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 109 毫秒
1.
The epicuticular wax covering on plant surface plays important roles in protecting plants against UV radiation. However, the role of epicuticular wax in affecting leaf gas exchange under enhanced ultraviolet-B (UV-B) radiation remains obscure. In the present study, different aged leaves of Brassica napus were used to analyze the responses of crystal structure and chemical constituents of epicuticular wax to UV-B radiation and the effects of such responses on gas exchange indices. Enhanced UV-B radiation significantly decreased the amount of esters in all leaves except the first leaf, amount of secondary alcohols in the second, third and fourth leaves, and amount of primary alcohols in the second and third leaves, while increased the amounts of ketones and aldehydes in the first leaf. Enhanced UV-B level had no significant effect on the amounts of alkanes and total wax in all leaves. Exposure to UV-B radiation resulted in wax fusion on adaxial leaf and stomata opening on abaxial leaf. Fusions of plates and rods on adaxial leaf surface covered most of the stomata, thereby influencing the photosynthesis in the upper mesophyll of leaves. Enhanced UV-B level significantly reduced the net photosynthesis rate (P N) but increased the stomata conductance (g s), concentrations of intercellular CO2 (C i ), and transpiration rate (E) in all leaves. Both UV-B radiation and the wax fusion induced by enhanced UV-B radiation resulted in different stomata status on abaxial and adaxial leaf surface, causing decrease of P N, and increase of g s, C i and E in leaves.  相似文献   

2.
Abstract. Epidermal (non-stomatally-controlled) conductance from the fourth leaf, first node leaf, flag leaf and ear of durum wheat (Triticum turgidum var durum L.) grown under Mediterranean field conditions has been measured, along with leaf stomatal frequency and the amount and distribution of epicuticular waxes. Measurements were carried out on varieties and land-races from the Middle East, North Africa, ‘Institut National de la Recherche Agricole’ (INRA) and ‘Centra Internacional de Mejora de Maiz y Trigo’ (CIMMYT). Significant differences were observed among genotypes in the epidermal conductances (ge) of the four organs. For each of the four organs tested, genotypes from the Middle East and CIMMYT showed higher ge. values than those from North Africa and INRA. Ears showed epidermal conductances that were more than four times higher than those of leaves when ge. values were expressed per unit dry weight. The amount of epicuticular waxes was higher in the fourth leaves, intermediate in the first node and flag leaves and lower in the ears. For each organ, ge differences among genotypes were unrelated with the amount of epicuticular waxes. Removal of epicuticular waxes by dipping the organs into chloroform significantly increased the epidermal conductance for the fourth and first node leaves and the ear. However, this did not occur for the flag leaf. For the fourth leaf, ge of intact leaves and ge of leaves in which epicuticular waxes were removed were unrelated (r = -0.265). The regression coefficient of this relation for the first node and flag leaves showed values of 0.666 and 0.650 (P > 0.05), respectively, and values were even higher in the ear (r > m 0.892, P > 0.01). Scanning electron microscope analysis showed that wax bloom decreased from the fourth leaf to the flag leaf, whereas the extent of amorphous wax increased. Wax bloom in leaves consisted mainly of deposits of thin wax plates. In the ears and the adaxial surface of flag leaves, fibrillar waxes predominated. In the first node and flag leaves, the wax deposits on the adaxial side cover the surface of the leaf more densely and uniformly than those on the abaxial side. There was no significant correlation between ge and total stomatal density, or between ge and either adaxial or abaxial stomatal density for any sample of the three different leaves. The contribution of epicuticular waxes plus total stomatal frequency only explained 42.4, 11.8, 28.3 and 16% of ge (per unit leaf area) variations for the fourth leaf, first node leaf, flag leaf and the combined variation of the three leaves together, respectively. From these results, it is concluded that complex interrelationship between different morphophysiological characteristics probably control ge differences among genotypes and that these interrelationships differ for each different plant part.  相似文献   

3.
Artichoke (Cynara scolymus L.) leaf size and shape, glandular and covering trichomes, stomatal density, stomata shape, pore area and epicuticular waxes during micropropagation stages were studied by scanning electron microscopy (SEM) and morphometric analysis with the aim to improve the survival rate after transfer to greenhouse conditions. Leaves from in vitro shoots at the proliferation stage showed a spatular shape, ring-shaped stomata, a large number of glandular trichomes and juvenile covering hairs, but failed to show any epicuticular waxes. Leaves from in vitro plants at the root elongation stage showed a lanceolated elliptic shape with a serrated border, elliptical stomata, decreased pore area percentage, stomatal density, and mature covering trichomes. One week after transfer to ex vitro conditions, epicuticular waxes appeared on the leaf surface and stomata and pore area were smaller as compared to in vitro plants. Artichoke acclimatization may be improved by hormonal stimulation of root development, since useful morphological changes on leaves occurred during root elongation.  相似文献   

4.
The weight of epicuticular wax on the surface of cabbage (Brassica oleracea var. Capitata `Market Prize') leaves was reduced by soil treatments of ethofumesate (2-ethoxy-2,3-dihydro-3,3-dimethyl-5-benzofuranyl methanesulfonate) and EPTC (S-ethyl dipropylthiocarbamate). Separation of epicuticular wax into major components by gas-liquid chromatography indicated that ethofumesate decreased the deposition of n-nonocosane and n-nonocosan-15-one on cabbage leaves but increased the deposition of a minor component, the long chain waxy esters. EPTC was less inhibitory to n-nonocosan-15-one deposition than was ethofumesate. EPTC did not increase long chain waxy ester deposition. Scanning electron micrographs revealed that ethofumesate almost totally eliminated the epicuticular wax on cabbage leaves while EPTC only diminished it. Cuticular transpiration was increased by ethofumesate but not by EPTC. Ethofumesate appears to be a more potent inhibitor of epicuticular wax deposition than EPTC.  相似文献   

5.
We show that induced synomones, emitted as a consequence of Murgantia histrionica activity on Brassica oleracea, are adsorbed by the epicuticular waxes of leaves and perceived by the egg parasitoid Trissolcus brochymenae. Leaves were exposed to M. histrionica females placed on the abaxial leaf surface. After 24 h, the leaves were treated mechanically using gum arabic, or chemically using chloroform, on the adaxial surface, and finally the adaxial surface was assayed with T. brochymenae by two‐choice tests in a closed arena. Wasp females responded to mechanically dewaxed cabbage leaf portions with feeding punctures and footprints (Ff) and with feeding punctures, oviposition and footprints (FOf), showing no effect of wax removal. In contrast, the removal of the epicuticular waxes from leaf portions close to FOf, and from leaves with oviposition and footprints (Of), determined the lack of responses by T. brochymenae. Solvent extracts of different treatments were bioassayed, but only FOf triggered parasitoid response. Thus the detection of oviposition‐induced synomones by the parasitoid depends on their adsorption by the epicuticular waxes. Mechanical wax removal from leaf portions contaminated with host footprints (f) also determined a lack of wasp responses, suggesting that the footprints might trigger the induction of a “footprint‐induced synomone” adsorbed onto the epicuticular waxes and exploited by the parasitoid. Leaf portions with the abaxial lamina previously dewaxed and then contaminated by footprints (D+f) of M. histrionica did not affect the parasitoid response, indicating that the abaxial epicuticular waxes are not directly involved in the chemicals induced by M. histrionica footprints.  相似文献   

6.
The diol constituent of Rhus cotinus leaf epicuticular wax has been identified as nonacosane-5,10-diol from chemical investigations of the free compound, the TMSi ether and the nonacosane-5,10-dione prepared from the diol by oxidation. The form and distribution of the crystalline waxes changed as the leaves expanded, dense clusters of short tubes covering the thin ribbons formed during the initial stages of growth. The diol content of the wax decreased by more than 50% over the same period.  相似文献   

7.
Plant cuticular waxes play a crucial role in limiting nonstomatal water loss. The goal of this study was to localize the transpiration barrier within the layered structure of cuticles of eight selected plant species and to put its physiological function into context with the chemical composition of the intracuticular and epicuticular wax layers. Four plant species (Tetrastigma voinierianum, Oreopanax guatemalensis, Monstera deliciosa, and Schefflera elegantissima) contained only very-long-chain fatty acid (VLCFA) derivatives such as alcohols, alkyl esters, aldehydes, and alkanes in their waxes. Even though the epicuticular and intracuticular waxes of these species had very similar compositions, only the intracuticular wax was important for the transpiration barrier. In contrast, four other species (Citrus aurantium, Euonymus japonica, Clusia flava, and Garcinia spicata) had waxes containing VLCFA derivatives, together with high percentages of alicyclic compounds (triterpenoids, steroids, or tocopherols) largely restricted to the intracuticular wax layer. In these species, both the epicuticular and intracuticular waxes contributed equally to the cuticular transpiration barrier. We conclude that the cuticular transpiration barrier is primarily formed by the intracuticular wax but that the epicuticular wax layer may also contribute to it, depending on species-specific cuticle composition. The barrier is associated mainly with VLCFA derivatives and less (if at all) with alicyclic wax constituents. The sealing properties of the epicuticular and intracuticular layers were not correlated with other characteristics, such as the absolute wax amounts and thicknesses of these layers.The plant cuticle is one of the major adaptations of vascular plants for life in the atmospheric environment. Accordingly, the primary function of cuticles is to limit nonstomatal water loss and, thus, to protect plants against drought stress (Burghardt and Riederer, 2006). However, plant cuticles also play roles in minimizing the adhesion of dust, pollen, and spores (Barthlott and Neinhuis, 1997), protecting tissues from UV radiation (Krauss et al., 1997; Solovchenko and Merzlyak, 2003), mediating biotic interactions with microbes (Carver and Gurr, 2006; Leveau, 2006; Hansjakob et al., 2010, 2011; Reisberg et al., 2012) as well as insects (Eigenbrode and Espelie, 1995; Müller and Riederer, 2005), and preventing deleterious fusions between different plant organs (Tanaka and Machida, 2013).Cuticles are composite (nonbilayer) membranes consisting of an insoluble polymer matrix and solvent-soluble waxes. The polymer matrix (MX) is mainly made of the hydroxy fatty acid polyester cutin (Nawrath, 2006) and also contains polysaccharides and proteins (Heredia, 2003). In contrast, cuticular waxes are complex mixtures of aliphatic compounds derived from very-long-chain fatty acids (VLCFAs) with hydrocarbon chains of C20 and more (Jetter et al., 2007). Wax quantities and compositions vary greatly between plant species and, in many cases, even between organs and developmental stages. Diverse VLCFA derivatives can be present, including free fatty acids, aldehydes, ketones, primary and secondary alcohols, alkanes, and alkyl esters. Besides, the cuticular waxes of many plant species also contain cyclic compounds such as triterpenoids and aromatics.In order to characterize the physiological function of cuticular waxes, methods have been developed for the isolation of astomatous cuticles and the measurement of transpiration rates under exactly controlled conditions, so that well-defined physical transport parameters such as permeances and resistances can be determined and compared across species and organs (Schönherr and Lendzian, 1981; Kerstiens, 1996; Riederer and Schreiber, 2001; Lendzian, 2006). With these methods, it was demonstrated that the cuticular water permeance increases by up to 3 orders of magnitude upon wax removal, thus showing the central role of waxes as a transpiration barrier (Schönherr, 1976). Permeances for water determined so far with astomatous isolated leaf cuticular membranes (CMs) or in situ leaf cuticles range over 2.5 orders of magnitude, from 3.63 × 10−7 m s−1 (Vanilla planifolia) to 7.7 × 10−5 m s−1 (Maianthemum bifolium; Riederer and Schreiber, 2001).The species-dependent differences of both wax composition and permeance led to a search for correlations between cuticle structure and function. If such a structure-function relationship could be established, then it would become possible to select or alter wax composition in order to improve cuticle performance in crop species (Kosma and Jenks, 2007). However, all attempts to understand cuticle permeance based on cuticle composition have failed so far: correlations between wax amounts and permeances could not be established, contrary to the common assumption that thicker wax layers must provide better protection against desiccation (Schreiber and Riederer, 1996; Riederer and Schreiber, 2001). Similarly, a correlation between wax quality (i.e. the relative portions of its constituents) and permeance could also not be established to date (Burghardt and Riederer, 2006). It is not clear how certain wax components contribute to the vital barrier function of the cuticle.Previous attempts to establish wax structure-function relationships may have failed because only bulk wax properties were studied and important effects of substructures were averaged out. However, distinct compartments of wax exist within the cuticle, most prominently as a layer of intracuticular wax embedded within the MX and a layer of epicuticular wax deposited on the outer surface of the polymer (Jeffree, 2006). Over the last years, methods have been developed that allow the selective removal of epicuticular wax by adhesive surface stripping, followed by equally selective extraction of intracuticular wax (Jetter et al., 2000; Jetter and Schäffer, 2001). Chemical analyses showed that, for most plant species investigated to date, both wax layers have distinct compositions (Buschhaus and Jetter, 2011). The most pronounced differences between the layers were found for the triterpenoids, which were localized predominantly (or even exclusively) in the intracuticular wax. These findings raised the possibility that the chemically distinct wax layers might also have distinct functions, leading back to the long-standing question of whether the water barrier function is exerted by the intracuticular and/or the epicuticular wax. There are only scant data to answer this question so far, mainly because methods allowing a distinction between epicuticular and intracuticular waxes were established only recently. Using these sampling techniques, it was recently found that, for leaves of Prunus laurocerasus, the epicuticular wax layer does not contribute to the transpiration barrier (Zeisler and Schreiber, 2016). In contrast, it had been reported that removal of the epicuticular wax layer from tomato (Solanum lycopersicum) fruit caused an approximately 2-fold increase in transpiration, suggesting that, in this species, the epicuticular layer constitutes an important part of the barrier (Vogg et al., 2004). Based on these conflicting reports, it is not clear to what extent the intracuticular or the epicuticular waxes contribute to the sealing function of the plant skin.The goal of this study was to localize the transpiration barrier within the cuticular membrane of selected plant species and to put the physiological function into context with the chemical composition of both the epicuticular and intracuticular wax layers. To this end, we selected eight species from which leaf cuticles could be isolated and methods for step-wise wax removal could be applied without damaging the cuticle. Preliminary studies had shown that the adaxial cuticles on leaves of Citrus aurantium (Rutaceae), Euonymus japonica (Celastraceae), Clusia flava (Clusiaceae), Garcinia spicata (Clusiaceae), Tetrastigma voinierianum (Vitaceae), Oreopanax guatemalensis (Araliaceae), Monstera deliciosa (Araceae), and Schefflera elegantissima (Araliaceae) were astomateous and showed wide chemical diversity. Therefore, these eight species were selected to address the following questions: (1) What are the amounts of epicuticular and intracuticular waxes? (2) Do compositional differences exist between the layers? (3) Where are the cuticular triterpenoids located? (4) How much do the epicuticular and intracuticular waxes contribute to the transpiration barrier? (5) Is the barrier associated with certain components of the intracuticular or epicuticular waxes?  相似文献   

8.
The leaf surface properties of 11 cuticular wax mutants of maize were characterized, and this information was used to identify the quantitative relations among distinct leaf surface traits. Compared with the wild‐type maize, these mutants were reduced 3–24% in their leaf surface hydrophobicity, 20–88% in the mass of cuticular waxes on their leaves, and 52–94% in the percentage of planar leaf surface area covered with epicuticular crystalline waxes. They also differed in the presence and abundance of the epicuticular crystalline waxes in each of seven structural classes. With the exception of one mutant, the mass of cuticular waxes produced by these mutants was positively correlated with the number of epicuticular crystalline waxes per unit area on their leaves. Furthermore, an increase of 0·4 mg of cuticular wax per gram of leaf (dry weight) was associated with a 1% increase in leaf surface area covered by epicuticular crystalline waxes, and this 1% increase was associated with a 2° increase in the contact angle of a water droplet on the leaf surface. Linear differences in the leaf surface hydrophobicity were associated with exponential differences in the mass of the cuticular waxes produced. Quantitative knowledge of these leaf surface properties is highly relevant to the interactions of leaves with environmental factors such as microbes, insects, agricultural chemicals, and pollutants.  相似文献   

9.
Numerous reports have indicated that glucosinolates are important stimulants for specialist herbivores feeding on Brassicaceae, and that these metabolites might be present on the plant surface and thereby detectable by an alighting insect. We investigated the outermost layer of leaves of two species of Brassicaceae, Brassica napus L. var. ‘Martina’ and Nasturtium officinale R. Br., using two highly selective extraction methods. When the epicuticular wax layer was mechanically removed with gum arabic, no glucosinolates were detectable in the lower and upper leaf surfaces. Extracting the leaf surfaces with a threefold short rinse with chloroform/methanol/water (2 : 1 : 1 vol/vol/vol) led to varying results, depending on the light conditions under which plants had been kept in the period prior to extraction. In plants kept under light, glucosinolates were detectable in a first extraction in minor concentrations, with increasing amounts in a second and third extraction. In plants kept in darkness, glucosinolates were almost absent in the first extraction. We postulate that the polar glucosinolates are washed from the inner leaf tissue through open stomata to the outside during solvent extraction, but are not naturally present in the outermost wax layer. The response of the crucifer specialist Phaedon cochleariae (F.) (Coleoptera: Chrysomelidae) to leaf surfaces of the host plants B. napus and N. officinale and to a glucosinolate was tested. Adults preferred both the adaxial and abaxial leaf surfaces of host plants that had been treated with gum arabic in order to remove the epicuticular waxes over intact surfaces. Waxes may therefore prevent direct contact with the stimulants. Sinigrin (allyl glucosinolate) and/or surface extracts of N. officinale leaves applied on Pisum sativum leaf discs did not evoke feeding, but feeding did occur when total leaf extracts of B. napus or N. officinale were applied on this non‐host. We conclude that glucosinolates might only act as feeding stimulants for P. cochleariae in concert with compounds other than surface waxes.  相似文献   

10.
The reaction of plants to environmental factors often varies with developmental stage. It was hypothesized, that also the cuticle, the outer surface layer of plants is modified during ontogenesis. Apple plantlets, cv. Golden Delicious, were grown under controlled conditions avoiding biotic and abiotic stress factors. The cuticular wax surface of adaxial apple leaves was analyzed for its chemical composition as well as for its micromorphology and hydrophobicity just after unfolding of leaves ending in the seventh leaf insertion. The outer surface of apple leaves was formed by a thin amorphous layer of epicuticular waxes. Epidermal cells of young leaves exhibited a distinctive curvature of the periclinal cell walls resulting in an undulated surface of the cuticle including pronounced lamellae, with the highest density at the centre of cells. As epidermal cells expanded during ontogenesis, the upper surface showed only minor surface sculpturing and a decrease in lamellae. With increasing leaf age the hydrophobicity of adaxial leaf side decreased significantly indicated by a decrease in contact angle. Extracted from plants, the amount of apolar cuticular wax per area unit ranged from only 0.9 microgcm(-2) for the oldest studied leaf to 1.5 microgcm(-2) for the youngest studied leaf. Differences in the total amount of cuticular waxes per leaf were not significant for older leaves. For young leaves, triterpenes (ursolic acid and oleanolic acid), esters and alcohols were the main wax components. During ontogenesis, the proportion of triterpenes in total mass of apolar waxes decreased from 32% (leaf 1) to 13% (leaf 7); absolute amounts decreased by more than 50%. The proportion of wax alcohols and esters, and alkanes to a lesser degree, increased with leaf age, whereas the proportion of acids decreased. The epicuticular wax layer also contained alpha-tocopherol described for the first time to be present also in the epicuticular wax. The modifications in the chemical composition of cuticular waxes are discussed in relation to the varying physical characteristics of the cuticle during ontogenesis of apple leaves.  相似文献   

11.
The composition and spatial arrangement of cuticular waxes on the leaves of Prunus laurocerasus were investigated. In the wax mixture, the triterpenoids ursolic acid and oleanolic acid as well as alkanes, fatty acids, aldehydes, primary alcohols and alcohol acetates were identified. The surface extraction of upper and lower leaf surfaces yielded 280 mg m ? 2 and 830 mg m ? 2, respectively. Protocols for the mechanical removal of waxes from the outermost layers of the cuticle were devised and evaluated. With the most selective of these methods, 130 mg m ? 2 of cuticular waxes could be removed from the adaxial surface before a sharp, physically resistant boundary was reached. Compounds thus obtained are interpreted as ‘epicuticular waxes’ with respect to their localization in a distinct layer on the surface of the cutin matrix. The epicuticular wax film can be transferred onto glass and visualized by scanning electron microscopy. Prunus laurocerasus epicuticular waxes consisted entirely of aliphatic compounds, whereas the remaining intracuticular waxes comprised 63% of triterpenoids. The ecological relevance of this layered structure for recognition by phytotrophic fungi and herbivorous insects that probe the surface composition for sign stimuli is discussed.  相似文献   

12.
BACKGROUND AND AIMS: The morphology of the epicuticular leaf waxes of Wollemia nobilis (Araucariaceae) was studied with special emphasis on the relationship between the microstructure of epicuticular wax crystals and their chemical composition. Wollemia nobilis is a unique coniferous tree of the family Araucariaceae and is of very high scientific value as it is the sole living representative of an ancient genus, which until 1994 was known only from fossils. METHODS: Scanning electron microscopy (SEM), gas chromatography (GC) combined with mass spectrometry (GC-MS) and nuclear magnetic resonance spectroscopy (NMR) were used for characterizing the morphology and the chemical structure of the epicuticular wax layer of W. nobilis needles. KEY RESULTS: The main component of the leaf epicuticular wax of W. nobilis is nonacosan-10-ol. This secondary alcohol together with nonacosane diols is responsible for the tubular habit of the epicuticular wax crystals. Scanning electron micrographs revealed differences in the fine structure of adaxial and abaxial leaf surfaces that could be explained by gas chromatographic studies after selective mechanical removal of the waxes. CONCLUSIONS: SEM investigations established the tubular crystalline microstructure of the epicuticular wax of W. nobilis leaves. GC-MS and NMR experiments showed that nonacosan-10-ol is the major constituent of the epicuticular wax of W. nobilis leaves.  相似文献   

13.
Total reflectance of ultraviolet and photosynthetically effective wavelengths was measured for a range of different leaf types. Two approaches were employed. Firstly, reflectance of monochromatic wavebands at 330 and 680 nm was measured for a total of 45 different species covering a wide range of genera. In the second, specific leaf types that displayed different degrees of reflectance were treated to remove hairs and waxes that contributed to their reflectance. Selected waxy and non‐waxy leaves were also studied in more detail over the spectral range 270–500 nm. It was found that both pubescence (presence of hairs) and glaucousness (presence of a thick epicuticular wax layer) had marked effects on total reflectance. Pubescent leaves tended to be more effective in reflecting longer wavelengths than ultraviolet radiation. The extent of this effect depended on hair type. Glaucous leaves demonstrated that surface waxes were very effective reflectors of both UV and longer wavelength radiation.  相似文献   

14.
Cuticular waxes play a pivotal role in limiting transpirational water loss across the plant surface. The correlation between the chemical composition of the cuticular waxes and their function as a transpiration barrier is still unclear. In the present study, intact tomato fruits (Lycopersicon esculentum) are used, due to their astomatous surface, as a novel integrative approach to investigate this composition- function relationship: wax amounts and compositions of tomato were manipulated before measuring unbiased cuticular transpiration. First, successive mechanical and extractive wax-removal steps allowed the selective modification of epi- and intracuticular wax layers. The epicuticular film consisted exclusively of very-long-chain aliphatics, while the intracuticular compartment contained large quantities of pentacyclic triterpenoids as well. Second, applying reverse genetic techniques, a loss-of-function mutation with a transposon insertion in a very-long-chain fatty acid elongase beta-ketoacyl-CoA synthase was isolated and characterized. Mutant leaf and fruit waxes were deficient in n-alkanes and aldehydes with chain lengths beyond C30, while shorter chains and branched hydrocarbons were not affected. The mutant fruit wax also showed a significant increase in intracuticular triterpenoids. Removal of the epicuticular wax layer, accounting for one-third of the total wax coverage on wild-type fruits, had only moderate effects on transpiration. By contrast, reduction of the intracuticular aliphatics in the mutant to approximately 50% caused a 4-fold increase in permeability. Hence, the main portion of the transpiration barrier is located in the intracuticular wax layer, largely determined by the aliphatic constituents, but modified by the presence of triterpenoids, whereas epicuticular aliphatics play a minor role.  相似文献   

15.
Elucidation of the mechanism of adsorption of particles suspended in the gas-phase (aerosol) to the outer surfaces of leaves provides useful information for understanding the mechanisms of the effect of aerosol particles on the growth and physiological functions of trees. In the present study, we examined the localization of artificially deposited sub-micron-sized carbon-based particles on the surfaces of needles of Cryptomeria japonica, a typical Japanese coniferous tree species, by field-emission scanning electron microscopy. The clusters (aggregates) of carbon-based particles were deposited on the needle surface regions where epicuticular wax crystals were sparsely distributed. By contrast, no clusters of the particles were found on the needle surface regions with dense distribution of epicuticular wax crystals. Number of clusters of carbon-based particles per unit area showed statistically significant differences between regions with sparse epicuticular wax crystals and those with dense epicuticular wax crystals. These results suggest that epicuticular wax crystals affect distribution of carbon-based particles on needles. Therefore, densely distributed epicuticular wax crystals might prevent the deposition of sub-micron-sized carbon-based particles on the surfaces of needles of Cryptomeria japonica to retain the function of stomata.  相似文献   

16.
Removal of trimethylamine (TMA) by 10 different living Sansevieria spp. and their dried leaf materials was studied. The results showed that living Sansevieria kirkii was the most effective plant while Sansevieria masoniana was the least effective in TMA removal. Two major pathways were involved in stomata opening and epicuticular wax on the leaf surface. In the presence of TMA, the stomata opening in Sansevieria spp. was induced, which enhanced TMA removal under light conditions. Dried leaf powders of Sansevieria spp. adsorbed TMA through their waxes. Therefore, both living and non-living Sansevieria spp. can be effectively used for removal of TMA.  相似文献   

17.
Plant surface characteristics were repeatedly shown to play a pivotal role in plant–pathogen interactions. The abaxial leaf surface of perennial ryegrass (Lolium perenne) is extremely glossy and wettable compared to the glaucous and more hydrophobic adaxial surface. Earlier investigations have demonstrated that the abaxial leaf surface was rarely infected by powdery mildew (Blumeria graminis), even when the adaxial surface was densely colonized. This led to the assumption that components of the abaxial epicuticular leaf wax might contribute to the observed impairment of growth and development of B. graminis conidia on abaxial surfaces of L. perenne. To re-assess this hypothesis, we analyzed abundance and chemical composition of L. perenne ab- and adaxial epicuticular wax fractions. While the adaxial epicuticular waxes were dominated by primary alcohols and esters, the abaxial fraction was mainly composed of n-alkanes and aldehydes. However, the major germination and differentiation inducing compound, the C26-aldehyde n-hexacosanal, was not present in the abaxial epicuticular waxes. Spiking of isolated abaxial epicuticular Lolium waxes with synthetically produced n-hexacosanal allowed reconstituting germination and differentiation rates of B. graminis in an in vitro germination assay using wax-coated glass slides. Hence, the absence of the C26-aldehyde from the abaxial surface in combination with a distinctly reduced surface hydrophobicity appears to be primarily responsible for the failure of normal germling development of B. graminis on the abaxial leaf surfaces of L. perenne.  相似文献   

18.
Epicuticular and intracuticular waxes from both adaxial and abaxial surfaces of the leaves of Kalanchoe daigremontiana were analyzed. All wax mixtures were found to contain approximately equal amounts of triterpenoids and very long chain fatty acid (VLCFA) derivatives. The triterpenoid fraction consisted of glutinol (8-19% of the total wax) and friedelin (4-9%), together with smaller amounts of glutanol, glutinol acetate, epifriedelanol, germanicol and β-amyrin. The VLCFA derivatives comprised C27-C35 alkanes (19-37% of the total wax), C32-C34 aldehydes (3-7%), C32 and C34 fatty acids (0.2-3%), C26-C36 primary alcohols (4-8%), and C42-C52 alkyl esters (2-9%). The wax layers were found to differ in triterpenoid amounts, with the intracuticular wax containing higher percentages of most triterpenoids than the epicuticular wax. Friedelin, the only triterpenoid ketone present, showed the opposite distribution with higher proportions in the epicuticular wax. VLCFA derivatives also accumulated to higher percentages in the epicuticular than in the intracuticular wax layer. Epicuticular wax crystals were observed on both the adaxial and abaxial leaf surfaces.  相似文献   

19.
广州市常见行道树种叶片表面形态与滞尘能力   总被引:16,自引:0,他引:16  
刘璐  管东生  陈永勤 《生态学报》2013,33(8):2604-2614
以广州市常见的18种行道树为对象,通过扫描电镜观察比较了行道树的叶表面形态结构、应用接触角测定仪测定了绿化树种叶片的接触角对滞尘能力的影响.结果表明:不同树种的滞尘量差异显著,18种植物叶片雨后第26天的最大滞尘量在0.066-1.831 g/m2,物种间相差达27倍以上.叶表面具有网状结构,气孔密度较大(20<气孔密度<60个)且气孔开口较大(如芒果)容易滞留粉尘;叶表面平滑具有蜡质层,气孔排列整齐,无明显起伏(如红花羊蹄甲、桃花心木、大叶紫薇、鹅掌藤),滞尘能力较弱.植物叶片接触角与滞尘量呈负相关(r=-0.614),接触角<90°的表现为亲水性.易润湿的植物叶片雨后第26天最大滞尘量在1.0-1.831 g/m2,叶片表面的形态结构凹凸不平,具有钩状或脊状褶皱、突起等且20<气孔密度<60范围内,测得的接触角较小(芒果、重阳木、高山榕),使得粉尘与植物叶片接触面积较大,粉尘不易从叶面脱落,滞尘能力较强.而接触角较大的盆架树、麻楝、大叶紫薇、鹅掌藤和红花羊蹄甲的滞尘量均<1.0g/m2,其特殊的表面结构和疏水的蜡质使颗粒物不易吸附在植物叶片上,因此滞尘能力较弱.由此可见,植物叶表面蜡质含量和气孔密度及其叶片接触角的大小是影响植物叶片滞尘能力的主要因素,在进行城市绿化时,适当考虑选择叶表面形态有利于滞尘的绿化树种,将可提高城市植被的环境效应.  相似文献   

20.
Anthyllis cytisoides L. is highly colonized by arbuscular mycorrhizal fungi (AMF) and behaves as a drought-avoider species in the field. Our objectives were: (1) to study the response of A. cytisoides when exposed to moderate (acclimation) or severe (peak) drought and subsequent rewatering under nursery conditions; and (2) to verify if AMF improved the adaptation of A. cytisoides to stress. The soil compactness in drought-acclimated treatments increased four times compared with that of well-watered controls, which could reinforce the effects of water deficit on plant physiology. Photosynthetic rates decreased by around 50% and 70% and leaf conductance decreased by 40% and 50% in drought-acclimated non-mycorrhizal and mycorrhizal plants, respectively. Peak drought limited plant growth, accelerated leaf senescence and induced the conversion of starch into soluble sugars in the leaves of stressed plants. The accumulation of sugars could contribute to a decrease in water potential in order to achieve the required tension to let water move from soil to shoot. Mycorrhizal plants showed a two-fold higher chlorotic leaf biomass than non-mycorrhizal plants under severe drought. Moreover, mycorrhizal A. cytisoides showed enhanced epicuticular waxes on the surfaces of the remaining green leaves. Increased leaf senescence, together with wax deposition, could reduce whole plant transpiration, thus allowing mycorrhizal plants to maintain a higher leaf relative water content (50%) than non-mycorrhizal plants (35%). After drought recovery, leaf abscission in stressed mycorrhizal plants was 10 times greater than that in non-mycorrhizal plants. The results suggest that AMF conferred greater responsiveness of A. cytisoides to drought. Enhanced wax deposition and leaf senescence could be an ecological adaptation to cope with severe water deficit.  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