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   51篇
  免费   12篇
  2021年   1篇
  2020年   1篇
  2016年   5篇
  2015年   1篇
  2014年   4篇
  2013年   5篇
  2012年   2篇
  2011年   2篇
  2010年   3篇
  2009年   2篇
  2008年   2篇
  2007年   2篇
  2006年   1篇
  2005年   4篇
  2004年   2篇
  2002年   2篇
  2001年   3篇
  2000年   2篇
  1993年   1篇
  1992年   1篇
  1991年   2篇
  1990年   3篇
  1988年   2篇
  1987年   4篇
  1985年   2篇
  1982年   1篇
  1973年   1篇
  1970年   1篇
  1968年   1篇
排序方式: 共有63条查询结果,搜索用时 15 毫秒
1.
The photoinduction period of Oenothera biennis L. seed germination was analyzed by varying the photoinduction temperature and by substituting red light pulses for continuous red light. At 24°C, seeds require 36 hours of continuous red light for maximal percent germination. The optimal photoinduction temperature is 32°C, with higher and lower temperatures being strongly inhibitory. A 30 minute exposure to far-red light, given immediately after a red light period of 1 to 36 hours, reduces germination by about 25%. Seeds escape from far-red inhibition with a half-time of 5 to 10 hours, depending on the length of the red exposure that precedes the far-red light. Periodic 15 minute pulses of red light can substitute for continuous red light in stimulating germination. Ted red light pulses, with 6 hours of darkness between successive pulses, cause maximal germination. The response to periodic red light is fully reversible by far-red light. Probit analysis of the periodic light response shows that as the length of the dark periods between successive pulses increases, less incident light is needed to induce germination but the population variance in light sensitivity remains constant. Probit analysis of the temperature response shows that as the photoinduction temperature increases from 16 to 32°C, less incident light is needed to induce germination and the population variance in light sensitivity also increases.  相似文献   
2.
Light is required for the herbicide activity of diphenyl ether herbicides. An action spectrum of acifluorfen-methyl activity with Chlamydomonas eugametos (Moewus) determined that cell death occurred at two peaks of light; 450 and 670 nanometers. These data indicate both chlorophyll and carotenoids, but not riboflavin, are involved in herbicide toxicity.  相似文献   
3.
Photosynthesis is not required for the toxicity of diphenyl ether herbicides, nor are chloroplast thylakoids the primary site of diphenyl ether herbicide activity. Isolated spinach (Spinacia oleracea L.) chloroplast fragments produced malonyl dialdehyde, indicating lipid peroxidation, when paraquat (1,1′-dimethyl-4,4′-bipyridinium ion) or diuron [3-(3,4-dichlorophenyl)-1,1-dimethylurea] were added to the medium, but no malonyl dialdehyde was produced when chloroplast fragments were treated with the methyl ester of acifluorfen (methyl 5-[2-chloro-4-(trifluoromethyl)phenoxy]-2-nitrobenzoic acid), oxyfluorfen [2-chloro-1-(3-ethoxy-4-nitrophenoxy)-4-(trifluoromethyl)benzene], or MC15608 (methyl 5-[2-chloro-4-(trifluoromethyl)phenoxy]-2-chlorobenzoate). In most cases the toxicity of acifluorfen-methyl, oxyfluorfen, or MC15608 to the unicellular green alga Chlamydomonas eugametos (Moewus) did not decrease after simultaneous treatment with diuron. However, diuron significantly reduced cell death after paraquat treatment at all but the highest paraquat concentration tested (0.1 millimolar). These data indicate electron transport of photosynthesis is not serving the same function for diphenyl ether herbicides as for paraquat. Additional evidence for differential action of paraquat was obtained from the superoxide scavenger copper penicillamine (copper complex of 2-amino-3-mercapto-3-methylbutanoic acid). Copper penicillamine eliminated paraquat toxicity in cucumber (Cucumis sativus L.) cotyledons but did not reduce diphenyl ether herbicide toxicity.  相似文献   
4.
In higher‐latitude trees, temperature and photoperiod control the beginning and end of the photosynthetically active season. Elevated temperature (ET) has advanced spring warming and delayed autumn cooling while photoperiod remains unchanged. We assessed the effects of warming on the length of the photosynthetically active season of three provenances of Pinus strobus L. seedlings from different latitudes, and evaluated the accuracy of the photochemical reflectance index (PRI) and the chlorophyll/carotenoid index (CCI) for tracking the predicted variation in spring and autumn phenology of photosynthesis among provenances. Seedlings from northern, local and southern P. strobus provenances were planted in a temperature‐free‐air‐controlled enhancement (T‐FACE) experiment and exposed to ET (+1.5/3°C; day/night). Over 18 months, we assessed photosynthetic phenology by measuring chlorophyll fluorescence, gas exchange, leaf spectral reflectance and pigment content. During autumn, all seedlings regardless of provenance followed the same sequence of phenological events with the initial downregulation of photosynthesis, followed by the modulation of non‐photochemical quenching and associated adjustments of zeaxanthin pool sizes. However, the timing of autumn downregulation differed between provenances, with delayed onset in the southern provenance (SP) and earlier onset in the northern relative to the local provenance, indicating that photoperiod at the provenance origin is a dominant factor controlling autumn phenology. Experimental warming further delayed the downregulation of photosynthesis during autumn in the SP. A provenance effect during spring was also observed but was generally not significant. The vegetation indices PRI and CCI were both effective at tracking the seasonal variations of energy partitioning in needles and the differences of carotenoid pigments indicative of the stress status of needles. These results demonstrate that PRI and CCI can be useful tools for monitoring conifer phenology and for the remote monitoring of the length of the photosynthetically active season of conifers in a changing climate.  相似文献   
5.
Rising global temperature and CO2 levels may sustain late-season net photosynthesis of evergreen conifers but could also impair the development of cold hardiness. Our study investigated how elevated temperature, and the combination of elevated temperature with elevated CO2, affected photosynthetic rates, leaf carbohydrates, freezing tolerance, and proteins involved in photosynthesis and cold hardening in Eastern white pine (Pinus strobus). We designed an experiment where control seedlings were acclimated to long photoperiod (day/night 14/10 h), warm temperature (22°C/15°C), and either ambient (400 μL L−1) or elevated (800 μmol mol−1) CO2, and then shifted seedlings to growth conditions with short photoperiod (8/16 h) and low temperature/ambient CO2 (LTAC), elevated temperature/ambient CO2 (ETAC), or elevated temperature/elevated CO2 (ETEC). Exposure to LTAC induced down-regulation of photosynthesis, development of sustained nonphotochemical quenching, accumulation of soluble carbohydrates, expression of a 16-kD dehydrin absent under long photoperiod, and increased freezing tolerance. In ETAC seedlings, photosynthesis was not down-regulated, while accumulation of soluble carbohydrates, dehydrin expression, and freezing tolerance were impaired. ETEC seedlings revealed increased photosynthesis and improved water use efficiency but impaired dehydrin expression and freezing tolerance similar to ETAC seedlings. Sixteen-kilodalton dehydrin expression strongly correlated with increases in freezing tolerance, suggesting its involvement in the development of cold hardiness in P. strobus. Our findings suggest that exposure to elevated temperature and CO2 during autumn can delay down-regulation of photosynthesis and stimulate late-season net photosynthesis in P. strobus seedlings. However, this comes at the cost of impaired freezing tolerance. Elevated temperature and CO2 also impaired freezing tolerance. However, unless the frequency and timing of extreme low-temperature events changes, this is unlikely to increase risk of freezing damage in P. strobus seedlings.Land surface temperature is increasing, particularly in the northern hemisphere (IPCC, 2014), which is dominated by boreal and temperate forests. At higher latitudes, trees rely on temperature and photoperiod cues to detect changing seasons and to trigger cessation of growth and cold hardening during the autumn (Ensminger et al., 2015). For boreal and temperate evergreen conifers, cold hardening involves changes in carbohydrate metabolism, down-regulation of photosynthesis, accumulation of cryoprotective metabolites, and development of freezing tolerance (Crosatti et al., 2013; Ensminger et al., 2015). These processes minimize freezing damage and enable conifers to endure winter stresses. However, rising temperatures result in asynchronous phasing of temperature and photoperiod characterized by delayed arrival of first frosts (McMahon et al., 2010), which may impact the onset and development of cold hardening during autumn.Short photoperiod induces the cessation of growth in many tree species (Downs and Borthwick, 1956; Heide, 1974; Repo et al., 2000; Böhlenius et al., 2006). As a consequence, carbon demand in sink tissue decreases toward the end of the growing season, and the bulk of photoassimilate is translocated from source tissues to storage tissues (Hansen and Beck, 1994; Oleksyn et al., 2000). In addition, cryoprotective soluble sugars, including sucrose, raffinose, and pinitol, accumulate in leaf tissues to enhance freezing tolerance (Strimbeck et al., 2008; Angelcheva et al., 2014). Thus, by winter, leaf nonstructural carbohydrates are mainly comprised of mono- and oligosaccharides, and only minimal levels of starch remain (Hansen and Beck, 1994; Strimbeck et al., 2008). The concurrent decrease of photoassimilate and demand for metabolites that occur during the cessation of growth also impacts the citric acid cycle that mediates between photosynthesis, respiration, and protein synthesis. The citric acid cycle generates NADH to fuel ATP synthesis via mitochondrial electron transport, as well as amino acid precursors (Shi et al., 2015). In C3 plants, the enzyme phosphoenolpyruvate carboxylase (PEPC) converts phosphoenolpyruvate to oxaloacetic acid in order to supplement the flow of metabolites to the citric acid cycle and thus controls the regulation of respiration and photosynthate partitioning (O’Leary et al., 2011).Cessation of growth, low temperature, and presumably short photoperiod decrease the metabolic sink for photoassimilates, resulting in harmful excess light energy (Öquist and Huner, 2003; Ensminger et al., 2006) and increased generation of reactive oxygen species (Adams et al., 2004). During autumn and the development of cold hardiness, conifers reconfigure the photosynthetic apparatus in order to avoid formation of excess light and reactive oxygen species. This involves a decrease in chlorophylls and PSII reaction center core protein D1 (Ottander et al., 1995; Ensminger et al., 2004; Verhoeven et al., 2009), as well as aggregation of light-harvesting complex proteins (Ottander et al., 1995; Busch et al., 2007). Additionally, photoprotective carotenoid pigments accumulate in leaves, especially the xanthophylls, zeaxanthin, and lutein that contribute to nonphotochemical quenching (NPQ) via thermal dissipation of excess light energy (Busch et al., 2007; Verhoeven et al., 2009; Demmig-Adams et al., 2012). Prolonged exposure to low temperature induces sustained nonphotochemical quenching (NPQS), where zeaxanthin constitutively dissipates excess light energy (Ensminger et al., 2004; Demmig-Adams et al., 2012; Fréchette et al., 2015).In conifers, freezing tolerance is initiated during early autumn in response to decreasing photoperiod (Rostad et al., 2006; Chang et al., 2015) and continues to develop through late autumn in response to the combination of short photoperiod and low temperature (Strimbeck and Schaberg, 2009; Chang et al., 2015). In addition to changes in carbohydrate content, freezing tolerance also involves the expression of specific dehydrins (Close, 1997; Kjellsen et al., 2013). Members of the dehydrin protein family are involved in responses to osmotic, salt, and freezing stress (Close, 1996). Dehydrins have been associated with improved freezing tolerance in many species including spinach (Kaye et al., 1998), strawberry (Houde et al., 2004), cucumber (Yin et al., 2006), peach (Wisniewski et al., 1999), birch (Puhakainen et al., 2004), and spruce (Kjellsen et al., 2013). In angiosperms, a characteristic Lys-rich dehydrin motif known as the K-segment interacts with lipids to facilitate membrane binding (Koag et al., 2003; Eriksson et al., 2011). Several in vitro studies have demonstrated dehydrin functions including prevention of aggregation and unfolding of enzymes (using Vitis riparia; Hughes and Graether, 2011), radical scavenging (using Citrus unshiu; Hara et al., 2004), and suppression of ice crystal formation (using Prunus persica; Wisniewski et al., 1999). To date, dehydrin functions have not been demonstrated in planta.Rising temperatures since the mid-twentieth century have delayed the onset of autumn dormancy and increased length of the growing season in forests across the northern hemisphere (Boisvenue and Running, 2006; Piao et al., 2007; McMahon et al., 2010). Studies have shown that elevated temperatures ranging from +4°C to +20°C above ambient can delay down-regulation of photosynthesis in several evergreen conifers. Consistent findings were apparent among climate-controlled chamber studies exposing Pinus strobus seedlings to a sudden shift in temperature and/or photoperiod (Fréchette et al., 2016), as well as chamber studies exposing Picea abies seedlings to simulated autumn conditions using a gradient of decreasing temperature and photoperiod (Stinziano et al., 2015). Similar findings were also demonstrated in open-top chamber experiments exposing mature Pinus sylvestris to a gradient of decreasing temperature and natural photoperiod (Wang, 1996). Elevated temperature (+4°C above ambient) also impaired cold hardening in Pseudotsuga menziesii seedlings (Guak et al., 1998) and mature P. sylvestris (Repo et al., 1996) exposed to a decreasing gradient of temperature and natural photoperiod using open-top chambers. In contrast, a recent study showed that smaller temperature increments (+1.5°C to +3°C) applied using infrared heaters did not delay down-regulation of photosynthesis or impair freezing tolerance in field-grown P. strobus seedlings that were acclimated to larger diurnal and seasonal temperature variations (Chang et al., 2015). For many tree species, photoperiod determines cessation of growth (Tanino et al., 2010; Petterle et al., 2013), length of the growing season (Bauerle et al., 2012), and development of cold hardiness (Welling et al., 1997; Li et al., 2003; Rostad et al., 2006). However, the effects of climate warming on tree phenology are complex and can be unpredictable due to species- and provenance-specific differences in sensitivity to photoperiod and temperature cues (Körner and Basler, 2010; Basler and Körner, 2012; Basler and Körner, 2014).The effect of elevated CO2 further increases uncertainties in the response of trees to warmer climate. Similar to warmer temperature, elevated CO2 may also delay the down-regulation of photosynthesis in evergreens and extend the length of the growing season, as demonstrated in mature P. sylvestris (Wang, 1996). Elevated CO2 increases carbon assimilation (Curtis and Wang, 1998; Ainsworth and Long, 2005) and biomass production (Ainsworth and Long, 2005) during the growing season. The effects could continue during the autumn if dormancy or growth cessation is delayed, which suggests that elevated CO2 may increase annual carbon uptake. However, long-term exposure to elevated CO2 can also down-regulate photosynthesis during the growing season (Ainsworth and Long, 2005). Prior studies that have attempted to determine the impact of a combination of elevated CO2 and/or temperature on cold hardening in evergreens have largely focused on freezing tolerance, with contrasting results. Open-top chamber experiments showed that a combination of elevated temperature and CO2 both delayed and impaired freezing tolerance of P. menziesii seedlings (Guak et al., 1998) and evergreen broadleaf Eucalyptus pauciflora seedlings (Loveys et al., 2006) but did not affect freezing tolerance of mature P. sylvestris (Repo et al., 1996). A recent field experiment examining mature trees revealed that Larix decidua, but not Pinus mugo, exhibited enhanced freezing damage following six years of exposure to combined soil warming and elevated CO2 (Rixen et al., 2012). In contrast, a climate-controlled study showed that exposure to elevated CO2 advanced the date of bud set and improved freezing tolerance in Picea mariana seedlings (Bigras and Bertrand, 2006). In a second study on similar seedlings conducted by the same authors, exposure of trees to elevated CO2 also enhanced freezing tolerance but impaired the accumulation of sucrose and raffinose (Bertrand and Bigras, 2006). These previous experiments used experimental conditions where temperature and photoperiod gradually decreased. While this approach aims to mimic natural conditions, it is difficult to distinguish specific responses to either photoperiod or temperature. Because of the contrasting findings from previous studies, we designed an experiment aiming to separate the effects of photoperiod, temperature, and CO2 on a wide range of parameters that are involved in cold hardening in conifers.Our study aimed to determine (1) how induction and development of the cold hardening process is affected by a shift from long to short photoperiod under warm conditions and (2) how the combination of warm air temperature and elevated CO2 affects photoperiod-induced cold hardening processes in Eastern white pine (P. strobus). To assess the development of cold hardening, we measured photosynthetic rates, changes in leaf carbohydrates, freezing tolerance, and proteins involved in photosynthesis and cold hardening over 36 d. Assuming that both low temperature and short photoperiod cues are required to induce cold hardening in conifers, we hypothesized that warm temperature and the combination of warm temperature and elevated CO2 would prevent seedlings growing under autumn photoperiod from down-regulating photosynthesis. We further hypothesized that warm temperature and the combination of warm temperature and elevated CO2 would impair the development of freezing tolerance, due to a lack of adequate phasing of the low temperature and short photoperiod signals.  相似文献   
6.
Clustered regularly interspaced short palindromic repeats with CRISPR‐associated gene (CRISPR‐Cas) systems are widely recognized as critical genome defense systems that protect microbes from external threats such as bacteriophage infection. Several isolates of the intracellular pathogen Legionella pneumophila possess multiple CRISPR‐Cas systems (type I‐C, type I‐F and type II‐B), yet the targets of these systems remain unknown. With the recent observation that at least one of these systems (II‐B) plays a non‐canonical role in supporting intracellular replication, the possibility remained that these systems are vestigial genome defense systems co‐opted for other purposes. Our data indicate that this is not the case. Using an established plasmid transformation assay, we demonstrate that type I‐C, I‐F and II‐B CRISPR‐Cas provide protection against spacer targets. We observe efficient laboratory acquisition of new spacers under ‘priming’ conditions, in which initially incomplete target elimination leads to the generation of new spacers and ultimate loss of the invasive DNA. Critically, we identify the first known target of L. pneumophila CRISPR‐Cas: a 30 kb episome of unknown function whose interbacterial transfer is guarded against by CRISPR‐Cas. We provide evidence that the element can subvert CRISPR‐Cas by mutating its targeted sequences – but that primed spacer acquisition may limit this mechanism of escape. Rather than generally impinging on bacterial fitness, this element drives a host specialization event – with improved fitness in Acanthamoeba but a reduced ability to replicate in other hosts and conditions. These observations add to a growing body of evidence that host range restriction can serve as an existential threat to L. pneumophila in the wild.  相似文献   
7.
Pathogens deliver complex arsenals of translocated effector proteins to host cells during infection, but the extent to which these proteins are regulated once inside the eukaryotic cell remains poorly defined. Among all bacterial pathogens, Legionella pneumophila maintains the largest known set of translocated substrates, delivering over 300 proteins to the host cell via its Type IVB, Icm/Dot translocation system. Backed by a few notable examples of effector–effector regulation in L. pneumophila, we sought to define the extent of this phenomenon through a systematic analysis of effector–effector functional interaction. We used Saccharomyces cerevisiae, an established proxy for the eukaryotic host, to query > 108,000 pairwise genetic interactions between two compatible expression libraries of ~330 L. pneumophila‐translocated substrates. While capturing all known examples of effector–effector suppression, we identify fourteen novel translocated substrates that suppress the activity of other bacterial effectors and one pair with synergistic activities. In at least nine instances, this regulation is direct—a hallmark of an emerging class of proteins called metaeffectors, or “effectors of effectors”. Through detailed structural and functional analysis, we show that metaeffector activity derives from a diverse range of mechanisms, shapes evolution, and can be used to reveal important aspects of each cognate effector's function. Metaeffectors, along with other, indirect, forms of effector–effector modulation, may be a common feature of many intracellular pathogens—with unrealized potential to inform our understanding of how pathogens regulate their interactions with the host cell.  相似文献   
8.
Linked unresponsiveness operates to induce specific unresponsiveness to fully mismatched vascularized allografts in recipients pretreated with anti-CD4 antibody and syngeneic bone marrow cells expressing a single donor MHC class I alloantigen. The aim of the study was to evaluate early post transplant cytokine expression in allografts where linked unresponsiveness was required for long term graft survival. CBA (H2(k)) mice were pretreated with CBK (H2(k)+K(b)) bone marrow cells under the cover of anti-CD4 antibody 28 days before transplantation of a CBK or a C57BL/10 (H2(b)) cardiac allograft. In both cases graft survival was prolonged (MST=100 days). Intragraft expression for interferon (IFN)-gamma, interleukin (IL)-2, IL-4, IL-10, IL-12(p40), IL-18, iNOS, transforming growth factor (TGF)-beta(1) and C-beta was determined on day 1.5, 3, 7 and 14 after transplantation. Whereas rejecting allografts displayed a sharp peak in IFN-gamma, IL-2, IL-4 and IL-10 expression, non-rejecting allografts were characterized by an initial TGF-beta(1) and IFN-gamma production. An increasing IL-4 expression towards day 14 was a unique feature of linked unresponsiveness. All non-rejecting allografts were characterized by an increasing IL-12(p40) production towards day 14. In summary, the early cytokine expression pattern in allografts after bone marrow induced operational tolerance is influenced by the quantity of donor alloantigens expressed on the graft as well as on the bone marrow inoculum.  相似文献   
9.
10.
Student evaluations of faculty are important components of the medical curriculum and faculty development. To improve the effectiveness and timeliness of student evaluations of faculty in the physiology course, we investigated whether evaluations submitted during the course differed from those submitted after completion of the course. A secure web-based system was developed to collect student evaluations that included numerical rankings (1-5) of faculty performance and a section for comments. The grades that students received in the course were added to the data, which were sorted according to the time of submission of the evaluations and analyzed by Pearson's correlation and Student's t-test. Only 26% of students elected to submit evaluations before completion of the course, and the average faculty ratings of these evaluations were highly correlated [r(14) = 0.91] with the evaluations submitted after completion of the course. Faculty evaluations were also significantly correlated with the previous year [r(14) = 0.88]. Concurrent evaluators provided more comments that were statistically longer and subjectively scored as more "substantive." Students who submitted their evaluations during the course and who included comments had significantly higher final grades in the course. In conclusion, the numeric ratings that faculty received were not influenced by the timing of student evaluations. However, students who submitted early evaluations tended to be more engaged as evidenced by their more substantive comments and their better performance on exams. The consistency of faculty evaluations from year to year and concurrent versus at the end of the course suggest that faculty tend not to make significant adjustments to student evaluations.  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