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   7906篇
  免费   752篇
  国内免费   7篇
  8665篇
  2022年   55篇
  2021年   118篇
  2020年   66篇
  2019年   91篇
  2018年   118篇
  2017年   111篇
  2016年   176篇
  2015年   304篇
  2014年   345篇
  2013年   444篇
  2012年   526篇
  2011年   541篇
  2010年   381篇
  2009年   356篇
  2008年   479篇
  2007年   487篇
  2006年   436篇
  2005年   421篇
  2004年   429篇
  2003年   390篇
  2002年   385篇
  2001年   81篇
  2000年   60篇
  1999年   96篇
  1998年   134篇
  1997年   83篇
  1996年   83篇
  1995年   75篇
  1994年   79篇
  1993年   73篇
  1992年   55篇
  1991年   60篇
  1990年   72篇
  1989年   49篇
  1988年   55篇
  1987年   56篇
  1986年   40篇
  1985年   61篇
  1984年   58篇
  1983年   58篇
  1982年   62篇
  1981年   58篇
  1980年   60篇
  1979年   47篇
  1978年   42篇
  1977年   42篇
  1976年   48篇
  1974年   37篇
  1973年   47篇
  1970年   24篇
排序方式: 共有8665条查询结果,搜索用时 25 毫秒
151.
The incorporation of uridine into the nucleotide pool of actinomycin-treated, mengovirus-infected Novikoff rat hepatoma cells in culture follows simple Michaelis-Menten kinetics, and the apparent V(max) and K(m) values are similar to those for uridine transport by uninfected cells. Incorporation of uridine into mengovirus-specific ribonucleic acid (RNA) also follows Michaelis-Menten kinetics, and the apparent K(m) (about 10 mum) is approximately the same as for uridine transport. Inhibition of uridine transport by the presence of adenosine, persantin, or phenethyl alcohol inhibits simultaneously and to the same extent the incorporation of uridine into the nucleotide pool and into viral RNA, without affecting viral RNA synthesis per se. Phenethyl alcohol, however, also inhibits virus maturation. The inhibition of uridine incorporation into the nucleotide pool and into viral RNA is of the simple competitive type, indicating that transport into the cells is the rate-limiting step in the incorporation of uridine into mengovirus RNA. The results also indicate that treatment with actinomycin D or mengovirus infection does not affect uridine transport.  相似文献   
152.
The products of the SOS-regulated umuDC operon are required for most UV and chemical mutagenesis in Escherichia coli, a process that results from a translesion synthesis mechanism. The UmuD protein is activated for its role in mutagenesis by a RecA-facilitated autodigestion that removes the N-terminal 24 amino acids. A previous genetic screen for nonmutable umuD mutants had resulted in the isolation of a set of missense mutants that produced UmuD proteins that were deficient in RecA-mediated cleavage (J. R. Battista, T. Ohta, T. Nohmi, W. Sun, and G. C. Walker, Proc. Natl. Acad. Sci. USA 87:7190–7194, 1990). To identify elements of the UmuD′ protein necessary for its role in translesion synthesis, we began with umuD′, a modified form of the umuD gene that directly encodes the UmuD′ protein, and obtained missense umuD′ mutants deficient in UV and methyl methanesulfonate mutagenesis. The D39G, L40R, and T51I mutations affect residues located at the UmuD′2 homodimer interface and interfere with homodimer formation in vivo. The D75A mutation affects a highly conserved residue located at one end of the central strand in a three-stranded β-sheet and appears to interfere with UmuD′2 homodimer formation indirectly by affecting the structure of the UmuD′ monomer. When expressed from a multicopy plasmid, the L40R umuD′ mutant gene exhibited a dominant negative effect on a chromosomal umuD+ gene with respect to UV mutagenesis, suggesting that the mutation has an effect on UmuD′ function that goes beyond its impairment of homodimer formation. The G129D mutation affects a highly conserved residue that lies at the end of the long C-terminal β-strand and results in a mutant UmuD′ protein that exhibits a strongly dominant negative effect on UV mutagenesis in a umuD+ strain. The A30V and E35K mutations alter residues in the N-terminal arms of the UmuD′2 homodimer, which are mobile in solution.  相似文献   
153.
Potassium (K+) is an important nutrient for plants. It serves as a cofactor of various enzymes and as the major inorganic solute maintaining plant cell turgor. In a recent study, an as yet unknown role of K+ in plant homeostasis was shown. It was demonstrated that K+ gradients in vascular tissues can serve as an energy source for phloem (re)loading processes and that the voltage-gated K+ channels of the AKT2-type play a unique role in this process. The AKT2 channel can be converted by phosphorylation of specific serine residues (S210 and S329) into a non-rectifying channel that allows a rapid efflux of K+ from the sieve element/companion cells (SE/CC) complex. The energy of this flux is used by other transporters for phloem (re)loading processes. Nonetheless, the results do indicate that post-translational modifications at S210 and S329 alone cannot explain AKT2 regulation. Here, we discuss the existence of multiple post-translational modification steps that work in concert to convert AKT2 from an inward-rectifying into a non-rectifying K+ channel.Key words: potassium, channel, potassium channel, AKT2, phloem (re)loading, post-translational modifications, potassium batteryPotassium (K+) is the most abundant mineral element in plants, and together with nitrogen and phosphorous, is limiting for plant production in many natural and agricultural habitats. Voltage-gated K+ channels are key players in the acquisition of K+ ions from the soil and in its redistribution within the plant.1 Structurally, these channels result from the assembly of four so-called α-subunits. The subunits are encoded by nine genes in Arabidopsis and both homo- and hetero-tetramers are expressed.2,3 The K+ channel α-subunits can be categorized into four different subfamilies, based on the voltage-gating characteristics of the exogenous K+ conductance when expressed in an appropriate heterologous expression system. Kin α-subunits form hyperpolarization-activated channels that mediate K+ uptake.47 Kout α-subunits form depolarization-activated channels that mediate K+ release from cells.810 Ksilent subunits appear unable to yield functional homomeric channels, but can combine with Kin subunits and fine-tune the K+-uptake properties of the resulting heteromeric channels.1114 Finally, Kweak α-subunits form channels with complex voltage-gating; they allow both K+ uptake and release.1519 In Arabidopsis, a single member is found in this subfamily, AKT2, and this channel can assemble in heteromeric channels with the Kin subunit KAT2.20To date, only scarce and speculative information has been obtained for the function of Kweak channels. When expressed in heterologous expression systems, two different subpopulations of AKT2 channels differing in their sensitivity to voltage were found.21 Channels of the first type showed gating properties and currents analogous to that of Kin channels, while the other sort enabled a non-rectified (leak-like) current; they were open over the entire physiological voltage range.A given channel can be converted from one type to the other by post-translational modifications.21 A voltage-dependent phosphorylation was found to be an essential step for this switch,22,23 although the kinase responsible for this conversion still needs to be uncovered.24 In biophysical studies, mutant versions of the Arabidopsis Kweak channel subunit AKT2 have been created that showed impaired gating mode settings.22,23 Recently, Gajdanowicz et al. generated transgenic Arabidopsis thaliana plants that express these mutant AKT2 channels in the background of the akt2-1 null-allele plant.25 The major conclusion from analyses of these mutants is that the status switching of AKT2 from an inward-rectifying to a non-rectifying channel is crucial for plants to overcome energy-limiting conditions. This function of AKT2 could be correlated to its expression in phloem tissues. Selective expression of AKT2 under the control of the phloem companion cell-specific AtSUC2 promoter rescued the akt2-1 line, but conversely, selective expression of AKT2 under the control of the guard cell-specific GC1 promoter,26 resulted in further impairment of plant growth (Fig. 1). By combining diverse experimental approaches with mathematical simulation methods, an existing model for phloem (re)loading18,27 was fundamentally improved. This allowed the uncovering of a novel and interesting role of K+ in phloem physiology: K+ gradients present between the sieve element/companion cell (SE/CC) complex and the apoplast can serve as an energy source in phloem (re)loading processes. This “potassium battery” can be tapped by means of AKT2 regulation. This clarifies the observation of Deeken et al.28 that in AKT2 loss-of-function mutant plants, assimilates leaking away from the sieve tube were not efficiently reloaded into the main phloem stream.Open in a separate windowFigure 1AKT2 expressed only in guard cells delays plant development. (A–C) Representative wild-type, akt2-1 and akt2-1+pGC1:AKT2 complementation plants grown for 7 weeks (A), 9 weeks (B) and 12 weeks (C) under 12-h day/12-h night conditions at normal light intensity (150 µmol m−2 s−1). (D) akt2-1+pGC1:AKT2 developed a similar number of leaves as the akt2-1 knock out plants, but bolting-time was delayed. (B and E) After 9 weeks, wild-type plants were at an advanced bolting stage, akt2-1 plants had started bolting, but only initial signs of bolting were visible in akt2-1+pGC1:AKT2 plants. (C and F) At 12 weeks, akt2-1 plants had caught up with the wild-type and akt2-1+pGC1:AKT2 was just starting to bolt, although rosette-leaves were showing clear signs of senescence. For the generation of akt2-1+pGC1:AKT2, the AKT2 cDNA was fused to the guard cell-specific GC1 promoter26 kindly provided by J.I. Schroeder, San Diego. The pGC1:AKT2 construct was cloned into pGreen0229-35S by replacing the 35S promoter and then transformed into the akt2-1 knockout plant. All seeds were cold-treated for 24 h at 4°C. Plants were grown on artificial substrate (type GS-90, Einheitserde). After 2 weeks, seedlings were transferred to single pots. Plants were grown in 60% relative humidity at 21°C during the day and 18°C at night. Phenotypical analyses were done in the middle of the day. Data are shown as means ± SD of n ≥ 9 plants. Statistical analyses using Student''s t test: (D, WT/akt2-1: p < 2e-08; D, WT/pGC-AKT2: p < 2e-08; D, akt2-1/pGC-AKT2: p < 5e-03; E, WT/akt2-1: p < 4e-06; E, WT/pGC-AKT2: p < 1e-10; E, akt2-1/pGC-AKT2: p < 5e-04; F, WT/akt2-1: p = 0.51; F, WT/pGC-AKT2: p < 1e-10; F, akt2-1/pGC-AKT2: p < 1e-10).AKT2 expression is especially abundant in phloem tissues and the root stele, both of which are characterized by a poor availability of oxygen.29,30 This local internal hypoxia impairs respiratory activity of the vascular tissue and concomitantly, respiratory ATP production is reduced.31 As a consequence, phloem transport is very susceptible to decreasing oxygen supply to the plant.29,32 It is therefore comprehensible that the above mentioned support by the K+ driving force for sucrose retrieval is especially relevant in the phloem. Indeed Gajdanowicz et al.25 showed that transgenic plants lacking the AKT2 K+ channel were severely impaired in growth when exposed to mild hypoxia (10% v:v), whereas growth of wild-type plants was unaffected by this treatment. These observations illustrate the importance of biochemical flexibility in plant cells to cope with the energetic consequences of the steep oxygen concentration gradients that generally occur in plant stems and roots.In fact, the role of K+ gradients in driving sugar, amino acid and organic acid transport across plant cell membranes was first suggested several decades ago.33,34 Experimental evidence for this concept was provided by various tests in which pieces of plant tissue were incubated in solutions with different K+ concentrations and pH levels.33,34 Unfortunately, at that time the lack of genetic information to support this hypothesis (e.g., identifying transporter proteins that could provide a molecular mechanism to explain the working mechanism of substrate transport driven by a K+-motive force) resulted in this idea falling into oblivion. Indeed, the unequivocal experimental observation of this new role of K+ gradients in phloem reloading is extremely challenging. Under normal experimental conditions, K+ fluxes and sucrose fluxes are coupled during phloem loading in source tissues and unloading in sink tissues. Nonetheless, computational simulations predict that under certain conditions, a local K+/Suc antiport is also thermodynamically possible. In this antiport system, the energy from the K+ gradient is used to transport Suc into the phloem. This process is only transient; flooding the apoplast with K+ will decrease the K+ gradient. However, the gradient can be maintained for longer if surrounding cells take up the apoplastic K+ for their own use. A K+/Suc antiport will not occur in obvious sink or source tissues since the energy balances in such cells are fundamentally different. Consequently, in these tissues only the coupled symport of K+ and Suc can be observed. However, the computational predictions allowed the identification of the experimental conditions under which the effect of the K+/Suc antiport system is empirically observable at the whole plant level.An essential role in the regulation of AKT2 is played by (de)phosphorylation events of serine residues at positions S210 and S329. The replacement of both serines by asparagine (AKT2-S210N-S329N) resulted in a K+-selective leak that is locked in a continuously open mode when the channels are expressed in Xenopus oocytes. Under certain conditions, plants expressing the AKT2-S210N-S329N mutation showed growth benefits over wild-type plants; akt2-1+AKT2-S210N-S329N plants reach the generative state faster, possess an increased number of leaves and increased fresh weight (Fig. 2). Intuitively, one would expect a continuously open channel to cause severe problems for the plant, not a benefit as was observed here. We therefore have to postulate that phosphorylation at residues AKT2-S210 and AKT2-S329 is insufficient for converting AKT2 from an inward-rectifying into a non-rectifying channel; other, as yet unknown mechanisms, must contribute to the switch in the AKT2 gating mode. Such a concept would correspond to results that would otherwise be hard to explain. For instance, when both serine residues were replaced by glutamate, the mutant AKT2-S210E-S329E still showed wild-type characteristics.22 The S to E substitution is expected to mimic the phosphorylated state better than the S to N replacement. Furthermore, position AKT2-K197 has a fundamental influence on the AKT2 gating mode.23 AKT2 mutants with that particular lysine substituted with a serine are far less sensitive towards (de)phosphorylation; they display the characteristics of a pure inward-rectifying K+ channel,23 and transgenic Arabidopsis plants expressing AKT2 channels with this substitution showed the characteristics of akt2-1 knock-out plants.25 Initially, it was proposed that the positive charge is important for sensitizing AKT2 to phosphorylation. However, the charge-conserving mutant AKT2-K197R is similar to the charge inverting mutant AKT2-K197D,23 a purely inward-rectifying channel (Fig. 3). We therefore need to take into account that in plants, K197 may also be a target of post-translational modification.35 At present, we can explain the beneficial effect of the AKT2-S210N-S329N mutant on plant growth only by a multiple step regulation of AKT2 (Fig. 4). The double-N mutation would then bypass the phosphorylation step, but AKT2-S210N-S329N could still be deregulated into an inward-rectifying channel. Thus, AKT2 can be considered as a highly specialized Kin channel that can be converted into a leak-like channel by a cascade of post-translational modification steps.Open in a separate windowFigure 2Plants expressing the AKT2-S210N-S329N mutant reach the generative state faster than wild-type plants. The mutant channel AKT2-S210N-S329N was expressed under the control of the native AKT2 promoter in the akt2-1 knock-out background. (A) Photos of representative Arabidopsis thaliana plants grown 7 weeks under short day conditions (12-h day/12-h night, light intensity = 150 µE m−2s−1). Seven weeks after sowing, plants expressing only AKT2-S210N-S329N mutant channels (n = 22) differed significantly (Student''s t test, p < 4e-05) from wild-type plants (n = 20) in the height of the main inflorescent stalk (B) and fresh weight (C). At later time points, these differences decrease.25Open in a separate windowFigure 3The mutant AKT2-K197R channel is inward-rectifying. Steady-state current-voltage characteristics measured at the end of activation voltage steps. Currents were normalized to the current values measured at −145 mV in 10 mM K+ and are shown as means ± SD (n = 6).Open in a separate windowFigure 4Minimal model for AKT2 gating-mode regulation. To switch AKT2 from an inward-rectifying into a non-rectifying channel, at least two post-translational steps are postulated. (1) Phosphorylation at residues AKT2-S210 and AKT2-S329 (transitions [1]→[2] and [3]→[4]) and (2) a yet unknown modification that most likely involves the residue AKT2-K197 (transitions [1]→[3] and [2]→[4]). Only after both modifications will AKT2 allow the efflux of K+ (state [4]).  相似文献   
154.
To determine the organization of presympathetic vasomotor drive by phenotypic populations of rostral ventrolateral medulla (RVLM) neurons, we examined the somatosympathetic reflex (SSR) evoked in four sympathetic nerves together with selective lesions of RVLM presympathetic neurons. Urethane-anesthetized (1.3 g/kg ip), paralyzed, vagotomized and artificially ventilated Sprague-Dawley rats (n = 41) were used. First, we determined the afferent inputs activated by sciatic nerve (SN) stimulation at graded stimulus intensities (50 sweeps at 0.5-1 Hz, 1-80 V). Second, we recorded sympathetic nerve responses (cervical, renal, splanchnic, and lumbar) to intensities of SN stimulation that activated A-fiber afferents (low) or both A- and C-fiber afferents (high). Third, with low-intensity SN stimulation, we examined the cervical SSR following RVLM microinjection of somatostatin, and we determined the splanchnic SSR in rats in which presympathetic C1 neurons were lesioned following intraspinal injections of anti-dopamine-β-hydroxylase-saporin (anti-DβH-SAP). Low-intensity SN stimulation activated A-fiber afferents and evoked biphasic responses in the renal, splanchnic, and lumbar nerves and a single peak in the cervical nerve. Depletion of presympathetic C1 neurons (59 ± 4% tyrosine hydroxylase immunoreactivity profiles lesioned) eliminated peak 2 of the splanchnic SSR and attenuated peak 1, suggesting that only RVLM neurons with fast axonal conduction were spared. RVLM injections of somatostatin abolished the single early peak of cervical SSR confirming that RVLM neurons with fast axonal conduction were inhibited by somatostatin. It is concluded that unmyelinated RVLM presympathetic neurons, presumed to be all C1, innervate splanchnic, renal, and lumbar but not cervical sympathetic outflows, whereas myelinated C1 and non-C1 RVLM neurons innervate all sympathetic outflows examined. These findings suggest that multiple levels of neural control of vasomotor tone exist; myelinated populations may set baseline tone, while unmyelinated neurons may be recruited to provide actions at specific vascular beds in response to distinct stressors.  相似文献   
155.
The cytoplasmic N-terminal domain of the human ether-a-go-go related gene (hERG) K+ channel is critical for the slow deactivation kinetics of the channel. However, the mechanism(s) by which the N-terminal domain regulates deactivation remains to be determined. Here we show that the solution NMR structure of the N-terminal 135 residues of hERG contains a previously described Per-Arnt-Sim (PAS) domain (residues 26-135) as well as an amphipathic α-helix (residues 13-23) and an initial unstructured segment (residues 2-9). Deletion of residues 2-25, only the unstructured segment (residues 2-9) or replacement of the α-helix with a flexible linker all result in enhanced rates of deactivation. Thus, both the initial flexible segment and the α-helix are required but neither is sufficient to confer slow deactivation kinetics. Alanine scanning mutagenesis identified R5 and G6 in the initial flexible segment as critical for slow deactivation. Alanine mutants in the helical region had less dramatic phenotypes. We propose that the PAS domain is bound close to the central core of the channel and that the N-terminal α-helix ensures that the flexible tail is correctly orientated for interaction with the activation gating machinery to stabilize the open state of the channel.  相似文献   
156.

Background

Neuron specific enolase (ENO2, γ-enolase) has been used as a biomarker to help identify neuroendocrine differentiation in breast cancer. The goal of the present study was to determine if ENO2 expression in the breast epithelial cell is influenced by the environmental pollutants, arsenite and cadmium. Acute and chronic exposure of MCF-10A cells to As+3 and Cd+2 sufficient to allow colony formation in soft agar, was used to determine if ENO2 expression was altered by these pollutants.

Results

It was shown that both As+3 and Cd+2 exposure caused significant increases in ENO2 expression under conditions of both acute and chronic exposure. In contrast, ENO1, the major glycolytic enolase in non-muscle and neuronal cells, was largely unaffected by exposure to either As+3 or Cd+2. Localization studies showed that ENO2 in the MCF-10A cells transformed by As+3 or Cd+2 had both a cytoplasmic and nuclear localization. In contrast, ENO1 was localized to the cytoplasm. ENO2 localized to the cytoplasm was found to co-localized with ENO1.

Conclusion

The results are the first to show that ENO2 expression in breast epithelial cells is induced by acute and chronic exposure to As+3 or Cd+2. The findings also suggest a possible link between As+3 and Cd+2 exposure and neuroendocrine differentiation in tumors. Overall, the results suggest that ENO2 might be developed as a biomarker indicating acute and/or chronic environmental exposure of the breast epithelial cell to As+3 and Cd+2.  相似文献   
157.
Eosinophilic esophagitis is characterized by increased infiltration and degranulation of eosinophils in the esophagus. Whether eosinophil-derived cationic proteins regulate esophageal sensory nerve function is still unknown. Using synthetic cationic protein to investigate such effect, we performed extracellular recordings from vagal nodose or jugular neurons in ex vivo esophageal-vagal preparations with intact nerve endings in the esophagus. Nerve excitabilities were determined by comparing action potentials evoked by esophageal distensions before and after perfusion of synthetic cationic protein poly-L-lysine (PLL) with or without pretreatment with poly-L-glutamic acid (PLGA), which neutralized cationic charges of PLL. Perfusion with PLL did not evoke action potentials in esophageal nodose C fibers but increased their responses to esophageal distension. This potentiation effect lasted for 30 min after washing out of PLL. Pretreatment with PLGA significantly inhibited PLL-induced mechanohyperexcitability of esophageal nodose C fibers. In esophageal nodose Aδ fibers, perfusion with PLL did not evoke action potentials. In contrast to nodose C fibers, both the spontaneous discharges and the responses to esophageal distension in nodose Aδ fibers were decreased by perfusion with PLL, which can be restored after washing out PLL for 30-60 min. Pretreatment with PLGA attenuated PLL-induced decrease in spontaneous discharge and mechanoexcitability of esophageal nodose Aδ fibers. In esophageal jugular C fibers, PLL neither evoked action potentials nor changed their responses to esophageal distension. Collectively, these data demonstrated that synthetic cationic protein did not evoke action potential discharges of esophageal vagal afferents but had distinctive sensitization effects on their responses to esophageal distension.  相似文献   
158.
Markers of stemness in equine mesenchymal stem cells: a plea for uniformity   总被引:1,自引:0,他引:1  
Mesenchymal stromal cells (MSC) are a very promising subpopulation of adult stem cells for cell-based regenerative therapies in veterinary medicine. Despite major progress in the knowledge on adult stem cells during recent years, a proper identification of MSC remains a challenge. In human medicine, the Mesenchymal and Tissue Stem Cell Committee of the International Society for Cellular Therapy (ISCT) recently proposed three criteria to define MSC. Firstly, cells must be plastic-adherent when maintained under standard culture conditions. Secondly, MSC must express CD73, CD90 and CD105, and lack expression of CD34, CD45, CD14 or CD11b, CD79α or CD19 and MHC class II antigens. Thirdly, MSC must be able to differentiate into osteoblasts, adipocytes and chondroblasts in vitro. Successful isolation and differentiation of equine MSC from different sources such as bone marrow, fat tissue, umbilical cord blood, Wharton's Jelly or peripheral blood has been widely reported. However, their unequivocal immunophenotyping is hampered by the lack of a single specific marker and the limited availability of monoclonal anti-horse antibodies, which are two major factors complicating successful research on equine MSC. Detection of gene expression on mRNA level is hereby a valuable alternative, although the need still exists to test several antibody clones in search for cross-reactivity. To date, commercial antibodies recognizing equine epitopes are only available for CD13, CD44 and MHC-II. Moreover, as the expression of certain adult stem cell markers may differ between species, it is mandatory to define a set of CD markers which can be uniformly applied for the identification of equine MSC.  相似文献   
159.
The effect of transforming growth factor beta-1 (TGF1) expression on fatty acid binding proteins was examined in control and two strains of gene targeted TGF1-deficient mice. Homozygous TGF1-deficient 129 × CF-1, expressing multifocal inflammatory syndrome, had 25% less liver fatty acid binding protein (L-FABP) when compared to control mice. The decrease in L-FABP expression was not due to multifocal inflammatory syndrome since homozygous TGF1-deficient/immunodeficient C3H mice on a SLID background had 36% lower liver L-FABP than controls. This effect was developmentally related and specific to liver, but not the proximal intestine, where L-FABP is also expressed. Finally, the proximal intestine also expresses intestinal-FABP (1-FABP) which decreased 3-fold in the TGF1-deficient/immunodeficient C3H mice only. Thus, TGF1 appears to regulate the expression of L-FABP and I-FABP in the liver and the proximal intestine, respectively.Abbreviations L-FABP liver fatty acid binding protein - I-FABP intestinal fatty acid binding protein - TGF1 transforming growth factor beta-1 - TNF- tumor necrosis factor- - MIP- macrophage inflammatory protein- - PMSF phenylmethyl sulfonyl fluoride - PBS phosphate buffered saline  相似文献   
160.
The sporophyte-gametophyte junction in Acaulon muticum is composed of the sporophyte foot, the surrounding gametophyte vaginula, and an intervening placental space. At an early stage of development the foot has a large basal cell, characterized by extensive wall ingrowths beginning at the lowermost tip of the basal cell and extending along its tangential walls. Sporophyte cells in contact with the basal cell develop ingrowths on their outer tangential walls and on radial walls in contact with the basal cell. All sporophyte cells at this stage are characterized by numerous mitochondria, strands of endoplasmic reticulum, and dictyosomes, particularly in the cytoplasm adjacent to areas of extensive wall development. Plastids typically contain abundant starch reserves. As development proceeds, wall ingrowths become more extensive on all walls in the sporophyte foot but are never found on the upper wall of the basal cell in contact with the remainder of the sporophyte. Plastids in the foot contain fewer starch reserves later in development. Wall ingrowths are not visible in the cells of the gametophyte vaginula until well after extensive development has occurred in the sporophyte foot. Stacks or layers of endoplasmic reticulum are characteristic of the cells of the gametophyte vaginula, along with numerous mitochondria, dictyosomes, and well-developed plastids. Starch reserves typically are less abundant in cells of the gametophyte. The early development of extensive wall elaborations in the cells of the sporophyte foot, and particularly in the basal cell, may favor the rapid movement of water and nutrients from the gametophyte into the sporophyte at a time when rapid development in this minute, ephemeral moss is critical.  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