suggest that the phenyl substituent of alkyne 1c increases the polarizability of the molecule, resulting in deactivating the alkyne moiety. As a result, phenylacetylene (1c) reacted in excellent yields with hydroxyimidoyl chlorides 1b.42 In addition, we observed that the electronic nature of the hydroxyimidoyl chloride substituent affects the reactivity of the nitrile oxide dipole. Hydroxyimidoyl chlorides 3a containing an aromatic substituent with strong electron-withdrawing groups decreased the reactivity of the nitrile oxide.52,53 Consequently, the nitrile oxide synthesized in situ from hydroxyimidoyl chlorides 3a could only react with tributyl(ethynyl)stannane (2a). On the other hand, hydroxyimidoyl chlorides 3b was the most reactive due to the bearing of a weaker electron-withdrawing group such as the ester functional group.43,52 Unfortunately, other alkynes containing substituents such as esters, pyridines, or substituted arenes were not tolerated under these conditions. Previous reports demonstrated the effect of copper catalyst or copper additives to accelerate the reaction and obtain the 3,5-isoxazoles in a regioselective manner.15,53–59 Therefore, we aimed to investigate the effect of copper additives or catalysts on this reaction.Open in a separate windowFig. 4Catalyst-free mechanochemical synthesis of 3,5-isoxazoles.Although the mechanochemical synthesis of 3,5-isoxazoles using copper(ii) catalyst is unprecedented, 1,2,3-triazoles have been synthesized in this way with copper(ii) salts and copper(ii) ions in alumina nanocomposites (Cu/Al2O3).60,61 We investigated the effect of Cu/Al2O3 (see ESI for XPS spectrum) and copper salts using methyl propiolate (2d) and (E,Z)-2-chloro-2-(hydroxyimino)acetate (3b) as model substrates (ii) in the synthesis of 3,5-isoxazolesac
EntryCu(ii)EquivalentsTime (min)Yieldb (%) 19e
1Cu/Al2O30.14 of Cu(ii)1073
2076
3079
4064
5056
2Cu(NO3)2·2.5H2O0.13078
3Cu(NO3)2·2.5H2O1.03084
4Cu(OAc)2·H2O1.03088
5Cu(OTf)21.03076
6CuCl2·H2O1.03076
7Cu2CO3(OH)22.03036
Open in a separate windowaReaction conditions: 0.220 mmol of 2d, 0.330 mmol of 3b, 0.220 mmol of Na2CO3, 0.440 mmol (14 mol%) of Cu/Al2O3, SS beaker (50 mL capacity), 8 × SS milling balls (10 mm diameter), 60 Hz.b 1H NMR yields were measured using 1,3,5-trimethoxybenzene as an internal standard.cSee ESI for solid-state characterization by FT-IR and MALDI-TOF-MS of reaction crude 1e.We observed a significant increase in yield and regioselective control when using sub-stoichiometric amounts of copper (0.14 equivalents or 14 mol%) of Cu/Al2O3 or 10 mol% of Cu(NO3)2·2.5H2O while milling the reagents for 30 minutes (entries 1 and 2, ii), it was observed that Cu(OAc)2·H2O performs similarly to Cu(NO3)2·2.5H2O (entry 4, 62–64We decided to continue our investigations using Cu/Al2O3 as the catalyst can be filtered and washed with solvent, thereby facilitating catalyst recovery and recycling (Fig. 5).60Open in a separate windowFig. 5(a) Filtration of the Cu/Al2O3 catalyst after the first run. (b) Colour change of the Cu/Al2O3 catalyst after recycling. From left to right. (left) Fresh catalyst: blue. (middle) First recycle: green. (right) Second recycle: brown.The Cu/Al2O3 catalyst effect was not exclusively beneficial for the cycloaddition with methyl propiolate (2d) (3,5-isoxazole 1e, Fig. 6). This system improves the reactivity of hydroxyimidoyl chlorides 3a, 3b, 3c, and 3d and other alkynes inaccessible under copper-free conditions, thus allowing access to a broader library of 3,5-isoxazoles (Fig. 6). Moreover, the presence of Cu/Al2O3 nanocomposite as part of the reaction conditions is not impaired by the presence of labile substituents such as silanes (1c, f–h), alkyl halides (1i–j), and boronic esters (1n) (Fig. 6). However, the presence of alkyl stannane substituents in the dipolarophile (2a) was not tolerated with Cu/Al2O3 catalyst, and no product was observed. Furthermore, Cu/Al2O3 enhances the reactivity of dipolarophiles bearing arenes with electron-donating substituents (EDG) (1o–q) and electron-withdrawing groups (EWG) (1n, 1r-2) when coupled with hydroxyimidoyl chlorides 3a and 3b. Additionally, pyridine substituents were more reactive towards the more reactive hydroxyimidoyl chlorides (3b) (3,5-isoxazole 1s, Fig. 6). Ethynyltrimethylsilane (1c) reacted efficiently with hydroxyimidoyl chlorides bearing EWG (3a, 3b, and 3c) to form the respective isoxazoles 1c,1f, and 1h, where silyl isoxazole 1c is obtained in higher yields compared to copper-free conditions (1c, Fig. 4). Hydroxyimidoil chloride bearing EDG (3d); resulted incompatible with terminal alkyne 2b and silyl isoxazole 1g was obtained in lower yields than with EWG in the hydroxyimidoyl chloride. However, terminal alkynes having an aliphatic substituent (2e and 2f) showed greater reactivity towards hydroxyimidoil chloride (3d) bearing EDG; consequently, aliphatic isoxazole 1j was obtained in higher yields than 1i. Then, we evaluated the impact of our conditions in the synthesis of 3,5-isoxazole 1f on a 1.0-gram scale (10.18 mmol). We were pleased to observe that the optimized Cu/Al2O3 conditions can be translated with excellent reproducibility from a 100 mg scale to a 1.0-gram scale without extending the milling time of the reagents (Fig. 6).Open in a separate windowFig. 6Mechanochemical synthesis of 3,5-isoxazoles reaction scope. aAll shown yields are isolated yields. bReaction performed in 1.0-gram scale.The practicality of the proposed methodology allows the recovery of Cu/Al2O3 nanocomposite catalyst directly after the milling of the reagents. In addition, the catalyst recovery allowed investigating the reusability of the recovered catalyst. The Cu/Al2O3 was reused on four occasions, and it was observed that 3,5-isoxazole 1f was obtained successfully with only a minimal drop in yield with each subsequent use for the first two recycling cycles (Fig. 7). The decrease in yield is explained by the decrease in the concentration of active Cu species in the Cu/Al2O3 nanocomposite (see ESI). ICP-MS analysis demonstrates that the Cu concentration of the first recycling represents a decrease of 1.24-fold (with respect to the fresh catalyst); thus, similar yields are obtained compared to the fresh catalyst (Fig. 7). However, the decrease in Cu concentration becomes more substantial for the second and third reuse with a decrease of 2.42 and 6.48-fold, respectively. Therefore, a considerable decrease in the yield of isoxazole 1f is observed. Furthermore, a change in the oxidation state and the bonding of the supported Cu(ii) ions. X-ray Photoelectron Spectroscopy (XPS) analysis of the first and second recycled catalyst reveals that the characteristic satellite signals of Cu(ii) found at about 942.8 eV are weak while the satellite signal at 963.2 eV is absent. Additionally, the 2p3/2 signal at about 933–934 eV is wider than in the fresh sample (see ESI for XPS spectra of the fresh, Fig. S3 for first recycling and Fig. S4 for second recycling). These observations suggest that the supported Cu(ii) is reduced to Cu(0) and CuO is formed with each subsequent recycling.65–67Open in a separate windowFig. 7Cu/Al2O3 efficiency study in the synthesis of 3,5-isoxazole 1f.Lastly, we evaluated the sustainability of the proposed mechanochemical 1,3-dipolar cycloaddition conditions by comparing E-factor for the synthesis of 3,5 isoxazoles 1d and 1f to previously reported solution-based conditions (Fig. 8).54,68 Using E-factor, the values calculated for the planetary ball milling conditions (pathway a and c, Fig. 8) demonstrate the sustainability of this methodology compared to solution-based reactions (pathway b and d) (see ESI for calculations). With our conditions, the absence of organic solvent is the most significant factor contributing to lowering the E-factor.69 Time differences were also another factor of comparison with previously reported solution-based conditions. Our mechanochemical conditions did not surpass 60 minutes, contrary to the reported solution-based conditions that require at least two hours to synthesize the desired 3,5-isoxazoles. Furthermore, our conditions did not show any sensitivity to oxygen or moisture present in the air as all reactions were performed in an open atmosphere.Open in a separate windowFig. 8Comparative green metrics of the proposed methodology to previously reported solution-based methodologies.  相似文献   
118.
119.
120.
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