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   9211篇
  免费   673篇
  国内免费   2篇
  9886篇
  2024年   4篇
  2023年   51篇
  2022年   126篇
  2021年   219篇
  2020年   132篇
  2019年   172篇
  2018年   226篇
  2017年   188篇
  2016年   312篇
  2015年   500篇
  2014年   581篇
  2013年   682篇
  2012年   819篇
  2011年   841篇
  2010年   550篇
  2009年   425篇
  2008年   555篇
  2007年   559篇
  2006年   484篇
  2005年   499篇
  2004年   421篇
  2003年   399篇
  2002年   346篇
  2001年   80篇
  2000年   56篇
  1999年   81篇
  1998年   90篇
  1997年   64篇
  1996年   54篇
  1995年   56篇
  1994年   45篇
  1993年   41篇
  1992年   24篇
  1991年   26篇
  1990年   27篇
  1989年   17篇
  1988年   13篇
  1987年   11篇
  1986年   13篇
  1985年   9篇
  1984年   11篇
  1983年   13篇
  1982年   15篇
  1981年   11篇
  1980年   5篇
  1979年   5篇
  1978年   7篇
  1977年   5篇
  1976年   5篇
  1975年   3篇
排序方式: 共有9886条查询结果,搜索用时 15 毫秒
91.
92.
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   
93.
We were interested whether PKC alpha, delta, epsilon or zeta is the isoform actually employed in the activation of hypertonicity-induced cation channels (HICCs) in primary cultures of rat hepatocytes. Quantitative SDS-page and Western-blot experiments revealed that PKC alpha, delta and epsilon were stimulated by Indolactam V (as a DAG substitute for activation of c and nPKCs) but that only PKC delta and epsilon did respond to hypertonic stress. Furthermore, chelation of intracellular Ca(++) by BAPTA-AM did not alter HICC activation in cable-analysis experiments whereas Indolactam V as well as V8 (an Indolactam derivative specific for PKC delta and epsilon) activated HICC currents under isotonic conditions. Finally, by use of Rottlerin (as an inhibitor exhibiting a slight preference for PKC delta over epsilon) PKC epsilon could be identified as the most likely isoform responsible for the activation of the HICC.  相似文献   
94.
We present a fast protocol that can be used to obtain highly purified cultures of proliferating adult human and rat Schwann cells accessible for non-viral transfection methods. The use of enriched genetically modified adult Schwann cells is of interest in the context of autologous cell transplantation within nerve transplants for peripheral nerve repair. Cell preparation from pre-degenerated adult peripheral nerves is described, together with the use of melanocyte growth medium plus forskolin, fibroblast growth factor-2 (FGF-2), pituitary extract and heregulin as a selective, serum-free culture medium and a subsequent cell enrichment step (cold jet). Proliferating adult Schwann cells can be efficiently genetically modified using optimized, non-viral electroporation protocols. The protocol results in Schwann cell cultures that are more than 90-95% pure, and transfection efficiencies vary depending on the initial cell constitution from 20 to 40%. The procedure takes up to 21 d, depending on the length of the pre-degeneration period.  相似文献   
95.
96.
The smaller isoform of the enzyme glutamic acid decarboxylase (GAD65) is a major islet autoantigen in autoimmune type 1 diabetes mellitus (T1DM). Transgenic plants expressing human GAD65 (hGAD65) are a potential means of direct oral administration of the islet autoantigen in order to induce tolerance and prevent clinical onset of disease. We have previously reported the successful generation of transgenic tobacco and carrot that express immunoreactive, full-length hGAD65. In the present study, we tested the hypothesis that the expression levels of recombinant hGAD65 in transgenic plants can be increased by targeting the enzyme to the plant cell cytosol and by mediating expression through the potato virus X (PVX) vector. By substituting the NH2-terminal region of hGAD65 with a homologous region of rat GAD67, a chimeric GAD671-87/GAD6588-585 molecule was expressed in transgenic tobacco plants. Immunolocalization analysis showed that immunoreactive GAD67/65 was found in the plant cell cytosol. By using a radio-immuno assay with human serum from a GAD65 autoantibody-positive T1DM patient, the highest expression level of the recombinant GAD67/65 protein was estimated to be 0.19% of the total soluble protein, compared to only 0.04% of wild-type hGAD65. Transient expression of wild-type, full-length hGAD65 in N. benthamiana mediated by PVX infection was associated with expression levels of immunoreactive protein as high as 2.2% of total soluble protein. This substantial improvement of the expression of hGAD65 in plants paves the way for immunoprevention studies of oral administration of GAD65-containing transgenic plant material in animal models of spontaneous autoimmune diabetes.  相似文献   
97.
OBJECTIVE: Fabry disease results from a deficiency in the activity of alpha-d-galactosidase A and subsequent accumulation of neutral glycosphingolipids in lysosomes. This study investigated whether lysosomal enzymes can indicate biochemical changes in the lysosomal apparatus induced by enzyme replacement therapy (ERT). DESIGN AND METHODS: Eight patients were monitored by clinical and biochemical tests and several lysosomal glycohydrolases were measured in plasma and leucocytes. RESULTS: Before starting ERT, beta-d-glucuronidase in leukocytes was markedly increased. After 1 month of therapy, enzyme levels dropped in all patients. In the patients who regularly followed the therapy, the enzyme levels remained stable for the next 20 months. In one patient who interrupted therapy for 2 months, the enzyme levels rose again. CONCLUSIONS: Lysosomal enzymes can be useful for monitoring biochemical changes in patients with Fabry disease receiving ERT. Though these findings refer to only a small number of patients, the correlation between beta-d-glucuronidase levels and ERT is interesting and might serve as a basis for further studies to define the potential of this enzyme in monitoring the effects of ERT in lysosomal storage disorders.  相似文献   
98.
PHOSPHO1, a phosphoethanolamine/phosphocholine phosphatase, is upregulated in mineralising cells and is thought to be involved in the generation of inorganic phosphate for bone mineralisation. PHOSPHO2 is a putative phosphatase sharing 42% sequence identity with PHOSPHO1. Both proteins contain three catalytic motifs, conserved within the haloacid dehalogenase superfamily. Mutation of Asp32 and Asp203, key residues within two motifs, abolish PHOSPHO1 activity and confirm it as a member of this superfamily. We also show that Asp43 and Asp123, residues that line the substrate-binding site in our PHOSPHO1 model, are important for substrate hydrolysis. Further comparative modelling reveals that the active sites of PHOSPHO1 and PHOSPHO2 are very similar, but surprisingly, recombinant PHOSPHO2 hydrolyses phosphoethanolamine and phosphocholine relatively poorly. Instead, PHOSPHO2 shows high specific activity toward pyridoxal-5-phosphate (V(max) of 633 nmol min(-1) mg(-1) and K(m) of 45.5 microM). Models of PHOSPHO2 and PHOSPHO1 suggest subtle differences in the charge distributions around the putative substrate entry site and in the location of potential H-bond donors.  相似文献   
99.
In vertebrates, homologous recombinational repair (HRR) requires RAD51 and five RAD51 paralogs (XRCC2, XRCC3, RAD51B, RAD51C and RAD51D) that all contain conserved Walker A and B ATPase motifs. In human RAD51D we examined the requirement for these motifs in interactions with XRCC2 and RAD51C, and for survival of cells in response to DNA interstrand crosslinks (ICLs). Ectopic expression of wild-type human RAD51D or mutants having a non-functional A or B motif was used to test for complementation of a rad51d knockout hamster CHO cell line. Although A-motif mutants complement very efficiently, B-motif mutants do not. Consistent with these results, experiments using the yeast two- and three-hybrid systems show that the interactions between RAD51D and its XRCC2 and RAD51C partners also require a functional RAD51D B motif, but not motif A. Similarly, hamster Xrcc2 is unable to bind to the non-complementing human RAD51D B-motif mutants in co-immunoprecipitation assays. We conclude that a functional Walker B motif, but not A motif, is necessary for RAD51D's interactions with other paralogs and for efficient HRR. We present a model in which ATPase sites are formed in a bipartite manner between RAD51D and other RAD51 paralogs.  相似文献   
100.
The male-specific regions of the Y chromosome (MSY) of the human and the chimpanzee (Pan troglodytes) are fully sequenced. The most striking difference is the dramatic rearrangement of large parts of their respective MSYs. These non-recombining regions include ampliconic gene families that are known to be important for male reproduction,and are consequently under significant selective pressure. However, whether the published Y-chromosomal pattern of ampliconic fertility genes is invariable within P. troglodytes is an open but fundamental question pertinent to discussions of the evolutionary fate of the Y chromosome in different primate mating systems. To solve this question we applied fluorescence in situ hybridisation (FISH) of testis-specific expressed ampliconic fertility genes to metaphase Y chromosomes of 17 chimpanzees derived from 11 wild-born males and 16 bonobos representing seven wild-born males. We show that of eleven P. troglodytes Y-chromosomal lines, ten Y-chromosomal variants were detected based on the number and arrangement of the ampliconic fertility genes DAZ (deleted in azoospermia) and CDY (chromodomain protein Y)—a so-far never-described variation of a species'' Y chromosome. In marked contrast, no variation was evident among seven Y-chromosomal lines of the bonobo, P. paniscus, the chimpanzee''s closest living relative. Although, loss of variation of the Y chromosome in the bonobo by a founder effect or genetic drift cannot be excluded, these contrasting patterns might be explained in the context of the species'' markedly different social and mating behaviour. In chimpanzees, multiple males copulate with a receptive female during a short period of visible anogenital swelling, and this may place significant selection on fertility genes. In bonobos, however, female mate choice may make sperm competition redundant (leading to monomorphism of fertility genes), since ovulation in this species is concealed by the prolonged anogenital swelling, and because female bonobos can occupy high-ranking positions in the group and are thus able to determine mate choice more freely.  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