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 31 毫秒
1.
In this study, interactions between Cry1Ac, a toxic crystal protein produced by Bacillus thuringiensis (Berliner), and Beauveria bassiana on the mortality and survival of Ostrinia furnacalis was evaluated in the laboratory. The results showed that Cry1Ac is toxic to O. furnacalis. Not only were larval growth and development delayed, but pupation, pupal weight and adult emergency also decreased when larvae were fed on artificial diet containing purified Cry1Ac toxin. When third instars O. furnacalis were exposed to combination of B. bassiana (1.8 × 105, 1.8 × 106 or 1.8 × 107 conidia ml−1) and Cry1Ac, (0.2 or 0.8 μg g−1), the effect on mortality was additive, however, the combinations of sublethal concentrations showed antagonism between Cry1Ac (3.2 or 13 μg g−1) and B. bassiana (1.8 × 105 or 1.8 × 106 conidia ml−1). When neonates were reared on sublethal concentrations of Cry1AC until the third instar, and survivors exposed B. bassiana conidial suspension, such treatments showed additive effect on mortality of O. furnacalis except for the combination of Cry1Ac (0.2 μg g−1) and B. bassiana (1.8 × 106 conidia ml−1) that showed antagonism.  相似文献   

2.
In this study, interactions on the mortality and debilitating effects between Cry1Ac, a toxic protein produced by Bacillus thuringiensis (Berliner) and HaCPV (Chinese strain) on first and third instars larvae of Helicoverpa armigera were evaluated in laboratory. When first instar was exposed to combination of Bt cotton leaf discs containing HaCPV (6 × 106, 1 × 107, and 3 × 107 PIB ml−1) the effect on mortality was additive, when such instar larvae exposed to combination of Cry1Ac (0.9, 2.7, or 8.1 μg g−1) and the same concentrations of HaCPV the effect on mortality was additive except for the combination of Cry1Ac (0.3 μg g−1) and HaCPV concentrations that showed synergism. When third instars of H. armigera were infected using a suspension containing both HaCPV and Cry1Ac, most combinations of them showed additive effect except for the combination of Cry1Ac (0.3 μg g−1) and HaCPV (3 × 107 PIB ml−1) that showed synergism. However, when they exposed to Bt cotton leaf discs and HaCPV the effect on mortality was synergism except combination of Bt cotton leaf discs and HaCPV (6 × 106 PIB ml−1) that showed additive. Most of the combinations are showed additive effect in the toxicity and in combinations of Cry1Ac at lowest and HaCPV at highest concentrations synergism is observed. Not only were larval growth and development delayed, but pupation and pupal weight also decreased when larvae were fed on artificial diet containing Cry1Ac and HaCPV or transgenic Bt cotton leaf discs specially in first instar.  相似文献   

3.
Two 1-aminocyclopropane-1-carboxylic acid (ACC) oxidase (ACO) genes have been cloned from RNA isolated from leaf tissue of apple (Malus domestica cv. Royal Gala). The genes, designated MD-ACO2 (with an ORF of 990 bp) and MD-ACO3 (966 bp) have been compared with a previously cloned gene of apple, MD-ACO1 (with an ORF of 942 bp). MD-ACO1 and MD-ACO2 share a close nucleotide sequence identity of 93.9% in the ORF but diverge in the 3′ untranslated regions (3′-UTR) (69.5%). In contrast, MD-ACO3 shares a lower sequence identity with both MD-ACO1 (78.5%) and MD-ACO2 (77.8%) in the ORF, and 68.4% (MD-ACO1) and 71% (MD-ACO2) in the 3′-UTR. Southern analysis confirmed that MD-ACO3 is encoded by a distinct gene, but the distinction between MD-ACO1 and MD-ACO2 is not as definitive. Gene expression analysis has shown that MD-ACO1 is restricted to fruit tissues, with optimal expression in ripening fruit, MD-ACO2 expression occurs more predominantly in younger fruit tissue, with some expression in young leaf tissue, while MD-ACO3 is expressed predominantly in young and mature leaf tissue, with less expression in young fruit tissue and least expression in ripening fruit. Protein accumulation studies using western analysis with specific antibodies raised to recombinant MD-ACO1 and MD-ACO3 produced in E. coli confirmed the accumulation of MD-ACO1 in mature fruit, and an absence of accumulation in leaf tissue. In contrast, MD-ACO3 accumulation occurred in younger leaf tissue, and in younger fruit tissue. Further, the expression of MD-ACO3 and accumulation of MD-ACO3 in leaf tissue is linked to fruit longevity. Analysis of the kinetic properties of the three apple ACOs using recombinant enzymes produced in E. coli revealed apparent Michaelis constants (Km) of 89.39 μM (MD-ACO1), 401.03 μM (MD-ACO2) and 244.5 μM (MD-ACO3) for the substrate ACC, catalytic constants (Kcat) of 6.6 × 10−2 (MD-ACO1), 3.44 × 10−2 (Md-ACO2) and 9.14 × 10−2 (MD-ACO3) and Kcat/Km (μM s−1) values of 7.38 × 10−4 μM s−1 (MD-ACO1), 0.86 × 10−4 M s−1 (MD-ACO2) and 3.8 × 10−4 μM s−1 (MD-ACO3). These results show that MD-ACO1, MD-ACO2 and MD-ACO3 are differentially expressed in apple fruit and leaf tissue, an expression pattern that is supported by some variation in kinetic properties.  相似文献   

4.
The uptake kinetics of phosphate (Pi) by Myriophyllum spicatum was determined from adsorption and absorption under light and dark conditions. Pi uptake was light dependent and showed saturation following the Michaelis-Menten relation (in light: V = 16.91 × [Pi](1.335 + [Pi]), R2 = 0.90, p < 0.001; in the dark: V = 5.13 × [Pi](0.351 + [Pi]), R2 = 0.77, p < 0.001). Around 77% of the loss of Pi in the water column was absorbed into the tissue of M. spicatum, and only 23% was adsorbed on the surface of the plant shoots. Our study shows that M. spicatum shoots have a much higher affinity (in light: 3.9 μmol g−1 dw h−1 μM−1; in the dark: 3.7 μmol g−1 dw h−1 μM−1) and Vmax (maximum uptake rate, shoot light) for Pi uptake than many other aquatic macrophytes (in light: 0.002-0.23 μmol g−1 dw h−1 μM−1; in the dark: 0.002-0.19 μmol g−1 dw h−1 μM−1), which may provide a competitive advantage over other macrophytes across a wide range of Pi concentrations.  相似文献   

5.
Human arylamine N-acetyltransferase 1 (NAT1) is a xenobiotic-metabolizing enzyme that biotransforms aromatic amine chemicals. We show here that biologically-relevant concentrations of inorganic (Hg2+) and organic (CH3Hg+) mercury inhibit the biotransformation functions of NAT1. Both compounds react irreversibly with the active-site cysteine of NAT1 (half-maximal inhibitory concentration (IC50) = 250 nM and kinact = 1.4 × 104 M−1 s−1 for Hg2+ and IC50 = 1.4 μM and kinact = 2 × 102 M−1 s−1 for CH3Hg+). Exposure of lung epithelial cells led to the inhibition of cellular NAT1 (IC50 = 3 and 20 μM for Hg2+ and CH3Hg+, respectively). Our data suggest that exposure to mercury may affect the biotransformation of aromatic amines by NAT1.  相似文献   

6.
Blue multicopper oxidases, laccases displayed on the surface of Bacillus spores were used to decolorize a widely used textile dyestuff, indigo carmine. The laccase-encoding gene of Bacillus subtilis, cotA, was cloned and expressed in B. subtilis DB104, and the expressed enzyme was spontaneously localized on Bacillus spores. B. subtilis spores expressing laccase exhibited maximal activity for the oxidation of 2,2′-azino-bis (3-ethylthiazoline-6-sulfonate) (ABTS) at pH 4.0 and 80 °C, and for the decolorization of indigo carmine at pH 8.0 and 60 °C. The displayed enzyme retained 80% of its original activity after pre-treatment with organic solvents such as 50% acetonitrile and n-hexane for 2 h at 37 °C. The apparent Km of the enzyme displayed on spores was 443 ± 124 μM for ABTS with a Vmax of 150 ± 16 U/mg spores. Notably, 1 mg of spores displaying B. subtilis laccase (3.4 × 102 U for ABTS as a substrate) decolorized 44.6 μg indigo carmine in 2 h. The spore reactor (0.5 g of spores corresponding to 1.7 × 105 U in 50 mL) in a consecutive batch recycling mode decolorized 223 mg indigo carmine/L to completion within 42 h at pH 8.0 and 60 °C. These results suggest that laccase displayed on B. subtilis spores can serve as a powerful environmental tool for the treatment of textile dye effluent.  相似文献   

7.
Techniques utilizing β-glucuronidase (GUS) activity as an indicator of Escherichia coli (E. coli) presence use labeled glucuronides to produce optical signals. Carboxyumbelliferyl-β-d-glucuronide (CUGlcU) is a fluorescent labeled glucuronide that is soluble and highly fluorescent at natural water pHs and temperatures and, therefore, may be an ideal reagent for use in an in situ optical sensor. This paper reports for the first time the Michaelis-Menten kinetic parameters for the binding of E. coli GUS with CUGlcU as Km = 910 μM, Vmax = 41.0 μM min−1, Vmax/Km 45.0 μmol L−1 min−1, the optimal pH as 6.5 ± 1.0, optimal temperature as 38 °C, and the Gibb's free energy of activation as 61.40 kJ mol−1. Additionally, it was found CUGlcU hydrolysis is not significantly affected by heavy solvents suggesting proton transfer and solvent addition that occur during hydrolysis are not limiting steps. Comparison studies were made with the more common fluorescent molecule methylumbelliferyl-β-d-glucuronide (MUGlcU). Experiments showed GUS preferentially binds to MUGlcU in comparison to CUGlcU. CUGlcU was also demonstrated in a prototype optical sensor for the detection of E. coli. Initial bench testing of the sensor produced detection of low concentrations of E. coli (1.00 × 103 CFU/100 mL) in 230 ± 15.1 min and high concentrations (1.05 × 105 CFU/100 mL) in 8.00 ± 1.01 min.  相似文献   

8.
Our study aimed to test the ability of aquatic plants to use bicarbonate when acclimated to three different bicarbonate concentrations. To this end, we performed experiments with the three species Ceratophyllum demersum, Egeria densa, Lagarosiphon major to determine photosynthetic rates under varying bicarbonate concentrations. We measured bicarbonate use efficiency, photosynthetic performance and respiration. For all species, our results revealed that photosynthetic rates were highest in replicates grown at low alkalinity. Thus, E. densa had approx. five times higher rates at low (264 ± 15 μmol O2 g−1 DW h−1) than at high alkalinity (50 ± 27 μmol O2 g−1 DW h−1), C. demersum had three times higher rates (336 ± 95 and 120 ± 31 μmol O2 g−1 DW h−1), and L. major doubled its rates at low alkalinity (634 ± 114 and 322 ± 119 μmol O2 g−1 DW h−1). Similar results were obtained for bicarbonate use efficiency by E. densa (136 ± 44 and 43 ± 10 μmol O2 mequiv. L−1 g−1 DW h−1) and L. major (244 ± 29 and 82 ± 24 μmol O2 mequiv. L−1 g−1 DW h−1). As to C. demersum, efficiency was high but unaffected by alkalinity, indicating high adaptation ability to varied alkalinities. A pH drift experiment supported these results. Overall, our results suggest that the three globally widespread worldwide species of our study adapt to low inorganic carbon availability by increasing their efficiency of bicarbonate use.  相似文献   

9.
An efficient and user-friendly bacterial transformation method by simple spreading cells with aminoclays was demonstrated. Compared to the reported transformation approaches using DNA adsorption or wrapping onto (in)organic fibers, the spontaneously generated clay-coated DNA suprastructures by mixing DNA with aminoclay resulted in transformants in both Gram-negative (Escherichia coli) and Gram-positive cells (Streptococcus mutans). Notably, the wild type S. mutans showed comparable transformation efficiency to that of the E. coli host for recombinant DNA cloning. This is a potentially promising result because other trials such as heat-shock, electroporation, and treatment with sepiolite for introducing DNA into the wild type S. mutans failed. Under defined conditions, the transformation efficiency of E. coli XL1-Blue and S. mutans exhibited ~ 2 × 105 and ~ 6 × 103 CFU/μg of plasmid DNA using magnesium-aminoclay. In contrast, transformation efficiency was higher in S. mutans than that in E. coli XL1-Blue for calcium-aminoclay. It was also confirmed that each plasmid transformed into E. coli and S. mutans was stably maintained and that they expressed the inserted gene encoding the green fluorescent protein during prolonged growth of up to 80 generations.  相似文献   

10.
Lead accumulation by free and immobilized cyanobacteria, Lyngbya majuscula and Spirulina subsalsa was studied. Exponentially growing biomass was exposed to 1-20 mg L−1 of Pb(II) solution at pH 6, 7 and 8 for time periods ranging from 10 min to 48 h. L. majuscula accumulated 10 times more Pb (13.5 mg g−1) than S. subsalsa (1.32 mg g−1) at pH 6 within 3 h of exposure to 20 mg L−1 Pb(II) solution and 76% of the Pb could be recovered using 0.1 M EDTA. This chelator (2 μM) did not influence Pb accumulation whereas 100 μM citrate increased that of S. subsalsa 6- to 8-fold. L. majuscula filaments enmeshed in a glass wool packed in a column removed 95.8% of the Pb from a 5 mg L−1 Pb solution compared to free and dead biomass which removed 64 and 33.6% Pb respectively. A 92.5% recovery of accumulated Pb from the immobilized biomass suggests that repeated absorption-desorption is possible.  相似文献   

11.
The ethanolic extract from Flourensia oolepis aerial parts showed strong antifeedant activity against the pest larvae, Epilachna paenulata, with an antifeedant index (AI%) of 99.1% at 100 μg/cm2. Based on chromatographic fractionation of the extract, guided by bioassays on E. paenulata, the flavanone pinocembrin (1) was isolated as the most active principle. In a choice assay, compound 1 showed strong antifeedant activity against E. paenulata, Xanthogaleruca luteola and Spodoptera frugiperda with an AI% of 90, 94 and 91% (p < 0.01) respectively, at 50 μg/cm2. The dosages necessary for 50% feeding inhibition of the insects (ED50) were 7.98, 6.13 and 8.86 μg/cm2, respectively. The feeding inhibitory activity of 1 against E. paenulata was compared with the activity of other structurally related flavonoids like naringenin, which was inactive up to 100 μg/cm2, catechin which was nearly 6 times less active than 1, and quercetin which was equally active as 1. The effect of these on the feeding behavior of E. paenulata was also studied.  相似文献   

12.
Yeast cell wall fractions have been proposed to bind enteropathogenic bacteria. The aim of this study was to develop a quantitative assay by measuring the optical density as growth parameter of adhering bacteria. The exponential growth phase of adhering bacteria was determined by optical density reading and compared with the colony count (CFU/mL). A linear regression was compiled and the bacterial number bound to the yeast cell wall product could be determined. Further focus was the investigation of a yeast cell wall from strain Trichosporon mycotoxinivorans (MTV) for its ability to bind gram negative Salmonella, E. coli and Campylobacter strains and gram positive probiotic bacteria of the genera lactobacilli and bifidobacteria as well as gram positive Clostridium perfringens quantitatively. The gram negative probiotic strain E. coli Nissle 1917 was also investigated. Seven out of 10 S. Typhimurium and S. Enteritidis strains adhered to the cell wall product with an amount between 103 and 104 CFU/10 μg. Four out of 7 E. coli strains showed an average binding capability (102 CFU/10 µg) whereas 4 × 103E. coli F4 cells bound per 10 μg yeast cell wall. E. coli 0149 K91, E. coli 0147 K89, C. jejuni and C. perfringens as well the genera lactobacilli and bifidobacteria did not bind to the yeast cell wall. E. coli Nissle 1917 was bound with 2 × 102 CFU/10 μg. These results demonstrate that cell wall from MTV can be used to differentially bind E. coli spp. and Salmonella spp. up to 8 × 104 CFU/10 μg. Thus certain yeast cell walls may prevent enteric infections caused by selective bacteria. This methodical approach would be an accurate tool in the feed industry for quality control of yeast cell wall products.  相似文献   

13.
14.
Microbial contamination on surfaces of food processing equipment is a major concern in industries. A new method to inoculate a single-cell layer (monolayer) of microorganisms onto polystyrene was developed, using a deposition with an airbrush. A homogeneous dispersion of Bacillus subtilis DSM 402 spores sprayed on the surface was observed using both plate count and scanning electron microscopy. No clusters were found, even with high spore concentrations (107 spores/inoculated surface). A monolayer of microorganisms was also obtained after deposition of 10 μL droplets containing 3 × 104 spores/spot on polystyrene disks, but not with a higher spore concentration. Pulsed light (PL) applied to monolayers of B. subtilis spores allowed log reductions higher than 6. As a consequence of clusters formation in spots of 10 μL containing more than 3 × 105 spores, log reductions obtained by PL were significantly lower. The comparative advantages of spot and spray depositions were discussed.  相似文献   

15.
In order for cryopreservation to become a practical tool for aquaculture, optimized protocols must be developed for each species and cell type. Knowledge of a cell’s osmotic tolerance and membrane permeability characteristics can assist in optimized protocol development. In this study, these characteristics were determined for Pacific oyster oocytes and modified methods for loading and unloading ethylene glycol (EG) were tested. Oocytes were found to behave as ideal osmometers and their osmotically inactive fraction (Vb) was calculated to be 0.48. Oocytes exposed to NaCl solutions of 0.6 to 2.3 Osm fertilized at rates equivalent to oocytes left in seawater. This corresponds to volume changes of +27.3 and −38.1 ± 1.2%. The permeability of the oocytes to water (Lp) was determined to be 3.8 ± 0.4 × 10−2, 5.7 ± 0.8 × 10−2, and 13.2 ± 1.3 × 10−2 μm min−1 atm−1, when measured at temperatures of 5, 10 and 20 °C. The respective EG permeability values (Ps) were 9.5 ± 0.1 × 10−5, 14.6 ± 1.2 × 10−5, and 41.7 ± 2.4 × 10−5 cm min−1. The activation energies for Lp and Ps were determined to be 14.5 and 17.5 kcal mol−1, respectively. Different models for EG loading and unloading from oocytes were developed and tested. Post-thaw fertilization did not differ significantly between a published step addition method and single step addition at 20 °C. This represents a considerable reduction in handling. The results of this study demonstrate that the cryobiological characteristics of a given cell type should be taken into account when developing cryopreservation methods.  相似文献   

16.
A microchip electrophoresis (MCE) method with chemiluminescence (CL) detection was developed for the determination of carnosine-related peptides, including carnosine, homocarnosine, and anserine, in biological samples. A simple integrated MCE-CL system was built to perform the assays. The highly sensitive CL detection was achieved by means of the CL reaction between hydrogen peroxide and N-(4-aminobutyl)-N-ethylisoluminol-tagged peptides in the presence of adenine as a CL enhancer and Co2+ as a catalyst. Experimental conditions for analyte labeling, MCE separation, and CL detection were studied. MCE separation of the above-mentioned three peptides took less than 120 s. Detection limits (signal/noise ratio [S/N] = 3) of 3.0 × 10−8, 2.8 × 10−8, and 3.4 × 10−8 M were obtained for carnosine, anserine, and homocarnosine, respectively. The current MCE-CL method was applied for the determination of carnosine, anserine, and homocarnosine in human cerebrospinal fluid (CSF) and canine plasma. Homocarnosine was detected at the micromolar (μM) level in the CSF samples analyzed, whereas the levels of carnosine and anserine in these samples were below the detection limit of the assay. Interestingly, both carnosine and anserine were detected in the canine plasma samples, whereas homocarnosine was not.  相似文献   

17.
GOX is the most widely used enzyme for the development of electrochemical glucose biosensors and biofuel cell in physiological conditions. The present work describes the production of a recombinant glucose oxidase from Penicillium amagasakiense (yGOXpenag) displaying a more efficient glucose catalysis (kcat/KM(glucose) = 93 μM−1 s−1) than the native GOX from Aspergillus niger (nGOXaspng), which is the most industrially used (kcat/KM(glucose) = 27 μM−1 s−1). Expression in Pichia pastoris allowed easy production and purification of the recombinant active enzyme, without overglycosylation. Its biotechnological interest was further evaluated by measuring kinetics of ferrocinium-methanol (FMox) reduction, which is commonly used for electron transfer to the electrode surface. Despite their homologies in sequence and structure, pH-dependant FMox reduction was different between the two enzymes. At physiological pH and temperature, we observed that electron transfer to the redox mediator is also more efficient for yGOXpenag than for nGOXaspng(kcat/KM(FMox) = 27 μM−1 s−1 and 17 μM−1 s−1 respectively). In our model system, the catalytic current observed in the presence of blood glucose concentration (5 mM) was two times higher with yGOXpenag than with nGOXaspng. All our results indicated that yGOXpenag is a better candidate for industrial development of efficient bioelectrochemical devices used in physiological conditions.  相似文献   

18.
An efficient method of regeneration for antidiabetic plant (Stevia rebaudiana) has been established for healthy biomass and main steviol glycosides (SGs) production, using different PGRs and agar concentrations. Higher callus induction (93.3%) was recorded when leaf explants were placed on an MS medium supplemented with 3.5 gL−1 agar and 2.0 mgL−1 2,4-D. The addition of 7.0 gL−1 agar and BA (1.0, 2.0 and 4.0 mgL−1) significantly (< 0.01) influences shooting response (100%). A maximum mean shoot length (13.03 cm) and 28 shoots per explant were observed on a medium containing 1.0 mgL−1 BA. However, the maximum number of leaves (132.67) was encouraged by the addition of BA (1.0 mgL−1) and Kin (1.0 mgL−1). Lower agar (3.5 gL−1), IAA (2.0 mgL−1), and NAA (2.0 mgL−1) concentrations significantly influence the rooting percent (100%), the mean root length (2.9 cm), and the number of roots per plantlet (26.3). These plantlets were successfully acclimatized in the soil. The BA (3.0 mgL−1) in combination with Kin (3.0 mgL−1) and 3.5 gL−1 agar increases dulcoside-A content (Dul-A; 71.8 μg/g-DW) in shoots compared to control (50.81 μg/g-DW). Similar PGRs with 7.0 gL−1 significantly increases the production of steviosides (Stev. 82.48 μg/g-DW). A higher rebaudioside-A content (Reb-A; 12.35 μg/g-DW) was observed in shoots that underwent the addition of BA (1.0 mgL−1) and 7.0 gL−1 agar than in control (07.39 μg/g-DW). Hereby, we developed an efficient and cost-effective method for regeneration and major SGs production, which could be helpful for future studies on this species.  相似文献   

19.
A granulosis virus (GV) was isolated from the diseased caterpillars of Arctornis submarginata (Walker) (Lymantriidae), a defoliating pest of tea from Darjeeling foothill region. The phase contrast and transmission electron microscopic studies identified the virus as granulosis virus. SDS-PAGE analysis of major protein of the occlusion bodies was found to be 31 kDa, characteristic for granulin. The total genomic DNA was isolated. The major band found was of molecular weight 16 kDa. Bioassay conducted with the occlusion bodies (OBs) of the virus showed LC50 value of 4.46 × 104 OBs/ml for the second instar caterpillars. Median lethal time (LT50) were 6.6 days for 1 × 10OBs/ml, 5.09 days for 1 × 105 OBs/ml, 4.45 days for 1 × 106 OBs/ml and 3.87 days for 1 × 107OBs/ml concentrations. The results indicated the potential of the virus for its future application as microbial pesticide against A. submarginata in future.  相似文献   

20.
Three Algerian populations of female Pistacia atlantica shrubs were investigated in order to check whether their terpenoid contents and morpho-anatomical parameters may characterize the infraspecific variability. The populations were sampled along a gradient of increasing aridity from the Atlas mountains into the northwestern Central Sahara.As evidenced by Scanning Electron Microscopy, tufted hairs could be found only on seedling leaves from the low aridity site as a population-specific trait preserved also in culture. Under common garden cultivation seedlings of the high aridity site showed a three times higher density of glandular trichomes compared to the low aridity site. Increased aridity resulted also in reduction of leaf sizes while their thickness increased. Palisade parenchyma thickness also increases with aridity, being the best variable that discriminates the three populations of P. atlantica.Analysis of terpenoids from the leaves carried out by GC-MS reveals the presence of 65 compounds. The major compounds identified were spathulenol (23 μg g−1 dw), α-pinene (10 μg g−1 dw), verbenone (7 μg g−1 dw) and β-pinene (6 μg g−1 dw) in leaves from the low aridity site; spathulenol (73 μg g−1 dw), α-pinene (25 μg g−1 dw), β-pinene (18 μg g−1 dw) and γ-amorphene (16 μg g−1 dw) in those from medium aridity and spathulenol (114 μg g−1 dw), α-pinene (49 μg g−1 dw), germacrene D (29 μg g−1 dw) and camphene (23 μg g−1 dw) in leaves from the high aridity site. Terpene concentrations increased with the degree of aridity: the highest mean concentration of monoterpenes (136 μg g−1 dw), sesquiterpenes (290 μg g−1 dw) and total terpenes (427 μg g−1 dw) were observed in the highest arid site and are, respectively, 3-, 5- and 4-fold higher compared to the lower arid site. Spathulenol and α-pinene can be taken as chemical markers of aridity. Drought discriminating compounds in low, but detectable concentrations are δ-cadinene and β-copaene. The functional roles of the terpenoids found in P. atlantica leaves and principles of their biosynthesis are discussed with emphasis on the mechanisms of plant resistance to drought conditions.  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