首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The flowers of 23 species of grass and herb plants were collected from a mesotrophic grassland to assess natural variability in bulk, monosaccharide and fatty acid δ13C values from one plant community and were compared with previous analyses of leaves from the same species. The total mean bulk δ13C value of flower tissues was −28.1‰, and there was no significant difference between the mean δ13Cflower values for grass (−27.8‰) and herb (−28.2‰) species. On average bulk δ13Cflower values were 1.1‰ higher than bulk δ13Cleaf values, however, the δ13Cflower and δ13Cleaf values of grasses did not differ between organs suggesting that carbon isotope discrimination is different in grass and herb species. The abundance of different monosaccharides abundance varied between plant types, i.e. xylose concentrations in the grass flowers were as high as 40%, compared with up to 15% in the herb species, but the general relationship δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose which had been observed in leaves was similar in flowers (total mean δ13C values = −25.9‰, −27.2‰, −28.8‰ and −28.1‰, respectively). However, the average 5.4‰ depletion in the δ13C values of the C16:0, C18:2 and C18:3 fatty acids in flowers compared to bulk tissue was significantly greater than observed for leaves. The trend C16:0 < C18:2 < C18:3 previously observed in leaves was also observed in grass flowers (δ13CC16:0 = −33.8‰; δ13CC18:2 = −33.1‰; δ13CC18:3 = −34.2‰) but not herb flowers (δ13CC16:0 = −34.1‰; δ13CC18:2 = −32.4‰; δ13CC18:3 = −34.5‰). We conclude: (i) that the biological processes influencing carbon isotope discrimination in grass flowers are different from herbs flowers; and, (ii) that a range of post-photosynthetic fractionation effects caused the observed differences between flower and leaf δ13C values, especially the significant 13C-depletion in flower fatty acid δ13C values.  相似文献   

2.
Resuspension cultures of Gibberella fujikuroi, strain GF-1a, were shown to metabolise potassium [3′-13C] mevalonate to 13C-enriched C19-gibberellins, plus 13CO2 (derived from the loss of carbon-20). The formation of [13C]-gibberellins could be observed in vivo using 13C NMR; however that of 13CO2 could not. In contrast, removal of the mycelium and concentration of the filtrate at pH 12 enabled the 13CO2 produced to be observed using 13C NMR. During incubations of H14CO2Na with this fungus, complete conversion to other radioactive products was observed, and the significance of these results in the light of previous work is discussed.  相似文献   

3.
Energy expenditure (EE) can be estimated based on respiratory gas exchange measurements, traditionally done in respiration chambers by indirect calorimetry (IC). However, the 13C-bicarbonate technique (13C-BT) might be an alternative minimal invasive method for estimation of CO2 production and EE in the field. In this study, four Shetland ponies were used to explore the effect of feeding on CO2 production and EE measured simultaneously by IC and 13C-BT. The ponies were individually housed in respiration chambers and received either a single oral or intravenous (IV) bolus dose of 13C-labelled sodium bicarbonate (NaH13CO3). The ponies were fed haylage 3 h before (T−3), simultaneously with (T0) or 3 h after (T+3) administration of 13C-bicarbonate. The CO2 produced and O2 consumed by the ponies were measured for 6 h with both administration routes of 13C-bicarbonate at the three different feeding times. Feeding time affected the CO2 production (P<0.001) and O2 consumption (P<0.001), but not the respiratory quotient (RQ) measured by IC. The recovery factor (RF) of 13C in breath CO2 was affected by feeding time (P<0.01) and three different RF were used in the calculation of CO2 production measured by 13C-BT. An average RQ was used for the calculations of EE. There was no difference between IC and 13C-BT for estimation of CO2 production. An effect of feeding time (P<0.001) on the estimated EE was found, with higher EE when feed was offered (T0 and T+3) compared with when no feed was available (T−3) during measurements. In conclusion, this study showed that feeding time affects the RF and measurements of CO2 production and EE. This should be considered when the 13C-BT is used in the field. IV administration of 13C-bicarbonate is recommended in future studies with horses to avoid complex 13C enrichment-time curves with maxima and shoulders as observed in several experiments with oral administration of 13C-bicarbonate.  相似文献   

4.
《Marine Micropaleontology》2006,58(4):243-258
We sampled the upper water column for living planktic foraminifera along the SW-African continental margin. The species Globorotalia inflata strongly dominates the foraminiferal assemblages with an overall relative abundance of 70–90%. The shell δ18O and δ13C values of G. inflata were measured and compared to the predicted oxygen isotope equilibrium values (δ18Oeq) and to the carbon isotope composition of the total dissolved inorganic carbon (δ13CDIC) of seawater. The δ18O of G. inflata reflects the general gradient observed in the predicted δ18Oeq profile, while the δ13C of G. inflata shows almost no variation with depth and the reflection of the δ13CDIC in the foraminiferal shell seems to be covered by other effects. We found that offsets between δ18Oshell and predicted δ18Oeq in the surface mixed layer do not correlate to changes in seawater [CO32−].To calculate an isotopic mass balance of depth integrated growth, we used the oxygen isotope composition of G. inflata to estimate the fraction of the total shell mass that is grown within each plankton tow depth interval of the upper 500 m of the water column. This approach allows us to calculate the Δδ13Cinterval added-DIC; i.e. the isotopic composition of calcite that was grown within a given depth interval. Our results consistently show that the Δδ13CIA-DIC correlates negatively with in situ measured [CO32−] of the ambient water. Using this approach, we found Δδ13CIA-DIC/[CO32−] slopes for G. inflata in the large size fraction (250–355 μm) of − 0.013‰ to 0.015‰ (μmol kg 1) 1 and of − 0.013‰ to 0.017‰ (μmol kg 1) 1 for the smaller specimens (150–250 μm). These slopes are in the range of those found for other non-symbiotic species, such as Globigerina bulloides, from laboratory culture experiments. Since the Δδ13CIA-DIC/[CO32−] slopes from our field data are nearly identical to the slopes established from laboratory culture experiments we assume that the influence of other effects, such as temperature, are negligibly small. If we correct the δ13C values of G. inflata for a carbonate ion effect, the δ13Cshell and δ13CDIC are correlated with an average offset of 2.11.  相似文献   

5.
Coenzyme Q (Q or ubiquinone) is a redox-active polyisoprenylated benzoquinone lipid essential for electron and proton transport in the mitochondrial respiratory chain. The aromatic ring 4-hydroxybenzoic acid (4HB) is commonly depicted as the sole aromatic ring precursor in Q biosynthesis despite the recent finding that para-aminobenzoic acid (pABA) also serves as a ring precursor in Saccharomyces cerevisiae Q biosynthesis. In this study, we employed aromatic 13C6-ring-labeled compounds including 13C6-4HB, 13C6-pABA, 13C6-resveratrol, and 13C6-coumarate to investigate the role of these small molecules as aromatic ring precursors in Q biosynthesis in Escherichia coli, S. cerevisiae, and human and mouse cells. In contrast to S. cerevisiae, neither E. coli nor the mammalian cells tested were able to form 13C6-Q when cultured in the presence of 13C6-pABA. However, E. coli cells treated with 13C6-pABA generated 13C6-ring-labeled forms of 3-octaprenyl-4-aminobenzoic acid, 2-octaprenyl-aniline, and 3-octaprenyl-2-aminophenol, suggesting UbiA, UbiD, UbiX, and UbiI are capable of using pABA or pABA-derived intermediates as substrates. E. coli, S. cerevisiae, and human and mouse cells cultured in the presence of 13C6-resveratrol or 13C6-coumarate were able to synthesize 13C6-Q. Future evaluation of the physiological and pharmacological responses to dietary polyphenols should consider their metabolism to Q.  相似文献   

6.
Leaves of 26 grass, herb, shrub and tree species were collected from mesotrophic grasslands to assess natural variability in bulk, fatty acid and monosaccharide δ13C values under different grazing management (cattle- or deer-grazed) on three sample dates (May, July and October) such that interspecific and spatiotemporal variations in whole leaf tissues and compound-specific δ13C values could be determined. The total mean leaf bulk δ13C value for plants was −28.9‰ with a range of values spanning 7.5‰. Significant interspecific variation between bulk leaf δ13C values was only determined in October (P = <0.001) when δ13C values of the leaf tissues from both sites was on average 1.5‰ depleted compared to during July and May. Samples from May were significantly different between fields (P = 0.03) indicating an effect from deer- or cattle-grazing in young leaves. The average individual monosaccharide δ13C value was 0.8‰ higher compared with whole leaf tissues. Monosaccharides were the most abundant components of leaf biomass, i.e. arabinose, xylose, mannose, galactose and glucose, and therefore, fluctuations in their individual δ13C values had a major influence on bulk δ13C values. An average depletion of ca. 1‰ in the bulk δ13C values of leaves from the deer-grazed field compared to the cattle-grazed field could be explained by a general depletion of 1.1‰ in glucose δ13C values, as glucose constituted >50% total leaf monosaccharides. In October, δ13C values of all monosaccharides varied between species, with significant variation in δ13C values of mannose and glucose in July, and mannose in May. This provided an explanation for the noted variability in the tissue bulk δ13C values observed in October 1999. The fatty acids C16:0, C18:2 and C18:3 were highly abundant in all plant species. Fatty acid δ13C values were lower than those of bulk leaf tissues; average values of −37.4‰ (C16:0), −37.0‰ (C18:2) and −36.5‰ (C18:3) were determined. There was significant interspecific variation in the δ13C values of all individual fatty acids during October and July, but only for C18:2 in May (P = <0.05). This indicated that seasonal trends observed in the δ13C values of individual fatty acids were inherited from the isotopic composition of primary photosynthate. However, although wide diversity in δ13C values of grassland plants ascribed to grazing management, interspecific and spatiotemporal influences was revealed, significant trends (P = <0.0001) for fatty acid and monosaccharide δ13C values: δ13C16:0 < δ13C18:2 < δ13C18:3 and δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose, respectively, previously described, appear consistent across a wide range of species at different times of the year in fields under different grazing regimes.  相似文献   

7.
The major radioactive products of the fixation of [13N]N2 by Azolla caroliniana Willd.-Anabaena azollae Stras. were ammonium, glutamine, and glutamate, plus a small amount of alanine. Ammonium accounted for 70 and 32% of the total radioactivity recovered after fixation for 1 and 10 minutes, respectively. The presence of a substantial pool of [13N]N2-derived 13NH4+ after longer incubation periods was attributed to the spatial separation between the site of N2-fixation (Anabaena) and a second, major site of assimilation (Azolla). Initially, glutamine was the most highly radioactive organic product formed from [13N]N2, but after 10 minutes of fixation glutamate had 1.5 times more radiolabel than glutamine. These kinetics of radiolabeling, along with the effects of inhibitors of glutamine synthetase and glutamate synthase on assimilation of exogenous and [13N]N2-derived 13NH4+, indicate that ammonium assimilation occurred by the glutamate synthase cycle and that glutamate dehydrogenase played little or no role in the synthesis of glutamate by Azolla-Anabaena.  相似文献   

8.
Our goal was to develop a field soil biodegradation assay using 13C-labeled compounds and identify the active microorganisms by analyzing 16S rRNA genes in soil-derived 13C-labeled DNA. Our biodegradation approach sought to minimize microbiological artifacts caused by physical and/or nutritional disturbance of soil associated with sampling and laboratory incubation. The new field-based assay involved the release of 13C-labeled compounds (glucose, phenol, caffeine, and naphthalene) to soil plots, installation of open-bottom glass chambers that covered the soil, and analysis of samples of headspace gases for 13CO2 respiration by gas chromatography/mass spectrometry (GC/MS). We verified that the GC/MS procedure was capable of assessing respiration of the four substrates added (50 ppm) to 5 g of soil in sealed laboratory incubations. Next, we determined background levels of 13CO2 emitted from naturally occurring soil organic matter to chambers inserted into our field soil test plots. We found that the conservative tracer, SF6, that was injected into the headspace rapidly diffused out of the soil chamber and thus would be of little value for computing the efficiency of retaining respired 13CO2. Field respiration assays using all four compounds were completed. Background respiration from soil organic matter interfered with the documentation of in situ respiration of the slowly metabolized (caffeine) and sparingly soluble (naphthalene) compounds. Nonetheless, transient peaks of 13CO2 released in excess of background were found in glucose- and phenol-treated soil within 8 h. Cesium-chloride separation of 13C-labeled soil DNA was followed by PCR amplification and sequencing of 16S rRNA genes from microbial populations involved with 13C-substrate metabolism. A total of 29 full sequences revealed that active populations included relatives of Arthrobacter, Pseudomonas, Acinetobacter, Massilia, Flavobacterium, and Pedobacter spp. for glucose; Pseudomonas, Pantoea, Acinetobacter, Enterobacter, Stenotrophomonas, and Alcaligenes spp. for phenol; Pseudomonas, Acinetobacter, and Variovorax spp. for naphthalene; and Acinetobacter, Enterobacter, Stenotrophomonas, and Pantoea spp. for caffeine.  相似文献   

9.
Local dynamics of interhelical loops in bacteriorhodopsin (bR), the extracellular BC, DE and FG, and cytoplasmic AB and CD loops, and helix B were determined on the basis of a variety of relaxation parameters for the resolved 13C and 15N signals of [1-13C]Tyr-, [15N]Pro- and [1-13C]Val-, [15N]Pro-labeled bR. Rotational echo double resonance (REDOR) filter experiments were used to assign [1-13C]Val-, [15N]Pro signals to the specific residues in bR. The previous assignments of [1-13C]Val-labeled peaks, 172.9 or 171.1 ppm, to Val69 were revised: the assignment of peak, 172.1 ppm, to Val69 was made in view of the additional information of conformation-dependent 15N chemical shifts of Pro bonded to Val in the presence of 13C-15N correlation, although no assignment of peak is feasible for 13C nuclei not bonded to Pro. 13C or 15N spin-lattice relaxation times (T1), spin-spin relaxation times under the condition of CP-MAS (T2), and cross relaxation times (TCH and TNH) for 13C and 15N nuclei and carbon or nitrogen-resolved, 1H spin-lattice relaxation times in the rotating flame (1H T) for the assigned signals were measured in [1-13C]Val-, [15N]Pro-bR. It turned out that V69-P70 in the BC loop in the extracellular side has a rigid β-sheet in spite of longer loop and possesses large amplitude motions as revealed from 13C and 15N conformation-dependent chemical shifts and T1, T2, 1H T and cross relaxation times. In addition, breakage of the β-sheet structure in the BC loop was seen in bacterio-opsin (bO) in the absence of retinal.  相似文献   

10.
The helix-coil transition of poly-l-lysine hydrochloride ((Lys)n) in aqueous solution has been studied by 13C Fourier-transform nuclear magnetic resonance spectroscopy. As reference compounds dodeca-l-lysine hydrobromide ((Lys)?12, tri-l-lysine hydrochloride ((Lys)3), and l-lysine hydrochloride (Lys), have been also studied by the same method. It is found that 13C spin-lattice relaxation times t1 of the carbonyl and the side-chain carbons decrease sharply at pD 10.2 which is the midpoint of the transition from the random-coil to the α-helix. Similarly the T1 values of the carbonyl groups of (Lys)?12 decrease at this point in a more moderate way, while no change is observed for those of the side-chain carbons. This is interpreted in terms of the reduced α-helicity involved for (Lys)?12.The variation of 13C chemical shifts with pD for (Lys)n and (Lys)?12 show the same trend:downfield shifts at higher pD. Furthermore, nonterminal and C-terminal residues of (Lys)3 show similar behavior. Thus it is concluded that the 13C chemical shift changes are caused mainly by the pD changes and not by the conformational transition. Conversion from α-helix to β-structure by elevation of temperature at pD 11.2 results in narrowing and downfield shifts of the 13C resonances of (Lys)n.  相似文献   

11.
《Inorganica chimica acta》1988,145(2):273-277
The 13C and 15SN NMR spectra of eleven cis-Fe(S2CNRR′)2(CO)2 complexes, where R and R′ are organic substituents, have been measured at ambient temperature in CDCl3 (0.08–0.16 M). The 13C absorptions for the carbonyl ligands correlate well with the force constants for the CO stretching vibrations in CHCl3 solution. Each of the parameters (13CO absorption and kcis for CO) correlate well with the aqueous solution pKa for+H2NRR′, corrected for the phenyl-containing substituents, high pKa values corresponding to high 13CO absorptions and low kcis CO force constants. [p ]Evidence was found in the 13C NMR spectra for hindered rotation about the CN bond in S2CNC2 in complexes with higher pKa(corr) values and in the 13C spectra of the corresponding thiuram disulfides. [p ]The 15N (natural abundance) NMR spectra for each of the complexes was determined. Each revealed a single sharp absorption in a region of the 15N NMR spectrum which indicates substantial CN double bond character, as one would expect for coordinated dithiocarbamate ligands.  相似文献   

12.
The biosynthesis of the iridoid glucoside lamalbid in Lamium barbatum, a plant species in the Lamiaceae, was investigated by administrating 13C-labeled intermediates of MVA and MEP pathways, respectively. The results demonstrated that [3,4,5-13C3]1-deoxy-d-xylulose 5-phosphate could be incorporated into lamalbid, whereas the incorporation of [2-13C1]mevalonolactone was not observed. Based on the 13C labeling pattern of lamalbid and the incorporation data, we deduce that the iridoid glucoside in L. barbatum is biosynthesized through the MEP pathway, whereas the classic MVA pathway is not utilized.  相似文献   

13.
We describe the use of carbon stable isotope (13C) labeled n-alkanes as a potential internal tracer to assess passage kinetics of ingested nutrients in ruminants. Plant cuticular n-alkanes originating from intrinsically 13C labeled ryegrass plants were pulse dosed intraruminally in four rumen-cannulated lactating dairy cows receiving four contrasting ryegrass silage treatments that differed in nitrogen fertilization level (45 or 90 kg nitrogen ha−1) and maturity (early or late). Passage kinetics through the gastrointestinal tract were derived from the δ13C (i.e. the ratio 13C:12C) in apparently undigested fecal material. Isotopic enrichment was observed in a wide range of long-chain n-alkanes (C27–C36) and passage kinetics were determined for the most abundant C29, C31 and C33 n-alkanes, for which a sufficiently high response signal was detected by combustion isotope ratio mass spectrometry. Basal diet treatment and carbon chain length of n-alkanes did not affect fractional passage rates from the rumen (K 1) among individual n-alkanes (3.71–3.95%/h). Peak concentration time and transit time showed a quantitatively small, significant (p≤0.002) increase with carbon chain length. K 1 estimates were comparable to those of the 13C labeled digestible dry matter fraction (3.38%/h; r = 0.61 to 0.71; p≤0.012). A literature review has shown that n-alkanes are not fermented by microorganisms in the rumen and affirms no preferential depletion of 13C versus 12C. Our results suggest that 13C labeled n-alkanes can be used as nutrient passage tracers and support the reliability of the δ13C signature of digestible feed nutrients as a tool to measure nutrient-specific passage kinetics.  相似文献   

14.
Coenzyme Q (ubiquinone or Q) is a crucial mitochondrial lipid required for respiratory electron transport in eukaryotes. 4-Hydroxybenozoate (4HB) is an aromatic ring precursor that forms the benzoquinone ring of Q and is used extensively to examine Q biosynthesis. However, the direct precursor compounds and enzymatic steps for synthesis of 4HB in yeast are unknown. Here we show that para-aminobenzoic acid (pABA), a well known precursor of folate, also functions as a precursor for Q biosynthesis. A hexaprenylated form of pABA (prenyl-pABA) is normally present in wild-type yeast crude lipid extracts but is absent in yeast abz1 mutants starved for pABA. A stable 13C6-isotope of pABA (p- amino[aromatic-13C6]benzoic acid ([13C6]pABA)), is prenylated in either wild-type or abz1 mutant yeast to form prenyl-[13C6]pABA. We demonstrate by HPLC and mass spectrometry that yeast incubated with either [13C6]pABA or [13C6]4HB generate both 13C6-demethoxy-Q (DMQ), a late stage Q biosynthetic intermediate, as well as the final product 13C6-coenzyme Q. Pulse-labeling analyses show that formation of prenyl-pABA occurs within minutes and precedes the synthesis of Q. Yeast utilizing pABA as a ring precursor produce another nitrogen containing intermediate, 4-imino-DMQ6. This intermediate is produced in small quantities in wild-type yeast cultured in standard media and in abz1 mutants supplemented with pABA. We suggest a mechanism where Schiff base-mediated deimination forms DMQ6 quinone, thereby eliminating the nitrogen contributed by pABA. This scheme results in the convergence of the 4HB and pABA pathways in eukaryotic Q biosynthesis and has implications regarding the action of pABA-based antifolates.  相似文献   

15.
Fractional passage rates form a fundamental element within modern feed evaluation systems for ruminants, but knowledge on feed-specific fractional passage is largely lacking. Commonly applied tracer techniques based on externally applied markers, such as chromium-mordanted neutral detergent fibre (Cr-NDF), have been criticised for behaving differently to feed particles. This study describes the use of the carbon stable isotope ratio (13C : 12C) as an internal digesta marker to quantify the fractional passage rate of concentrates through the digestive tract of dairy cows. In a crossover study, five dairy cows were fed low (24.6%) and high (52.6%) levels of concentrates (dry matter (DM) basis) and received a pulse-dosed Cr-NDF and 13C isotopes. The latter was administered orally by exchanging part of the dietary concentrates of low 13C natural abundance with a pulse dose of maize bran-based concentrates of high 13C natural abundance. Fractional passage rates from the rumen (K1) and from the large intestine (K2) were determined from faecal marker concentrations of Cr-NDF and of 13C in the DM (13C-DM), NDF (13C-NDF) and neutral detergent soluble (13C-NDS). No differences in K1 estimates were found for the two concentrate levels fed but significant differences between markers (P<0.001) were observed. Faecal Cr-NDF excretions gave lower K1 estimates (0.037–0.039/h) than 13C-DM (0.054–0.056/h) and 13C-NDF (0.061–0.063/h). The 13C-NDS was calculated by the difference of 13C in the DM and NDF, and K1 values (0.039–0.043/h) were comparable to Cr-NDF. Total mean retention time was considerably higher for Cr-NDF (40.9–42.0 h) as compared to 13C-DM and 13C-NDF (32.0–33.5 h; P<0.001). The accuracy of the curve fits for Cr-NDF and 13C-DM and 13C-NDF was overall good (mean prediction error of 9.9–13.9%). Fractional passage rate of Cr-NDF was comparable to studies where this marker was assumed to represent the fractional passage of roughages. However, K1 estimates based on the 13C : 12C ratio varied considerably from studies based on external markers. Our results suggest that the use of 13C isotopes as digesta passage markers can provide feed component-specific K1 estimates for concentrates and provides new insight into passage kinetics of NDF from technologically treated compound feed.  相似文献   

16.
《FEBS letters》1987,216(1):4-6
The competitive oxidation of 13CH3OH and 13CD3OH has been observed using in vivo 13C NMR spectroscopy. Simultaneous 1H and 2H decoupling gave isotopically shifted 13C singlets for the two methanol isotopomers. The measured enzymic isotope effect, kH/kD is approx. 1.8, indicating that CH bond cleavage is not rate-determining.  相似文献   

17.
Biao Zhu  Weixin Cheng 《Plant and Soil》2011,342(1-2):277-287
Stable carbon isotopes are used extensively to partition total soil CO2 efflux into root-derived rhizosphere respiration or autotrophic respiration and soil-derived heterotrophic respiration. However, it remains unclear whether CO2 from rhizosphere respiration has the same δ13C value as root biomass. Here we investigated the magnitude of 13C isotope fractionation during rhizosphere respiration relative to root biomass in six plant species. Plants were grown in a carbon-free sand-perlite medium inoculated with microorganisms from a farm soil for 62 days inside a greenhouse. We measured the δ13C value of rhizosphere respiration using a closed-circulation 48-hour CO2 trapping method during 40~42 and 60~62 days after sowing. We found a consistent depletion in 13C (0.9~1.7‰) of CO2 from rhizosphere respiration relative to root biomass in three C3 species (Glycine max L. Merr., Helianthus annuus L. and Triticum aestivum L.), but a relatively large depletion in 13C (3.7~7.0‰) in three C4 species (Amaranthus tricolor L., Sorghum bicolor (L.) Moench and Zea mays L. ssp. mays). Overall, our results indicate that CO2 from rhizosphere respiration is more 13C-depleted than root biomass. Therefore, accounting for this 13C fractionation is required for accurately partitioning total soil CO2 efflux into root-derived and soil-derived components using natural abundance stable carbon isotope methods.  相似文献   

18.
In plants, jasmonic acid (JA) and its derivatives are thought to be involved in mobile forms of defense against biotic and abiotic stresses. In this study, the distal transport of JA-isoleucine (JA-Ile) that is synthesized de novo in response to leaf wounding in tomato (Solanum lycopersicum) plants was investigated. JA-[13C6]Ile was recovered in distal untreated leaves after wounded leaves were treated with [13C6]Ile. However, as [13C6]Ile was also recovered in the distal untreated leaves, whether JA-Ile was synthesized in the wounded or in the untreated leaves was unclear. Hence, stem exudates were analyzed to obtain more detailed information. When [13C6]Ile and [2H6]JA were applied separately into the wounds on two different leaves, JA-[13C6]Ile and [2H6]JA-Ile were detected in the stem exudates but [2H6]JA–[13C6]Ile was not, indicating that JA was conjugated with Ile in the wounded leaf and that the resulting JA-Ile was then transported into systemic tissues. The [2H3]JA-Ile that was applied exogenously to the wounded tissues reached distal untreated leaves within 10 min. Additionally, applying [2H3]JA-Ile to the wounded leaves at concentrations of 10 and 60 nmol/two leaves induced the accumulation of PIN II, LAP A, and JAZ3 mRNA in the distal untreated leaves of the spr2 mutant S. lycopersicum plants. These results demonstrate the transportation of de novo synthesized JA-Ile and suggest that JA-Ile may be a mobile signal.  相似文献   

19.

Background and Aims

Carbon isotope discrimination (Δ13C) in C3 plants used as an indirect measure of water-use efficiency (WUE) provides a tool for selecting crops with high WUE under dry environments.

Methods

We evaluated the physiology and Δ13C of a set of 8 F5 recombinant inbred lines (RILs) with contrasting levels of leaf Δ13C derived from two parents, ‘W89001002003’ (low Δ13C) and ‘I60049’ (high Δ13C) of six-row barley (Hordeum vulgare L.) in a greenhouse and under field conditions in three locations (Lacombe, Vegreville and Castor). In the greenhouse experiment, seven days of water deficit was imposed at the stem elongation stage followed by re-watering to pre-deficit level.

Results

A significant negative relationship between WUE and leaf Δ13C was observed. Under water-deficit conditions, both photosynthetic rate (A) and stomatal conductance (g s ) were significantly reduced with a strong positive correlation (r = 0.89) between the two, and the variation in g s was proportionally greater than A. The low leaf-Δ13C RIL ‘147’ maintained the highest A and g s among ten genotypes (RILs and parents) under water-deficit conditions. Leaf Δ13C was positively correlated with biomass and grain yield in the field trials. Multivariate analysis of leaf Δ13C, harvest index and plant height discriminated genotypes into three clusters: drought sensitive, drought tolerant and an intermediate type.

Conclusions

The study suggests that it is possible to select low Δ13C lines such as RIL ‘147’, which is able to maintain or produce high yields under low moisture conditions on the Canadian Prairies  相似文献   

20.
We have shown that the yeast-Escherichia coli shuttle vector YEp 13 contains, as part of its yeast chromosomal segment, a tRNA3Leu gene. We have also isolated and characterized a variant of YEp13, namely YEp13-a, which is capable of suppressing a variety of yeast amber-suppressible alleles in vivo. YEp13-a differs from YEp13 by a single point mutation, which changes the three-nucleotide, plus-strand sequence corresponding to the tRNA3Leu anticodon from the normal C-A-A to C-T-A. This nucleotide change creates a site for the restriction enzyme XbaI in the suppressor tRNA3Leu gene. We have taken advantage of the correlation between the suppressor mutation and the XbaI site formation, to show that the tRNA3Leu gene on YEp13 corresponds to the genetically characterized yeast chromosomal amber suppressor SUP53. We have also shown that SUP53 is located just centromere-distal to LEU2 on chromosome III. Finally, comparison of the DNA sequence of SUP53 and its flanking regions with the sequences of other cloned yeast tRNA3Leu genes has revealed considerable sequence homology in the immediate 5′-flanking regions of these genes.  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