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   8202篇
  免费   629篇
  国内免费   8篇
  8839篇
  2023年   44篇
  2022年   89篇
  2021年   176篇
  2020年   121篇
  2019年   128篇
  2018年   215篇
  2017年   176篇
  2016年   257篇
  2015年   438篇
  2014年   420篇
  2013年   547篇
  2012年   719篇
  2011年   629篇
  2010年   416篇
  2009年   322篇
  2008年   477篇
  2007年   445篇
  2006年   432篇
  2005年   411篇
  2004年   415篇
  2003年   386篇
  2002年   309篇
  2001年   97篇
  2000年   71篇
  1999年   75篇
  1998年   83篇
  1997年   47篇
  1996年   53篇
  1995年   50篇
  1994年   51篇
  1993年   35篇
  1992年   51篇
  1991年   51篇
  1990年   31篇
  1989年   35篇
  1988年   31篇
  1987年   28篇
  1986年   29篇
  1985年   45篇
  1984年   35篇
  1983年   30篇
  1982年   24篇
  1981年   25篇
  1980年   30篇
  1979年   27篇
  1978年   26篇
  1977年   24篇
  1976年   31篇
  1975年   25篇
  1974年   19篇
排序方式: 共有8839条查询结果,搜索用时 15 毫秒
71.

Background

Tick-borne relapsing fever spirochetes are maintained in endemic foci that involve a diversity of small mammals and argasid ticks in the genus Ornithodoros. Most epidemiological studies of tick-borne relapsing fever in West Africa caused by Borrelia crocidurae have been conducted in Senegal. The risk for humans to acquire relapsing fever in Mali is uncertain, as only a few human cases have been identified. Given the high incidence of malaria in Mali, and the potential to confuse the clinical diagnosis of these two diseases, we initiated studies to determine if there were endemic foci of relapsing fever spirochetes that could pose a risk for human infection.

Methodology/Principal Findings

We investigated 20 villages across southern Mali for the presence of relapsing fever spirochetes. Small mammals were captured, thin blood smears were examined microscopically for spirochetes, and serum samples were tested for antibodies to relapsing fever spirochetes. Ornithodoros sonrai ticks were collected and examined for spirochetal infection. In total, 11.0% of the 663 rodents and 14.3% of the 63 shrews tested were seropositive and 2.2% of the animals had active spirochete infections when captured. In the Bandiagara region, the prevalence of infection was higher with 35% of the animals seropositive and 10% infected. Here also Ornithodoros sonrai were abundant and 17.3% of 278 individual ticks tested were infected with Borrelia crocidurae. Fifteen isolates of B. crocidurae were established and characterized by multi-locus sequence typing.

Conclusions/Significance

The potential for human tick-borne relapsing fever exists in many areas of southern Mali.  相似文献   
72.
We used botulinum neurotoxins (BoNT) to examine whether differences in the secretory activity of noradrenergic and adrenergic chromaffin cells are related to differences in the exocytotic machinery of these two types of bovine adrenal medulla cells. Cleavage of syntaxin and SNAP-25 by BoNT/C1 decreased in a dose-dependent way the release of both noradrenaline and adrenaline, but noradrenaline release was more sensitive to BoNT/C1. Cleavage of SNAP-25 by BoNT/A also had a larger inhibitory effect on noradrenaline release than on adrenaline release. Neither BoNT/C1 nor BoNT/A affected the intracellular Ca2+ responses induced by K+-depolarisation, and the extent of the inhibition of K+-evoked catecholamine release by selective blockers of voltage-gated Ca2+ channels was not affected by BoNT/C1. Therefore, our data do not support the hypothesis of a regulatory effect of syntaxin or SNAP-25 on the activity of Ca2+ channels. The lower sensitivity of adrenaline release to BoNT was not due to a reduced ability of the toxins to enter or to cleave their protein targets in adrenergic cells, since immunoblot analysis showed the cleavage of a larger fraction of syntaxin 1A in adrenergic cells, as compared to the cleavage in noradrenergic cells. The immunoblot analysis also showed larger amounts of syntaxin 1A in noradrenergic chromaffin cells than in adrenergic cells. Thus, in spite of a greater cleavage of syntaxin 1A in adrenergic cells by BoNT/C1, adrenaline release was less sensitive to BoNT/C1, suggesting that the release process in noradrenergic cells might be more dependent on syntaxin 1A and SNAP-25, as compared to adrenergic cells.  相似文献   
73.
Predatory bacteria seek and consume other live bacteria. Although belonging to taxonomically diverse groups, relatively few bacterial predator species are known. Consequently, it is difficult to assess the impact of predation within the bacterial realm. As no genetic signatures distinguishing them from non-predatory bacteria are known, genomic resources cannot be exploited to uncover novel predators. In order to identify genes specific to predatory bacteria, we developed a bioinformatic tool called DiffGene. This tool automatically identifies marker genes that are specific to phenotypic or taxonomic groups, by mapping the complete gene content of all available fully-sequenced genomes for the presence/absence of each gene in each genome. A putative ‘predator region’ of ~60 amino acids in the tryptophan 2,3-dioxygenase (TDO) protein was found to probably be a predator-specific marker. This region is found in all known obligate predator and a few facultative predator genomes, and is absent from most facultative predators and all non-predatory bacteria. We designed PCR primers that uniquely amplify a ~180bp-long sequence within the predators’ TDO gene, and validated them in monocultures as well as in metagenetic analysis of environmental wastewater samples. This marker, in addition to its usage in predator identification and phylogenetics, may finally permit reliable enumeration and cataloguing of predatory bacteria from environmental samples, as well as uncovering novel predators.  相似文献   
74.
75.
76.
Specific activities of eight enzymes involved in glycerol metabolism were determined in crude extracts of three strains ofNeurospora crassa after growth on six different carbon sources. One of the strains was wild type, which grew poorly on glycerol as sole carbon source; the other two were mutant strains which were efficient glycerol utilizers. A possible basis for this greater effeciency of glycerol utilization was catabolite repression of glyceraldehyde kinase by glycerol in wild type, and two-fold higher glycerate kinase activity in the mutant strains after growth on glycerol, thus apparently allowing two routes for glyceraldehyde to enter the glycolytic pathway in the mutant strains but only one in wild type. The preferential entry of glyceraldehyde to the glycolytic pathway through glycerate was suggested by the lack of glyceraldehyde kinase in all three strains after growth on one or more of the carbon sources and the generally higher levels of aldehyde dehydrogenase and of glycerate kinase than of glyceraldehyde kinase.  相似文献   
77.
78.
Amrinone is a bipyridine compound with characteristic effects on the force-velocity relationship of fast skeletal muscle, including a reduction in the maximum shortening velocity and increased maximum isometric force. Here we performed experiments to elucidate the molecular mechanisms for these effects, with the additional aim to gain insight into the molecular mechanisms underlying the force-velocity relationship. In vitro motility assays established that amrinone reduces the sliding velocity of heavy meromyosin-propelled actin filaments by 30% at different ionic strengths of the assay solution. Stopped-flow studies of myofibrils, heavy meromyosin and myosin subfragment 1, showed that the effects on sliding speed were not because of a reduced rate of ATP-induced actomyosin dissociation because the rate of this process was increased by amrinone. Moreover, optical tweezers studies could not detect any amrinone-induced changes in the working stroke length. In contrast, the ADP affinity of acto-heavy meromyosin was increased about 2-fold by 1 mm amrinone. Similar effects were not observed for acto-subfragment 1. Together with the other findings, this suggests that the amrinone-induced reduction in sliding velocity is attributed to inhibition of a strain-dependent ADP release step. Modeling results show that such an effect may account for the amrinone-induced changes of the force-velocity relationship. The data emphasize the importance of the rate of a strain-dependent ADP release step in influencing the maximum sliding velocity in fast skeletal muscle. The data also lead us to discuss the possible importance of cooperative interactions between the two myosin heads in muscle contraction.Muscle contraction, as well as several other aspects of cell motility, results from cyclic interactions between myosin II motors and actin filaments. These force-generating interactions are driven by the hydrolysis of ATP at the myosin active site as outlined in Scheme 1 (13). In the absence of actin, the Pi and ADP release steps (k4 and k5) are rate-limiting for the entire cycle at high (>12 °C) and low temperatures, respectively (46). In the presence of actin, the rate of Pi release increases significantly, and the overall cycle is accelerated more than 2 orders of magnitude. The sliding velocity of myosin-propelled motors is generally believed to be rate-limited by actomyosin dissociation (rate constant k5, k6, or k2 in Scheme 1) (7). Alternatively, some studies (8, 9) have suggested that the sliding velocity is determined by the fraction of myosin heads in the weak-binding states, AM4 ATP and AM ADP Pi. However, it is worth emphasizing that KT is very low under physiological conditions (1, 3) with low population of these states. For the same reason, the rate of dissociation of the AM complex is governed by K1 and k2.Open in a separate windowSCHEME 1.Simplified kinetics scheme for MgATP turnover by myosin (lower row) and actomyosin (upper row). Inorganic phosphate is denoted by Pi; MgATP is denoted by ATP, and MgADP is denoted by ADP; myosin is denoted by M. The states AM*ADP and AM ADP correspond to myosin heads with their nucleotide binding pocket in a partially closed and open conformation, respectively (7, 52). Rate constants are indicated by lowercase letters (rightward transitions, k2k5 and k2k5, or leftward transitions, k−2k−5 and k−2k−5) and equilibrium constants by uppercase letters (K1, K1, KT, K3, K3, K6, k6, and KDP). The equilibrium constants are association constants except for simple bimolecular reactions where they are defined as ki/ki.For the study of contractile mechanisms in both muscle and other types of cells, drugs may be useful as pharmacological tools affecting different transitions or states in the force-generating cycle. Whereas the use of drugs as tools may be less specific than site-directed mutagenesis, it also has advantages. The motor protein function may be studied in vivo, with maintained ordering of the protein components, e.g. as in the muscle sarcomere, allowing more insight into the relationship between specific molecular events and contractile properties of muscle. A drug that has been used quite extensively in this context is butanedione monoxime. The usefulness of this drug is based on firm characterization of its effect on actomyosin function on the molecular level (3, 1013). More recently other drugs, like N-benzyl-p-toluene sulfonamide (14, 15) and blebbistatin (16), have been found to affect myosin function, and their effects at the molecular level have also been elucidated in some detail (14, 15, 17, 18). Both these drugs appear to affect the actomyosin interaction in a similar way as butanedione monoxime by inhibiting a step before (or very early in) the myosin power stroke, leading to the inhibition of actomyosin cross-bridge formation and force production.In contrast to the reduced isometric force, caused by the above mentioned drugs, the bipyridine compound amrinone (Fig. 1A) has been found to increase the isometric force production of fast intact skeletal muscles of the frog (19, 20) and mouse (21) and also of fast (but much less slow) skinned muscle fibers of the rat (22). In all the fast myosin preparations, the effect of about 1 mm amrinone on isometric force was associated with characteristic changes of the force-velocity relationship (Fig. 1B), including a reduced maximum velocity of shortening (1922) and a reduced curvature of the force-velocity relationship (1922). The latter effect was accompanied (20, 21) by a less pronounced deviation of the force-velocity relationship from the hyperbolic shape (23) at high loads. There have been different interpretations of the drug effects. It has been proposed (2022) that amrinone might competitively inhibit the MgATP binding by myosin. However, more recently, results from in vitro motility assay experiments (24) challenged this idea. These results showed that amrinone reduces the sliding velocity (Vmax) at saturating MgATP concentrations but not at MgATP concentrations close to, or below, the Km value for the hyperbolic relationship between MgATP concentration and sliding velocity. Such a combination of effects is consistent with a reduced MgADP release rate (24) but not with competitive inhibition of substrate binding. However, effects of amrinone on the MgADP release rate have not been directly demonstrated. Additionally, in view of the uncertainty about what step actually determines the sliding velocity at saturating [MgATP] (see above and Refs. 79), it is of interest to consider other possible drug effects that could account for the data of Klinth et al. (24). These include the following: 1) an increased drag force, e.g. because of enhancement of weak actomyosin interactions; 2) a reduced step length; and 3) effects of the drug on the rate of MgATP-induced dissociation of actomyosin.Open in a separate windowFIGURE 1.A, structure of amrinone. B, experimental force-velocity data obtained in the presence (filled symbols) and absence (open symbols) of 1.1 mm amrinone. The data, from intact single frog muscle fibers, were obtained at 2 °C and fitted by Hill''s (42) hyperbola (lines) for data truncated at 80% of the maximum isometric force. Filled line, equation fitted to control data, a/P0* = 0.185; P0*/P0 = 1.196. Dashed line, amrinone, a/P0* = 0.347; P0*/P0 = 1.009. Force-velocity data were obtained in collaboration with Professor K. A. P. Edman. Same data as in Fig. 8 of Ref. 20. Note a decrease in maximum sliding velocity and curvature of the force-velocity relationship at low force, in response to amrinone. Also note that amrinone caused increased isometric force and a reduced deviation of the force-velocity relationship from the Hill''s hyperbola at high force. All changes of the force-velocity relationship were statistically significant (20), and similar changes were later also observed in intact mouse muscle and skinned rat muscle fibers. Data in Fig. 1 are published by agreement with Professor K. A. P. Edman.To differentiate between these hypotheses for the amrinone effects, and to gain more general insight into fundamental aspects of muscle function (e.g. mechanisms underlying the force-velocity relationship), we here study the molecular effects of amrinone on fast skeletal muscle myosin preparations in the presence and absence of actin.In vitro motility assay studies at different ionic strengths suggest that drag forces, caused by increased fraction of myosin heads in weak binding states, are not important for the effect of amrinone on sliding velocity. Likewise, optical tweezers studies showed no effect of the drug on the myosin step length. Finally, ideas that amrinone should reduce sliding velocity by reduced rate of MgATP-induced dissociation could be discarded because the drug actually increased the rate of this process. Instead, we found an amrinone-induced increase in the MgADP affinity of heavy meromyosin (HMM) in the presence of actin. Interestingly, similar effects of amrinone were not observed using myosin S1. As discussed below, this result and other results point to an amrinone-induced reduction in the rate of a strain-dependent MgADP release step. Simulations, using a model modified from that of Edman et al. (25), support this proposed mechanism of action. The results are discussed in relation to fundamental mechanisms underlying the force-velocity relationship of fast skeletal muscle, including which step determines shortening velocity and the possible importance of inter-head cooperativity.  相似文献   
79.
Arabidopsis thaliana (L.) Heynh. Columbia wild type and a root hair-less mutant RM57 were grown on iron-containing and iron-deficient nutrient solutions. In both genotypes, ferric chelate reductase (FCR) of intact roots was induced upon iron deficiency and followed a Michaelis-Menten kinetic with a K m of 45 and 54 M FeIII-EDTA and a V max of 42 and 33 nmol Fe2+·(g FW)–1·min–1 for the wild type and the mutant, respectively. The pH optimum for the reaction was around pH 5.5. The approximately four fold stimulation of FCR activity was independent of formation of root hairs and/or transfer cells induced by iron deficiency. Iron-deficiency-induced chlorosis and the development of a rigid root habit disappeared when ferric chelate was applied to the leaves, while FCR activity remained unchanged. The time course of the responses to iron deficiency showed that morphological and physiological responses were controlled separately.Abbreviations FCR ferric chelate reductase - FW fresh weight Thanks are due to Klaas Sjollema (Department of Electronmicroscopy, University of Groningen, The Netherlands) for help with the electron microscopy sample preparation and especially to Dr. Uwe Santore (Heinrich-Heine-University for electron microscopy. This work was supported by the SCIENCE programm of the European community; P.R.M.) and a Personal Research Grant by the Ministerium für Wissenschaft und Forschung of Nordrhein-Westfalen (P.R.M.) and last, not least by the productive discussions in ECOTRANS B.V.  相似文献   
80.
The biogeochemical and stoichiometric signature of vegetation fire may influence post‐fire ecosystem characteristics and the evolution of plant ‘fire traits’. Phosphorus (P), a potentially limiting nutrient in many fire‐prone environments, might be particularly important in this context; however, the effects of fire on P cycling often vary widely. We conducted a global‐scale meta‐analysis using data from 174 soil studies and 39 litter studies, and found that fire led to significantly higher concentrations of soil mineral P as well as significantly lower soil and litter carbon:P and nitrogen:P ratios. These results demonstrate that fire has a P‐rich signature in the soil–plant system that varies with vegetation type. Further, they suggest that burning can ease P limitation and decouple the biogeochemical cycling of P, carbon and nitrogen. These effects resemble a transient reversion to an earlier stage of ecosystem development, and likely underpin at least some of fire's impacts on ecosystems and organisms.  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