首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 312 毫秒
1.
The carbon dioxide concentrating system in C4 photosynthesis allows high net photosynthetic rates (P N) at low internal carbon dioxide concentrations (C i), permitting higher P N relative to stomatal conductance (g s) than in C3 plants. This relation would be reflected in the ratio of C i to external ambient (C a) carbon dioxide concentration, which is often given as 0.3 or 0.4 for C4 plants. For a C a of 360 μmol mol−1 that would mean a C i about 110–140 μmol mol−1. Our field observations made near midday on three weedy C4 species, Amaranthus retroflexus, Echinochloa crus-galli, and Setaria faberi, and the C4 crop Sorghum bicolor indicated mean values of C i of 183–212 μ mol mol−1 at C a = 360 μmol mol−1. Measurements in two other C4 crop species grown with three levels of N fertilizer indicated that while midday values of C i at high photon flux were higher at limiting N, even at high nitrogen C i averaged 212 and 196 μmol mol−1 for Amaranthus hypochondriacus and Zea mays, respectively. In these two crops midday C i decreased with increasing leaf to air water vapor pressure difference. Averaged over all measurement days, the mean C i across all C4 species was 198 μmol mol−1, for a C i/C a ratio of 0.55. Prior measurements on four herbaceous C3 species using the same instrument indicated an average C i/C a ratio of 0.69. Hence midday C i values in C 4 species under field conditions may often be considerably higher and more similar to those of C3 species than expected from measurements made on plants in controlled environments. Reducing g s in C4 crops at low water vapor pressure differences could potentially improve their water use efficiency without decreasing P N.  相似文献   

2.
We studied the energy flow from C3 and C4 plants to higher trophic levels in a central Amazonian savanna by comparing the carbon stable-isotope ratios of potential food plants to the isotope ratios of species of different consumer groups. All C4 plants encountered in our study area were grasses and all C3 plants were bushes, shrubs or vines. Differences in δ13C ratios among bushes ( = −30.8, SD = 1.2), vines ( = −30.7, SD = 0.46) and trees ( = −29.7, SD = 1.5) were small. However the mean δ13C ratio of dicotyledonous plants ( = −30.4, SD = 1.3) was much more negative than that of the most common grasses ( = −13.4, SD = 0.27). The insect primary consumers had δ13C ratios which ranged from a mean of −29.5 (SD = 0.47) for the grasshopper Tropidacris collaris to a mean of −14.7 (SD = 0.56) for a termite (Nasutitermes sp.), a range similar to that of the vegetation. However, the common insectivorous and omnivorous vertebrates had intermediate values for δ13C, indicating that carbon from different autotrophic sources mixes rapidly as it moves up the food chain. Despite this mixing, the frogs and lizards generally had higher values of δ13C ( = −21.7, SD = 1.6;  = −21.9, SD = 1.8, respectively) than the birds ( = −24.8, SD = 1.8) and the only species of mammal resident in the savanna ( = −25.4), indicating that they are generally more dependent on, or more able to utilise, food chains based on C4 grasses. Received: 7 May 1998 / Accepted: 30 November 1998  相似文献   

3.
A controlled growth chamber experiment was conducted to investigate the short-term water use and photosynthetic responses of 30-d-old carrot seedlings to the combined effects of CO2 concentration (50–1 050 μmol mol−1) and moisture deficits (−5, −30, −55, and −70 kPa). The photosynthetic response data was fitted to a non-rectangular hyperbola model. The estimated parameters were compared for effects of moisture deficit and elevated CO2 concentration (EC). The carboxylation efficiency (α) increased in response to mild moisture stress (−30 kPa) under EC when compared to the unstressed control. However, moderate (−55 kPa) and extreme (−70 kPa) moisture deficits reduced α under EC. Maximum net photosynthetic rate (P Nmax) did not differ between mild water deficit and unstressed controls under EC. Moderate and extreme moisture deficits reduced P Nmax by nearly 85 % compared to controls. Dark respiration rate (R D) showed no consistent response to moisture deficit. The CO2 compensation concentration (Γ) was 324 μmol mol−1 for −75 kPa and ranged 63–93 μmol mol−1 for other moisture regimes. Interaction between moisture deficit and EC was noticed for P N, ratio of intercellular and ambient CO2 concentration (C i/C a), stomatal conductance (g s ), and transpiration rate (E). P N was maximum and C i/C a was minimum at −30 kPa moisture deficit and at C a of 350 μmol mol−1. The g s and E showed an inverse relationship at all moisture deficit regimes and EC. Water use efficiency (WUE) increased with moisture deficit up to −55 kPa and declined thereafter. EC showed a positive influence towards sustaining P N and increasing WUE only under mild moisture stress, and no beneficial effects of EC were noticed at moderate or extreme moisture deficits.  相似文献   

4.
The metabolic pathway of primary carbon fixation was studied in a peculiar pennate marine diatom, Haslea ostrearia (Bory) Simonsen, which synthesizes and accumulates a blue pigment known as “marennine”. Cells were cultured in a semi-continuous mode under saturating [350 μmol(photon) m−2 s−1] or non-saturating [25 μmol(photon) m−2 s−1] irradiance producing “blue” (BC) and “green” (GC) cells, characterized by high and low marennine accumulation, respectively. Growth, pigment contents (chlorophyll a and marennine), 14C accumulation in the metabolites, and the carbonic anhydrase (CA) activity of the cells were determined during the exponential growth phase. Growth rate and marennine content were closely linked to irradiance during growth: higher irradiance increased both growth rate and marennine content. On the other hand, the Chl a concentration was lower under saturating irradiance. The distribution between the Calvin-Benson (C3) and β-carboxylation (C4) pathways was very different depending on the irradiance during growth. Metabolites of the C3 cycle contained about 70 % of the total fixed radioactivity after 60 s of incorporation into cells cultured under the non-saturating irradiance (GC), but only 47 % under saturating irradiance (BC). At the same time, carbon fixation by β-carboxylation was 24 % in GC versus about 41 % in BC, becoming equal to that in the C3 fixation pathway in the latter. Internal CA activity remained constant, but the periplasmic CA activity was higher under low than high irradiance.  相似文献   

5.
The catalytically competent Mn(II)-loaded form of the argE-encoded N-acetyl-l-ornithine deacetylase from Escherichia coli (ArgE) was characterized by kinetic, thermodynamic, and spectroscopic methods. Maximum N-acetyl-l-ornithine (NAO) hydrolytic activity was observed in the presence of one Mn(II) ion with k cat and K m values of 550 s−1 and 0.8 mM, respectively, providing a catalytic efficiency (k cat/K m) of 6.9 × 105 M−1 s−1. The ArgE dissociation constant (K d) for Mn(II) was determined to be 0.18 μM, correlating well with a value obtained by isothermal titration calorimetry of 0.30 μM for the first metal binding event and 5.3 μM for the second. An Arrhenius plot of the NAO hydrolysis for Mn(II)-loaded ArgE was linear from 15 to 55 °C, suggesting the rate-limiting step does not change as a function of temperature over this range. The activation energy, determined from the slope of this plot, was 50.3 kJ mol−1. Other thermodynamic parameters were ΔG = 58.1 kJ mol−1, ΔH = 47.7 kJ mol−1, and ΔS = –34.5 J mol−1 K−1. Similarly, plots of lnK m versus 1/T were linear, suggesting substrate binding is controlled by a single step. The natural product, [(2S,3R)-3-amino-2-hydroxy-4-phenylbutanoyl]leucine (bestatin), was found to be a competitive inhibitor of ArgE with a K i value of 67 μM. Electron paramagnetic resonance (EPR) data recorded for both [Mn(II)_(ArgE)] and [Mn(II)Mn(II)(ArgE)] indicate that the two Mn(II) ions form a dinuclear site. Moreover, the EPR spectrum of [Mn(II)Mn(II)(ArgE)] in the presence of bestatin indicates that bestatin binds to ArgE but does not form a μ-alkoxide bridge between the two metal ions.  相似文献   

6.
We determined the interactive effects of irradiance, elevated CO2 concentration (EC), and temperature in carrot (Daucus carota var. sativus). Plants of the cv. Red Core Chantenay (RCC) were grown in a controlled environmental plant growth room and exposed to 3 levels of photosynthetically active radiation (PAR) (400, 800, 1 200 μmol m−2 s−1), 3 leaf chamber temperatures (15, 20, 30 °C), and 2 external CO2 concentrations (C a), AC and EC (350 and 750 μmol mol−1, respectively). Rates of net photosynthesis (P N) and transpiration (E) and stomatal conductance (g s ) were measured, along with water use efficiency (WUE) and ratio of internal and external CO2 concentrations (C i/C a). P N revealed an interactive effect between PAR and C a. As PAR increased so did P N under both C a regimes. The g s showed no interactive effects between the three parameters but had singular effects of temperature and PAR. E was strongly influenced by the combination of PAR and temperature. WUE was interactively affected by all three parameters. Maximum WUE occurred at 15 °C and 1 200 μmol m−2 s− 1 PAR under EC. The C i /C a was influenced independently by temperature and C a. Hence photosynthetic responses are interactively affected by changes in irradiance, external CO2 concentration, and temperature. EC significantly compensates the inhibitory effects of high temperature and irradiance on P N and WUE.  相似文献   

7.
为了解巴西橡胶树(Hevea brasiliensis)栽培种质的变异情况,以53份在云南植胶区综合性状表现较好的巴西橡胶树栽培种质为材料,采用流式细胞术测定了基因组C值,并进行了变异分析。结果表明,浅绿色嫩叶是巴西橡胶树流式细胞术测定的最适样品。53份巴西橡胶树栽培种质的细胞核DNA含量和基因组C值存在一定差异,基因组的平均C值是1.531 696×109 bp,最小的是CRTG-272种质(1.465 908×10~9 bp),最大的是CRTG-83种质(1.600 381×10~9 bp),变异系数较小(CV=0.035 5)。53份巴西橡胶树栽培种质中有47份为二倍体,6份为三倍体。在已测定基因组大小的40种大戟科(Euphorbiaceae)植物中,基因组大小变异较大(CV=1.248 6),与"C值悖论"观点相一致。因此,应用流式细胞术能快速、准确地测定巴西橡胶树细胞核DNA含量、基因组C值和染色体倍性。  相似文献   

8.
A nucleopolyhedrovirus (MaviMNPV) was isolated from diseased larvae of legume pod borer (LPB), Maruca vitrata, at Tainan in Taiwan. Electron microscopical studies on the ultrastructure of MaviMNPV occlusion bodies (OBs) showed several virions (up to 19) with multiple nucleocapsids (up to 6) packaged within a single viral envelope. The diameter of OBs was 0.9 to 1.3 μm with a mean of 1.152±0.116 μm. The complete sequence of the MaviMNPV polyhedrin (Polh) gene contained 735 nucleotides (GenBank accession number DQ399596). Phylogenetic analyses using the complete sequence of the Polh gene of MaviMNPV indicated that this virus clusters with Group I NPVs. The genome size of MaviMNPV estimated with restriction enzymes viz., HindIII, EcoRI, BglII and PstI was 113.41 ± 1.50 kbp. First instar LPB larvae were the most susceptible stage (LC50 2.053 × 102 OBs/ml) followed by second, third and fourth instars with the median lethal concentrations (LC50s) 1.410 × 103, 2.390 × 103 and 2.636 × 103 OBs/ml, respectively. This is the first record of this virus from this region. The first and second authors have equal contributions in this paper  相似文献   

9.
The bioelectrochemistry of the blue copper protein, pseudoazurin, at glassy carbon and platinum electrodes that were modified with single-wall carbon nanotubes (SWNTs) was investigated by multiple scan rate cyclic voltammetry. The protein showed reversible electrochemical behavior at both bare glassy carbon electrodes (GCEs) and SWNT-modified GCEs (SWNT|GCEs); however, direct electrochemistry was not observed at any of the platinum electrodes. The effect of the carbon nanotubes at the GCE was to amplify the current response 1000-fold (nA at bare GCE to μA at SWNT|GCE), increase the apparent diffusion coefficient D app of the solution-borne protein by three orders of magnitude, from 1.35 × 10−11 at bare GCE to 7.06 × 10−8 cm2 s-1 at SWNT|GCE, and increase the heterogeneous electron transfer rate constant k s threefold, from 1.7 × 10−2 cm s−1 at bare GCE to 5.3 × 10−2 cm s−1 at SWNT|GCE. Pseudoazurin was also found to spontaneously adsorb onto the nanotube-modified GCE surface. Well-resolved voltammograms indicating quasi-reversible faradaic responses were obtained for the adsorbed protein in phosphate buffer, with I pc and I pa values now greater than corresponding values for solution-borne pseudoazurin at SWNT|GCEs and with significantly reduced ΔE p values. The largest electron transfer rate constant of 1.7 × 10−1 cm s−1 was achieved with adsorbed pseudoazurin at the SWNT|GCE surface in deaerated buffer solution consistent with its presumed role in anaerobic respiration of some bacteria.  相似文献   

10.
Summary Short-term culture of rainbow trout (Onchorhynchus mykiss) hepatocytes was used to examine the effect of dexamethasone (DEX) on microsomal CYP 1A1 protein content and 7-ethoxyresorufin-O-deethylase (EROD) activity in vitro. Hepatocytes prepared by controlled collagenase digestion and plated at a density of 0.25 × 106 cells/cm2 in plastic culture dishes precoated with trout skin extract (7.6 μg skin protein/cm2) to facilitate cell attachment were maintained at 16° C. Cells were treated with DEX (10−9 to 10−7 M) or vehicle (dimethyl sulfoxide, DMSO) at 24 h. Microsomal CYP 1A1 protein content and EROD activities were measured at 72 h. Both CYP 1A1 protein as measured by Western blots using CYP 1A1 specific anti-sera and EROD activity were significantly lower in DEX (10−8 to 10−7 M)-treated hepatocytes compared to untreated (control) or DMSO-treated cells. The effect was dose dependent in that a gradual decrease of CYP 1A1 protein and EROD activities were seen with increasing doses of DEX (10−8 to 10−7 M). DEX at 10−9 M was ineffective. Concomitant addition of 10−6 M RU486, a type II specific glucocorticoid receptor antagonist, to hepatocytes treated with 10−7 M DEX abolished the DEX effect. RU486 at 10−8 M was ineffective. Spironolactone (10−8 to 10−6 M), a type I specific glucocorticoid receptor antagonist, did not counteract the DEX effect. RU486 or spironolactone (10−6 M) alone had no effect on CYP 1A1 under similar conditions. DEX thus down regulates CYP 1A1 in fish cultured hepatocytes and this regulation is mediated through the type II glucocorticoid receptor(s).  相似文献   

11.
Genome size has been studied for the first time in the Colchicum genus. Values obtained by flow cytometry were quite stable and specific to each taxon: C. autumnale L., 2C=5.89±0.22 pg, equivalent to 5.7×10 bp (2 n=4 x=36); C. alpinum DC., 2C=8.06±0.24 pg, equivalent to 7.8×10 bp (2 n=6 x=56); C. lusitanum Brot., 2C=10.7±0.67 pg, equivalent to 10.3×10 bp (2 n= approx. 10 x=90–92 and 94–96); C. multiflorum Brot., 2C=16.5±0.69 pg, equivalent to 15.9×10 bp, (2 n= approx. 16 x=140–148); C. corsicum Baker, 2C=21.3±0.99 pg, equivalent to 20.6×10 bp (2 n=22 x= approx. 198±2). These values are well below those published for Liliaceae stricto sensu. In Colchicum species from the western Mediterranean area, genome size was highly correlated with ploidy level ( R 2=0.99, P<0.001). This relationship is consistent with most previous results, with our new chromosome number count for C. corsicum and with our correction of published erroneous counts for C. lusitanum (2 n=102–108). The Calabrian population appeared to be distinct from all of the other plants of the C. alpinum group. Reproducible and accurate, cytometry appears to be a particularly appropriate method for studying this polyploid genus in Western Europe, taking into account that chromosome numbers are difficult to enumerate. It can guide future taxonomic research because it reveals similarities or differences between taxa within this difficult complex.  相似文献   

12.
The genome size and base composition of diploid plant species from three genera of the Casuarinaceae family were determined by flow cytometry. Casuarina glauca Sieb. ex Spring. and Gymnostoma deplancheana (Miq.) L. Johnson showed a small genome with 2C = 0.70 pg, 58.6% AT, 40.5% GC for the first species and 2C = 0.75 pg, 58.7% AT, 40.5% GC for the second. Allocasuarina verticillata (Lam.) L. Johnson had a larger genome: 2C = 1.90 pg, 59.3% AT, 41.1% GC. One haploid genome of C. glauca is therefore about 340×106 base pairs. In leaves, roots or bark of these three species, polysomaty was virtually absent: a maximum frequency of 4C nuclei of only 0.08 was found in bark of C. glauca. The genome sizes of C. glauca and G. deplancheana are among the smallest described for higher plants. Small genome size, diploidy and the absence of polysomaty are advantageous traits for facilitating molecular approaches to improvement of these actinorhizal plants and developing the study of their symbiotic interactions with Frankia. Received: 20 December 1997 / Revision received: 13 March 1998 / Accepted: 30 March 1998  相似文献   

13.
Flow cytometric analysis performed on two different crosses of dura×pisifera oil palm gave an accurate estimation of nuclear DNA content. The genome size of Elaeis guineensis was found to be 2C=3.76±0.09 pg and therefore ca. 3.4×109 bp. Embryogenic calli and plants showed the same ploidy level, but the measured 2C DNA values differed significantly. No variation in the ploidy level between three different types of calli originating from foliar explants, namely nodular compact callus, fast-growing callus and friable callus was observed. Since fast-growing callus (FGC), already identified as a source of `mantled' phenotype variants, did not show any difference in their ploidy level, these results are consistent with the hypothesis of an epigenetic origin for this type of somaclonal variant. Received: 17 February 1997 / Revision received: 13 May 1997 / Accepted: 22 May 1997  相似文献   

14.
Nuclear genome size of conifers as measured by flow cytometry with propidium iodide was investigated, striving to collect at least a single species from each genus. 64 out of 67 genera and 172 species were measured. Of the 67 genera, 21 are reported here for the first time and the same is true for 76 species. This nearly doubles the number of measured genera and adds 50% to the number of analyzed species. Conifers have chromosome numbers in the range of n = (7)10–12(19). However, the nuclear DNA content (2C‐value) is shown here to range from 8.3 to 71.6 picogram. The largest genome contains roughly 6 × 1010 more base pairs than the smallest genome. Genome sizes are evaluated and compared with available taxonomic treatments. For the mainly (sub)tropical Podocarpaceae small genome sizes were found with a 2C‐value of only 8–28 pg, with 13.5 pg on average. For the Taxaceae 2C‐values from 23–60 pg were determined. Not surprisingly, the genus Pinus with 97 species (39 species measured here) has a broad range with 2C = 38–72 pg. A factor of 2 difference is also found in the Cupressaceae (136 species) with nuclear DNA contents in the range 18–35 pg. Apart from the allohexaploid Sequoia, ploidy plays a role only in Juniperus and some new polyploids are found. The data on genome size support conclusions on phylogenetic relationships obtained by DNA sequencing. Flow cytometry is applicable even to young plants or seeds for the monitoring of trade in endangered species.  相似文献   

15.
 A correlation between genome size and agronomically important traits has been observed in many plant species. The goal of the present research was to determine the relationship between genome size, seed size, and leaf width and length in soybean [Glycine max (L.) Merr.] Twelve soybean strains, representing three distinct seed size groups, were analyzed. Flow cytometry was used to estimate their 2C nuclear DNA contents. Data on seed size and leaf size of the 12 strains were obtained from 1994 and 1995 field experiments. Variation of 2C nuclear DNA among the 12 soybean strains was 4.6%, ranging from 2.37 pg for a small-seed strain to 2.48 pg for a large-seed strain. Strain seed size was positively associated with leaf width (r=0.92) and leaf length (r=0.93). Genome size was highly correlated with seed size (r=0.97), leaf width (r=0.90) , and leaf length (r=0.93). The results of our study indicate that there is a significant correlation between genome size and leaf and seed size in soybean. It is possible that selection for greater seed size either leads to, or results from, greater genome size. If so, this relationship might be worth exploring at a more fundamental level. Received: 5 April 1997 / Accepted: 9 January 1998  相似文献   

16.
毛竹基因组大小测定   总被引:1,自引:0,他引:1  
李潞滨  武静宇    胡陶  杨学文  彭镇华 《植物学报》2008,25(5):574-578
毛竹(Phyllostachys edulis)属禾本科(Poaceae)竹亚科(Bambusoideae)刚竹属(Phyllostachys), 是我国分布和栽培面积最大的经济竹种, 有着广泛的开发前景。本实验以水稻(Oryza sativa)为内标, 用流式细胞仪对水稻和竹子样品的PI发射荧光强度进行测定, 通过比较水稻与毛竹样品峰值的倍数关系, 计算出毛竹的基因组大小。对24组样品进行重复测试, 测得毛竹基因组大小为 2 075.025±13.08 Mb, 即2 C DNA含量为4.24 pg(以1 pg DNA = 0.978×109 bp计算)。毛竹基因组大小测定为毛竹基因组文库的建立及其基因组学研究奠定了重要基础。  相似文献   

17.
Nuclear DNA content (2C) is used as a new criterion to investigate nearly all species of the genus Nerine Herb. The species have the same chromosome number (2n = 2x = 22), with the exception of three triploid plants found. The nuclear DNA content of the diploids, as measured by flow cytometry with propidium iodide, is demonstrated to range from 18.0–35.3 pg. This implies that the largest genome contains roughly 2 × 1010 more base pairs than the smallest. The species, arranged according to increasing genome size, fell apart in three groups if growth cycle and leaf width were also considered. A narrow-leafed, evergreen group with a DNA content between 18.0 and 24.6 pg contains thirteen species, a broad-leaved winter growing group with four species has a DNA content from 25.3–26.2 pg and a broad-leafed summer growing group has a DNA content of 26.8–35.3 pg and contains six species. If the presence of filament appendages and hairiness of the pedicels were also considered, the thirteen evergreen species could be further divided into a group without filament appendages or hairy pedicels with a DNA content of 18.0–18.7 pg. A second group without filament appendages but with hairy pedicels had a DNA content of 19.7–22.3 pg. And a third group with both filament appendages and hairy pedicels had a DNA content of 22.0–24.6 pg. The exception is N. marincowitzii that, despite a low DNA content and narrow leaves is summer growing. The broad-leafed group is further characterised by the absence of filament appendages and the absence of strongly hairy pedicels. The exception here is N. pusilla that, despite a high DNA content, has narrow leaves and minutely hairy pedicels. Nuclear DNA content as measured by flow cytometry is shown to be relevant to throw new light on the relationships between Nerine species.  相似文献   

18.
Limited Genome Size Variation in Sesleria albicans   总被引:2,自引:1,他引:1  
The extent and significance of intraspecific genome size variationin plants continues to be a matter of discussion: in some speciesconsiderable variation has been described, while no variationhas been detected in other taxa. In the present study, intraspecificgenome size variation was analysed in a perennial grass Sesleriaalbicans Kit. ex Schult. (Poaceae). Flow cytometry was usedfor the analysis of nuclear DNA content in ten geographicallyisolated populations ofS. albicans . Despite long-term isolationand lack of gene-flow between the populations, only negligibleinter-population differences were found. Although the differencesbetween the populations were statistically significant, themaximum inter-population difference reached only 1.6% of themean 2C value (9.78 ± 0.04 pg). The variation was notcorrelated with geographical location or with altitude of thepopulations analysed. The present study clearly demonstratesthat S. albicans belongs to the plant taxa with a highly stablegenome size. Copyright 2000 Annals of Botany Company Sesleria albicans, genome size, nuclear DNA content, intraspecific variation, flow cytometry, Europe  相似文献   

19.
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