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   19678篇
  免费   1967篇
  国内免费   2篇
  2023年   99篇
  2022年   207篇
  2021年   422篇
  2020年   253篇
  2019年   314篇
  2018年   383篇
  2017年   331篇
  2016年   571篇
  2015年   1052篇
  2014年   1061篇
  2013年   1312篇
  2012年   1677篇
  2011年   1595篇
  2010年   1023篇
  2009年   882篇
  2008年   1140篇
  2007年   1134篇
  2006年   1086篇
  2005年   1034篇
  2004年   1018篇
  2003年   896篇
  2002年   826篇
  2001年   261篇
  2000年   196篇
  1999年   228篇
  1998年   192篇
  1997年   128篇
  1996年   120篇
  1995年   86篇
  1994年   96篇
  1993年   88篇
  1992年   108篇
  1991年   117篇
  1990年   103篇
  1989年   76篇
  1988年   81篇
  1987年   68篇
  1986年   73篇
  1985年   71篇
  1984年   82篇
  1983年   64篇
  1982年   61篇
  1980年   50篇
  1978年   55篇
  1977年   63篇
  1976年   48篇
  1975年   58篇
  1974年   54篇
  1973年   77篇
  1972年   51篇
排序方式: 共有10000条查询结果,搜索用时 312 毫秒
81.
82.
Alginate Production by Plant-Pathogenic Pseudomonads   总被引:10,自引:4,他引:6       下载免费PDF全文
Eighteen plant-pathogenic and three non-plant-pathogenic pseudomonads were tested for the ability to produce alginic acid as an exopolysaccharide in vitro. Alginate production was demonstrated for 10 of 13 fluorescent plant-pathogenic pseudomonads tested with glucose or gluconate as the carbon source, but not for all 5 nonfluorescent plant pathogens and all 3 non-plant pathogens tested. With sucrose as the carbon source, some strains produced alginate while others produced both polyfructan (levan) and alginate. Alginates ranged from <1 to 28% guluronic acid, were acetylated, and had number-average molecular weights of 11.3 × 103 to 47.1 × 103. Polyfructans and alginates were not elicitors of the soybean phytoalexin glyceollin when applied to wounded cotyledon surfaces and did not induce prolonged water soaking of soybean leaf tissues. All or most pseudomonads in rRNA-DNA homology group I may be capable of synthesizing alginate as an exopolysaccharide.  相似文献   
83.
K M Lee  A G Marshall 《Biochemistry》1986,25(25):8245-8252
In this paper we report the first 1H NMR study of the base-paired secondary structure of yeast 5.8S RNA. On the basis of a combination of homonuclear Overhauser enhancements and temperature dependence of the proton 500-MHz NMR spectrum, we are able to identify and assign eight of the nine base pairs in the most thermally stable helical arm: G116.C137-C117.G136-C118.G135- C119.G134-C120.G133-U121.G132- U122.A131-G123.C130. This arm contains an unusually temperature-stable (to 71 degrees C) segment of four consecutive G.C base pairs. This work constitutes the most direct evidence to date for the existence and base-pair sequence of the GC-rich helix, which is common to most currently popular secondary structural models for eukaryotic 5.8S ribosomal RNA.  相似文献   
84.
Summary Suspensions of LLC-PK1 cells (a continuous epitheliod cell line with renal characteristics) are examined for mechanisms of intracellular pH regulation using the fluorescent probe BCECF. Initial experiments determine suitable calibration procedures for use of the BCECF fluorescent signal. They also determine that the cell suspension contains cells which (after 4 hr in suspension) have Na+ and K+ gradients comparable to those of cells in monolayer culture. The steady-state intracellular pH (7.05±0.01,n=5) of cells which have recovered in (pH 7.4) Na+-containing medium is not affected over several minutes by addition of 100 M amiloride or removal of extracellular Na+ (Na o + /H i + and Na i + /H o + exchange reactions are functionally inactive (compared to cellular buffering capacity). In contrast, Na o + /H i + exchange is activated by an increased cellular acid load. This activation may be observed directly either as a stimulation of net H+ efflux or net Na+ influx with decreasing intracellular pH. The extrapolation of this latter data suggests a set point of Na+/H+ exchange of approximately pH 7.0, consistent with the observed resting intracellular pH of approximately 7.05.  相似文献   
85.
Summary We have mapped and sequenced the globin gene and seven surrounding Alu repeat sequences in the orangutan globin gene cluster and have compared these and other orangutan sequences to orthologously related human sequences. Noncoding flanking and intron sequences, synonymous sites of , , and globin coding regions, and Alu sequences in human and orangutan diverge by 3.2%, 2.7%, and 3.7%, respectively. These values compare to 3.6% from DNA hybridizations and 3.4% from the globin gene region. If as suggested by fossil evidence and molecular clock calculations, human and orangutan lineages diverged about 10–15 MYA, the rate of noncoding DNA evolution in the two species is 1.0–1.5×10–9 substitutions per site per year. We found no evidence for either the addition or deletion of Alu sequences from the globin gene cluster nor is there any evidence for recent concerted evolution among the Alu sequences examined. Both phylogenetic and phenetic distance analyses suggest that Alu sequences within the and globin gene clusters arose close to the time of simian and prosimian primate divergence (about 50–60 MYA). We conclude that Alu sequences have been evolving at the rate typical of noncoding DNA for the majority of primate history.Presented at the FEBS Symposium on Genome Organization and Evolution, held in Crete, Greece, September 1–5, 1986  相似文献   
86.
Summary Injection of depolarizing current into vegetative cells of the water moldBlastocladiella emersonii elicits a regenerative response that has the electrical characteristics of an action potential. Once they have been taken past a threshold of about –40 mV, cells abruptly depolarize to +20 mV or above; after an interval ranging from several hundred milliseconds to a few seconds, the cells spontaneously return to their resting potential near –100 mV. When the action potential was analyzed with voltage-clamp recording, it proved to be biphasic. The initial phase reflects an influx of calcium ions through voltage-sensitive channels that also carry Sr2+ ions. The delayed, and more extended, phase of inward current results from the efflux of chloride and other anions. The anion channels are broadly selective, passing chloride, nitrate, phosphate, acetate, succinate and even PIPES. The anion channels open in response to the entry of calcium ions, but do not recognize Sr2+. Calcium channels, anion channels and calcium-specific receptors that link the two channels appear to form an ensemble whose physiological function is not known. Action potentials rarely occur spontaneously but can be elicited by osmotic downshock, suggesting that the ion channels may be involved in the regulation of turgor.  相似文献   
87.
Ten glycoproteins of molecular weights of 180,000, 150,000, 130,000, 115,000, 97,000, 77,000, 74,000, 64,000, 55,000, and 45,000 (designated as 180K, 150K, etc.) and a single nonglycosylated 107,000-molecular-weight (107K) protein were quantitatively removed from purified bovine herpesvirus 1 (BHV-1) virions by detergent treatment. Immunoprecipitations with monospecific and monoclonal antibodies showed that three sets of coprecipitating glycoproteins, 180K/97K, 150K/77K, and 130K/74K/55K, were the major components of the BHV-1 envelope. These glycoproteins were present in the envelope of the virion and on the surface of BHV-1-infected cells and reacted with neutralizing monoclonal and monospecific antibodies. Antibodies to 150K/77K protein had the largest proportion of virus-neutralizing antibodies, followed by antibodies to 180K/97K protein. Monoclonal antibodies to 130K/74K/55K protein were neutralizing but only in the presence of complement; however, monospecific antisera produced with 55K protein did not have neutralizing activity. Analysis under nonreducing conditions showed that the 74K and 55K proteins interact through disulfide bonds to form the 130K molecule. Partial proteolysis studies showed that the 180K protein was a dimeric form of the 97K protein and that the 150K protein was a dimer of the 77K protein, but these dimers were not linked by disulfide bonds. The 107K protein was not glycosylated and induced antibodies that did not neutralize BHV-1. The 64K protein was not precipitated by anti-BHV-1 convalescent antisera, and monospecific antisera to this protein precipitated several polypeptides from uninfected cell lysates, suggesting that 64K is a protein of cellular origin associated with the BHV-1 virion envelope.  相似文献   
88.
Rat ovarian tissue has been shown to contain high-affinity gonadotropin-releasing hormone (GnRH) receptors, and synthetic GnRH analogues have been shown to inhibit steroid production by rat corpora lutea in vivo and in vitro. These results raise the possibility that an ovarian GnRH-like peptide may be involved in normal luteal regression. We have examined binding of D-Ala6-des-Gly10-GnRH ethylamide (D-Ala) to rabbit corpora lutea, and have investigated the luteolytic activity of this analogue in hypophysectomized, pseudopregnant rabbits. Three hypophysectomized estrogen-treated rabbits were injected with 0.25 mg D-Ala s.c. every 6 h for 48 h during mid-pseudopregnancy, and three were injected with vehicle only. Treatment with D-Ala produced no acute changes in serum progesterone, nor was the time of luteal regression altered. Rabbit anterior pituitary tissue was found to contain high-affinity GnRH receptors (Ka = 7.0 X 10(9) M-1; 188.2 +/- 35.6 fmol/mg protein). However, no similar high-affinity GnRH receptors were detected in rabbit luteal tissue from any stage of pseudopregnancy. Some apparent low-affinity binding was observed, but this displaceable binding was subsequently observed in all control tissues tested. Thus, a potent GnRH analogue does not have any detectable direct effect on steroidogenesis in the rabbit corpus luteum, nor are high-affinity GnRH binding sites present in rabbit luteal tissue.  相似文献   
89.
Comparison of antisera from sheep during primary infection and following vaccination and challenge with Trichostrongylus colubriformis, with antisera obtained following primary infection of high- and low-responder guinea pigs, failed to reveal different antigenic patterns in proteins separated from fourth stage larval extracts by two-dimensional electrophoresis and probed by the immunoblot technique.Generally, serum IgG reacted specifically with worm antigens of mol. wt greater than 94,000, whereas protection against challenge infection was elicited most effectively in the guinea pig by fractions in the 67,000–94,000 range.Most distinct separations of larval proteins by SDS-polyacrylamide gel electrophoresis were obtained by extraction of live larvae and the extracts used within 2–3 days.  相似文献   
90.
Summary The experiments reported here evaluate the capability of isolated intestinal epithelial cells to accomplish net H+ transport in response to imposed ion gradients. In most cases, the membrane potential was kept constant by means of a K+ plus valinomycin voltage clamp in order to prevent electrical coupling of ion fluxes. Net H+ flux across the cellular membrane was examined at pH 6.0 (the physiological lumenal pH) and at pH 7.4 using methylamine distribution or recordings of changes in media pH. Results from both techniques suggest that the cells have an Na+/H+ exchange system in the plasma membrane that is capable of rapid and sustained changes in intracellular pH in response to an imposed Na+ gradient. The kinetics of the Na+/H+ exchange reaction at pH 6.0 [K t for Na+=57mm,V max=42 mmol H+/liter 3OMG (3-O-methylglucose) space/min] are dramatically different from those at pH 7.4 (K t for Na+=15mm,V max=1.7 mmol H+/liter 3OMG space/min). Experiments involving imposed K+ gradients suggest that these cells have negligible K+/H+ exchange capability. They exhibit limited but measurable H+ conductance. Anion exchange for base equivalents was not detected in experiments performed in media nominally free of bicarbonate.  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