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   303篇
  免费   23篇
  2023年   2篇
  2022年   7篇
  2021年   21篇
  2020年   7篇
  2019年   7篇
  2018年   11篇
  2017年   9篇
  2016年   10篇
  2015年   19篇
  2014年   25篇
  2013年   35篇
  2012年   30篇
  2011年   18篇
  2010年   7篇
  2009年   16篇
  2008年   13篇
  2007年   9篇
  2006年   9篇
  2005年   11篇
  2004年   16篇
  2003年   3篇
  2002年   6篇
  2001年   2篇
  2000年   2篇
  1999年   1篇
  1998年   4篇
  1997年   1篇
  1996年   3篇
  1994年   1篇
  1993年   2篇
  1992年   2篇
  1991年   1篇
  1990年   1篇
  1987年   1篇
  1986年   2篇
  1984年   2篇
  1983年   2篇
  1982年   1篇
  1981年   3篇
  1980年   1篇
  1979年   1篇
  1975年   1篇
  1964年   1篇
排序方式: 共有326条查询结果,搜索用时 31 毫秒
291.
Omega‐3 (also called n‐3) long‐chain polyunsaturated fatty acids (≥C20; LC‐PUFAs) are of considerable interest, based on clear evidence of dietary health benefits and the concurrent decline of global sources (fish oils). Generating alternative transgenic plant sources of omega‐3 LC‐PUFAs, i.e. eicosapentaenoic acid (20:5 n‐3, EPA) and docosahexaenoic acid (22:6 n‐3, DHA) has previously proved problematic. Here we describe a set of heterologous genes capable of efficiently directing synthesis of these fatty acids in the seed oil of the crop Camelina sativa, while simultaneously avoiding accumulation of undesirable intermediate fatty acids. We describe two iterations: RRes_EPA in which seeds contain EPA levels of up to 31% (mean 24%), and RRes_DHA, in which seeds accumulate up to 12% EPA and 14% DHA (mean 11% EPA and 8% DHA). These omega‐3 LC‐PUFA levels are equivalent to those in fish oils, and represent a sustainable, terrestrial source of these fatty acids. We also describe the distribution of these non‐native fatty acids within C. sativa seed lipids, and consider these data in the context of our current understanding of acyl exchange during seed oil synthesis.  相似文献   
292.

Objectives

We compared the accuracy of NOGA endocardial mapping for delineating transmural and non-transmural infarction to the results of cardiac magnetic resonance imaging (cMRI) with late gadolinium enhancement (LE) for guiding intramyocardial reparative substance delivery using data from experimental myocardial infarction studies.

Methods

Sixty domestic pigs underwent diagnostic NOGA endocardial mapping and cMRI-LE 60 days after induction of closed-chest reperfused myocardial infarction. The infarct size was determined by LE of cMRI and by delineation of the infarct core on the unipolar voltage polar map. The sizes of the transmural and non-transmural infarctions were calculated from the cMRI transmurality map using signal intensity (SI) cut-offs of>75% and>25% and from NOGA bipolar maps using bipolar voltage cut-off values of <0.8 mV and <1.9 mV. Linear regression analysis and Bland-Altman plots were used to determine correlations and systematic differences between the two images. The overlapping ratios of the transmural and non-transmural infarcted areas were calculated.

Results

Infarct size as determined by 2D NOGA unipolar voltage polar mapping correlated with the 3D cMRI-LE findings (r = 0.504, p<0.001) with a mean difference of 2.82% in the left ventricular (LV) surface between the two images. Polar maps of transmural cMRI and bipolar maps of NOGA showed significant association for determining of the extent of transmural infarction (r = 0.727, p<0.001, overlap ratio of 81.6±11.1%) and non-transmural infarction (r = 0.555, p<0.001, overlap ratio of 70.6±18.5%). NOGA overestimated the transmural scar size (6.81% of the LV surface) but slightly underestimated the size of the non-transmural infarction (−3.04% of the LV surface).

Conclusions

By combining unipolar and bipolar voltage maps, NOGA endocardial mapping is useful for accurate delineation of the targeted zone for intramyocardial therapy and is comparable to cMRI-LE. This may be useful in patients with contraindications for cMRI who require targeted intramyocardial regenerative therapy.  相似文献   
293.
Aging leads to a proinflammatory state within the vasculature without disease, yet whether this inflammatory state occurs during atherogenesis remains unclear. Here, we examined how aging impacts atherosclerosis using Ldlr?/? mice, an established murine model of atherosclerosis. We found that aged atherosclerotic Ldlr?/? mice exhibited enhanced atherogenesis within the aorta. Aging also led to increased LDL levels, elevated blood pressure on a low‐fat diet, and insulin resistance after a high‐fat diet (HFD). On a HFD, aging increased a monocytosis in the peripheral blood and enhanced macrophage accumulation within the aorta. When we conducted bone marrow transplant experiments, we found that stromal factors contributed to age‐enhanced atherosclerosis. To delineate these stromal factors, we determined that the vasculature exhibited an age‐enhanced inflammatory response consisting of elevated production of CCL‐2, osteopontin, and IL‐6 during atherogenesis. In addition, in vitro cultures showed that aging enhanced the production of osteopontin by vascular smooth muscle cells. Functionally, aged atherosclerotic aortas displayed higher monocyte chemotaxis than young aortas. Hence, our study has revealed that aging induces metabolic dysfunction and enhances vascular inflammation to promote a peripheral monocytosis and macrophage accumulation within the atherosclerotic aorta.  相似文献   
294.
Arl13b belongs to the ADP-ribosylation factor family within the Ras superfamily of regulatory GTPases. Mutations in Arl13b cause Joubert syndrome, which is characterized by congenital cerebellar ataxia, hypotonia, oculomotor apraxia, and mental retardation. Arl13b is highly enriched in cilia and is required for ciliogenesis in multiple organs. Nevertheless, the precise role of Arl13b remains elusive. Here we report that the exocyst subunits Sec8, Exo70, and Sec5 bind preferentially to the GTP-bound form of Arl13b, consistent with the exocyst being an effector of Arl13b. Moreover, we show that Arl13b binds directly to Sec8 and Sec5. In zebrafish, depletion of arl13b or the exocyst subunit sec10 causes phenotypes characteristic of defective cilia, such as curly tail up, edema, and abnormal pronephric kidney development. We explored this further and found a synergistic genetic interaction between arl13b and sec10 morphants in cilia-dependent phenotypes. Through conditional deletion of Arl13b or Sec10 in mice, we found kidney cysts and decreased ciliogenesis in cells surrounding the cysts. Moreover, we observed a decrease in Arl13b expression in the kidneys from Sec10 conditional knockout mice. Taken together, our results indicate that Arl13b and the exocyst function together in the same pathway leading to functional cilia.  相似文献   
295.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   
296.
Selection and transport of objects to use as tools at a distant site are considered to reflect planning. Ancestral humans transported tools and tool-making materials as well as food items. Wild chimpanzees also transport selected hammer tools and nuts to anvil sites. To date, we had no other examples of selection and transport of stone tools among wild nonhuman primates. Wild bearded capuchins (Cebus libidinosus) in Boa Vista (Piauí, Brazil) routinely crack open palm nuts and other physically well-protected foods on level surfaces (anvils) using stones (hammers) as percussive tools. Here we present indirect evidence, obtained by a transect census, that stones suitable for use as hammers are rare (study 1) and behavioral evidence of hammer transport by twelve capuchins (study 2). To crack palm nuts, adults transported heavier and harder stones than to crack other less resistant food items. These findings show that wild capuchin monkeys selectively transport stones of appropriate size and hardness to use as hammers, thus exhibiting, like chimpanzees and humans, planning in tool-use activities.  相似文献   
297.
Providing accurate animals’ phylogenies rely on increasing knowledge of neglected phyla. Tardigrada diversity evaluated in broad phylogenies (among phyla) is biased towards eutardigrades. A comprehensive phylogeny is demanded to establish the representative diversity and propose a more natural classification of the phylum. So, we have performed multilocus (18S rRNA and 28S rRNA) phylogenies with Bayesian inference and maximum likelihood. We propose the creation of a new class within Tardigrada, erecting the order Apochela (Eutardigrada) as a new Tardigrada class, named Apotardigrada comb. n. Two groups of evidence support its creation: (a) morphological, presence of cephalic appendages, unique morphology for claws (separated branches) and wide‐elongated buccopharyngeal apparatus without placoids, and (b) phylogenetic support based on molecular data. Consequently, order Parachela is suppressed and its superfamilies erected as orders within Eutardigrada, maintaining their current names. We propose a new classification within the family Echiniscidae (Echiniscoidea, Heterotardigrada) with morphological and phylogenetic support: (a) subfamily Echiniscinae subfam. n., with two tribes Echiniscini tribe n. and Bryodelphaxini tribe n.; (b) subfamily Pseudechiniscinae subfam. n., with three tribes Cornechiniscini tribe n., Pseudechiniscini tribe n. and Anthechiniscini tribe n.; and (c) subfamily Parechiniscinae subfam. n., with two tribes Parechiniscini tribe n. and Novechiniscini tribe n. Reliable biodiversity selection for tardigrades in broad phylogenies is proposed due to biased analyses performed up to now. We use our comprehensive molecular phylogeny to evaluate the evolution of claws in the clawless genus Apodibius and claw reduction across the Tardigrada tree of life. Evolutionary consequences are discussed.  相似文献   
298.
Molecular Diagnosis & Therapy - Fetal growth restriction is a pathological condition occurring when the fetus does not reach the genetically determined growth potential. The etiology of fetal...  相似文献   
299.
The biodiversity crisis has had particularly harsh impacts on marine environments. However, there is still considerable uncertainty about how many species have been seriously impacted and the effectiveness of protection measures (e.g., marine protected areas or MPAs) due to high levels of cryptic species in many taxa. Here, we employ an integrative taxonomy approach to mullet species in the genus Mugil. In addition to its high economic value, this genus is notable for having diversified ~29 million years ago without marked morphological and ecological divergence. We obtained 129 specimens of Mugil from the Coral Coast MPA, the largest of its kind in the Tropical Southwestern Atlantic marine province. Although morphometric and meristic traits revealed six taxonomically recognized species, only five mitochondrial lineages were observed. All individuals morphologically identified as M. incilis belonged to the mitochondrial lineage of Mugil curema, which is consistent with misidentification of morphologically similar species and an overestimation of species diversity. Remarkably, Mugil species in our sample that diverged up to ~23 million years ago are also the most morphologically similar (M. curema and M. rubrioculus), suggesting extreme morphological conservatism, possibly driven by similarities in habitat use and life‐history traits. This study demonstrates the potential utility of integrative taxonomy (including DNA barcoding) for contributing to the conservation and sustainable use of natural resources.  相似文献   
300.
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