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 surface pressure isotherms of chlorophyll a, monogalactosyldiacylglycerol and phytol at the air-water interface were studied on a Langmuir trough at 20.0±0.5°C. The subphase was a phosphate buffer, 10?3 M at pH 8.0. The extrapolated limiting areas per molecule are 115, 82 and 38 Å2/molecule, respectively. The isotherms of eight mixtures of chlorophyll a with monogalactosyldiacylglycerol and eight mixtures of chlorophyll a with phytol, covering in both cases the whole range of molar fractions have also been measured. The results for the mixed monolayers were analysed in terms of the additivity rule. They show that a small negative deviation with respect to ideality is observed upon mixing chlorophyll a with monogalactosyldiacylglycerol. However, chlorophyll a forms an ideal two-dimensional solution when mixed with phytol. The excess free energies of mixing of chlorophyll a with monogalactosyldiacylglycerol as a function of concentration were calculated from the surface pressure isotherms at 10, 15, 20 and 25 mN·m?1. The values are negative, reflecting the interaction prevailing between these components in the monolayers. For the four surface pressures studied, the excess free energies are symmetrical with respect to the mode fraction. The values for an equimolar mixture range from ?300 to ?540 J·mol?1 at 10 and 25 mN·m?1, respectively. A comparison between the thermodynamics of mixing of chlorophyll a with monogalactosyldiacylglycerol and phytol suggests that the polar head of monogalactosyldiacylglycerol together with the polar groups of chlorophyll a are probably involved in the interaction. However, this does not completely rule out the possibility that structural effects due to a different packing of chlorophyll a with monogalactosyldiacylglycerol and phytol may also be involved. Furthermore it is shown that the small interactions between these constituents are not inconsistent with the specific coupling existing between the apoprotein of the chlorophyll a-protein complexes and chlorophyll a.  相似文献   

2.
As part of a study to investigate the effect of oil seeps on intertidal organisms, oil extracts of Blackstone oil shale from Kimmeridge on the Dorset coast were used in laboratory experiments to test their effect on the settlement of the barnacle Balanus balanoides (L.). Thin films of oil extract painted on the surface of pits in slate panels had no effect on cyprid settlement when applied up to a surface density of 2.8 g · m?2, representing a thickness of 3.3 μm. Larger surface densities of oil stimulated cyprids to settle in far greater numbers than on unoiled panels. The maximum effect was obtained at a surface density of between 14.0 and 56.0 g · m ?2, representing a thickness of 16.5 μm and 66.0 μm. With higher concentration of oil in the pits, stimulation to settle was reduced although cyprid settlement was still encouraged at a surface density of oil of 112g · m?2 or 132 μm thickness.The unfractionated crude oil shale extract was a less powerful stimulus for barnacle settlement than a partially purified solution of the integumental protein arthropodin, another strong settlement inducer for barnacle cyprids.  相似文献   

3.
Bao Y J  Li Z H  Han X G  Han G D  Zhong Y K 《农业工程》2007,27(11):4443-4451
The caloric contents of 42 species and their composition in a Leymus chinensis steppe community of Inner Mongolia, China were determined and analyzed based on the field experiment for 11 years. The caloric content (x ± SD) of aboveground parts of plant species varied from (13156 ± 1141) J·g?1 (ash contained) to (18141 ± 527) J·g?1. The average caloric content of all species was (16899 ± 840) J·g?1 and the inter-specific CV (coefficient of variation) was 4.9%. Of all the species, Caragana microphylla had the highest caloric content (18142 J·g?1). Grasses had a higher average caloric content ((17425 ± 291) J·g?1) than forbs ((16734 ± 844)J·g?1). When the herbaceous species were classified into subgroups according to life-form and growth-form, the order of average caloric contents, from high to low, was: tall grasses ((17717 ± 92) J·g?1) > legume ((17228 ± 433) J·g?1) > short grasses ((17250 ± 218) J·g?1) > remaining forbs ((16784 ± 529) J·g?1) > subshrubs ((16719 ± 69) J·g?1) > annuals and biennials ((15911 ± 1759) J·g?1). There was a positive correlation (P < 0.05) between the caloric contents of 42 species and their relative biomass in the community. When all species were classified into 3 groups according to their composition in the community, the average caloric contents, weighed by the species relative biomass, followed the order: dominant (17740 J·g?1) > companion (17244 J·g?1) > incidental (16653 J·g?1). The plants with high caloric contents were more competitive, which allowed them to gain a dominant status, whereas the competitive abilities of plants with low caloric contents were generally weak. The latter made up the companion or incidental species in a steppe community.  相似文献   

4.
The caloric contents of 42 species and their composition in a Leymus chinensis steppe community of Inner Mongolia, China were determined and analyzed based on the field experiment for 11 years. The caloric content (x ± SD) of aboveground parts of plant species varied from (13156 ± 1141) J·g?1 (ash contained) to (18141 ± 527) J·g?1. The average caloric content of all species was (16899 ± 840) J·g?1 and the inter-specific CV (coefficient of variation) was 4.9%. Of all the species, Caragana microphylla had the highest caloric content (18142 J·g?1). Grasses had a higher average caloric content ((17425 ± 291) J·g?1) than forbs ((16734 ± 844)J·g?1). When the herbaceous species were classified into subgroups according to life-form and growth-form, the order of average caloric contents, from high to low, was: tall grasses ((17717 ± 92) J·g?1) > legume ((17228 ± 433) J·g?1) > short grasses ((17250 ± 218) J·g?1) > remaining forbs ((16784 ± 529) J·g?1) > subshrubs ((16719 ± 69) J·g?1) > annuals and biennials ((15911 ± 1759) J·g?1). There was a positive correlation (P < 0.05) between the caloric contents of 42 species and their relative biomass in the community. When all species were classified into 3 groups according to their composition in the community, the average caloric contents, weighed by the species relative biomass, followed the order: dominant (17740 J·g?1) > companion (17244 J·g?1) > incidental (16653 J·g?1). The plants with high caloric contents were more competitive, which allowed them to gain a dominant status, whereas the competitive abilities of plants with low caloric contents were generally weak. The latter made up the companion or incidental species in a steppe community.  相似文献   

5.
Penetration of 1-alkanols into monolayers of hydrophobic polypeptides, poly(ε-benzyloxycarbonyl-l-lysine) and poly(ε-benzyloxycarbonyl-dl-lysine), was compared with their adsorption on the air/water interface in the absence of monolayers. The polypeptide prepared from l-lysine is generally considered to be in the α-helical form whereas dl-copolymer polypeptide contains random-coiled portions due to the structural incompatibility between the two isomers. The free energy of adsorption of 1-alkanols on the air/water interface at dilute concentrations was ?0.68 kcal·mol?1 per methylene group and 0.15 kcal·mol?1 for the hydroxyl group at 25°C. In the close-packed state, the surface area occupied by each molecule of 1-alkanols of varying carbon chain-lengths showed nearly a constant value of about 27.2 Å2, indicating perpendicular orientation of the alkanol molecules at the interface. About 75% of the water surface was covered by 1-butanol in this close-packed state. The mode of adsorption of 1-alkanols on the vacant air/water interface followed the Gibbs surface excess while the mode on the polypeptide membranes followed the Langmuir adsorption isotherm, indicating that the latter is characterized by the presence of a finite number of binding sites. The free energies of adsorption of 1-alkanols on the l-polymer monolayers were more negative than those on the vacant air/water interface and less negative than those on the dl-copolymer monolayers. Thus, the affinity of 1-alkanols to the interface was in the order of vacant air/water interface <l-polymer <dl-copolymer. The difference between the air/water interface and l-polymer was about 0.54 kcal·mol?1 and that between l-polymer and dl-copolymer was 0.17 kcal·mol?1 at 25°C: the adsorption of 1-alkanols to the dl-copolymer was favored compared to the l-polymer. The polar moieties of the backbone of the dl-copolymer may be exposed to the aqueous phase at the disordered portion. Dipole interaction between this portion and 1-alkanol molecules may account for the enhanced adsorption of the alkanols to the dl-copolymer.  相似文献   

6.
The exothermic effects observed on wetting pectins with water and aliphatic alcohols were studied using a microcalorimeter.The heat released on wetting 1 g pectin with water was found to be 171 ± 7·5 J g?1. It was experimentally established that 1 g of dry pectin exothermically bonded up to 0·57 g of water.By using the Gibbs-Helmholtz-Young equation which relates the heat released by wetting to the area of the wetted surface, it was estimated that the surface accessible to water in 1 g of pectin was 1·46 × 103 m2 g?1. The heat of hydration was independent of the degree of esterification of the pectin. The experimental results revealed that there were about six molecules of energetically bonded water per monomer unit of pectin.A specific interaction between methanol and the methoxyl groups of pectin was observed on wetting pectins with methanol and dependence was established between the released heat and the degree of esterification. No similar dependence was reported for the remaining aliphatic alcohols.  相似文献   

7.
To verify the validity of thermodynamic approaches to the prediction of cellular behavior, cell spreading of three different cell types on solid substrata was determined in vitro. Solid substrata as well as cell types were selected on the basis of their surface free energies, calculated from contact angle measurements. The surface free energies of the solid substrata ranged from 18–116 erg cm−2. To measure contact angles on cells, a technique was developed in which a multilayer of cells was deposited on a filter and air dried. Cell surface free energies ranged from 60 erg cm−2 for fibroblasts, and 57 for smooth muscle cells, to 91 for HeLa epithelial cells. After adsorption of serum proteins, cell surface free energies of all three cell types converged to approx 74 erg cm−2. The spreading of these cell types from RPMI 1640 medium on the various solid substrata showed that both in the presence and in the absence of serum proteins in the medium, cells spread poorly on low energy substrata (Y s <50 erg cm−2), whereas good cell spreading was observed on the higher energy substrata. Calculations of the interfacial free energy of adhesion (ΔF adh) show that ΔF adh decreases with increasingY s , and equals zero around 45 erg cm−2 for all three cell types in the presence of serum proteins and for HeLa epithelium cells in the absence of serum proteins. This explains the spreading of these cells on the various substrata upon a thermodynamic basis. The results clearly show that substratum surface free energy has a predictive value with respect to cell spreading in vitro, both in the presence and absence of serum proteins. It is noted, however, that interfacial thermodynamics fail to explain the behavior of fibroblasts and smooth muscle cells in the absence of serum proteins, most likely because of the relatively high surface charges of these two cell types.  相似文献   

8.
The small intestine is known to possess mechanisms for intact transport and membrane hydrolysis of oligopeptides. To determine the relative role of these processes in peptide assimilation the fate of two model peptides known to be high-affinity substrates for the brush border aminooligopeptidase were studied in rat small intestine in vivo. Both 20 mM Gly-L-Pro, a potent inhibitor of peptide transport, and specific inhibitors of the aminopeptidase, 10 mM L-Ala-β-naphthylamide or the phthalimido derivative of 0.1 mM L-leucine bromomethyl ketone, reduced assimilation of L-Leu-Gly-Gly and L-Leu-L-Leu. Further inhibition was found when both transport and peptidase inhibitors were included in the intestinal perfusate suggesting that the model di- and tripeptides utilize both intact transport and surface hydrolysis for their assimilation. Although comparative kinetic parameters of intact transport (Km = 22 mM; V = 1.9 · 10?3μmol · s?1 · cm?2) and surface hydrolysis (Km = 8.7; V = 1.1 · 10?3) for l-Leu-l-Leu differed markedly, the relationship of peptide concentration to assimilation rate was nearly identical for intact transport and surface hydrolysis in the physiological range of 1–10 mM substrate. Both intact peptide transport and surface hydrolysis appear to be efficient and complementary processes that promote efficient assimilation of dipeptides and tripeptides. The relative importance of each assimilation process appears to depend upon the amino acid composition of the peptide nutrient.  相似文献   

9.
Three photosynthetic parameters of 7 species of marine diatoms were studied using Na214CO3 at 5–8 C using log phase axenic cultures. The cell volumes of the different species varied from 70 μm3 to 40 × 105μm3. The present experiment is consistent with the interpretation that the initial slope α (mg C · [mg chl a]?1· h?1· w?1· m2) of photosynthesis vs. light curves is controlled by self-shading of chlorophyll a in the cell. Pm, the rate of photosynthesis at light saturation (mg C · [mg cell, C]?1· h?1) and R, the intercept at zero light intensity (mg C · [mg cell C]?1· H?1) are both dependent on the ratio of surface area to volume of cell.  相似文献   

10.
The purpose of this study was to elucidate the binding of paeonol to human serum albumin (HSA) through spectroscopic methods. The fluorescence quenching of HSA by paeonol was a result of the formation of the HSA–paeonol complex with low binding affinity (K = 4.45 × 103 M?1 at 298 K). Thermodynamic parameters (ΔG = –2.08 × 104 J·mol?1, ΔS = 77.9 J·mol?1·K?1, ΔH = 2.41 × 103 J·mol?1, kq = 9.67 × 1012 M?1·s?1) revealed that paeonol mainly binds HSA through hydrophobic force following a static quenching mode. The binding distance was estimated to be 1.91 nm by fluorescence resonant energy transfer. The conformation of HSA was changed and aggregates were formed in the presence of paeonol, revealed by synchronous fluorescence, circular dichroism, Fourier transform infrared spectroscopy, three‐dimensional fluorescence spectroscopy, and resonance light scattering results.  相似文献   

11.
Ulothrix zonata (Weber and Mohr) Kütz. is an unbranched filamentous green alga found in rocky littoral areas of many northern lakes. Field observations of its seasonal and spatial distribution indicated that it should have a low temperature and a high irradiance optimum for net photosynthesis, and at temperatures above 10°C it should show an increasingly unfavorable energy balance. Measurements of net photosynthesis and respiration were made at 56 combinations of light and temperature. Optimum conditions were 5°C and 1100 μE·m?2·s?1 at which net photosynthesis was 16.8 mg O2·g?1·h?1. As temperature increased above 5° C optimum irradiance decreased to 125 μE·m?2·s?1 at 30°C. Respiration rates increased with both temperature and prior irradiance. Light-enhanced respiration rates were significantly greater than dark respiration rates following irradiance exposures of 125 μE·m?2·s?1 or greater. Polynomials were fitted to the data to generate response surfaces. Polynomial equations represent statistical models which can accurately predict photosynthesis and respiration for inclusion in ecosystem models.  相似文献   

12.
Mutations at protein–protein recognition sites alter binding strength by altering the chemical nature of the interacting surfaces. We present a simple surface energy model, parameterized with empirical values, yielding mean energies of ?48 cal mol?1 Å?2 for interactions between hydrophobic surfaces, ?51 to ?80 cal mol?1 Å?2 for surfaces of complementary charge, and 66–83 cal mol?1 Å?2 for electrostatically repelling surfaces, relative to the aqueous phase. This places the mean energy of hydrophobic surface burial at ?24 cal mol?1 Å?2. Despite neglecting configurational entropy and intramolecular changes, the model correlates with empirical binding free energies of a functionally diverse set of rigid‐body interactions (r = 0.66). When used to rerank docking poses, it can place near‐native solutions in the top 10 for 37% of the complexes evaluated, and 82% in the top 100. The method shows that hydrophobic burial is the driving force for protein association, accounting for 50–95% of the cohesive energy. The model is available open‐source from http://life.bsc.es/pid/web/surface_energy/ and via the CCharpPPI web server http://life.bsc.es/pid/ccharppi/ . Proteins 2015; 83:640–650. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
A dense community of shade adapted microalgae dominated by the diatom Trachyneis aspera is associated with a siliceous sponge spicule mat in McMurdo Sound, Antarctica. Diatoms at a depth of 20 to 30 m were found attached to spicule surfaces and in the interstitial water between spicules. Ambient irradiance was less than 0.6 μE · m?2· s?1 due to light attenuation by surface snow, sea ice, ice algae, and the water column. Photosynthesis-irradiance relationships determined by the uptake of NaH14CO3 revealed that benthic diatoms beneath annual sea ice were light-saturated at only 11 μE·m?2·s?1, putting them among the most shade adapted microalgae reported. Unlike most shade adapted microalgae, however, they were not photoinhibited even at irradiances of 300 μE·m?2·s?1. Although in situ primary production by benthic diatoms was low, it may provide a source of fixed carbon to the abundant benthic invertebrates when phytoplankton or ice algal carbon is unavailable.  相似文献   

14.
The caseinate-induced competitive displacement of whey protein from planar air-water interfaces was investigated based on atomic force microscopy (AFM) imaging and that from the surfaces of oil droplets immersed in aqueous solution based on AFM force spectroscopy. After the addition of sodium caseinate to the sub-phase, the surface pressure of planar interfacial films of pre-adsorbed whey protein increased from 8 mN/m to up to 21 mN/m. The thicknesses of interfacial films were uniform and remained to be approximately 2 nm at relatively low surface pressures up to 18 mN/m, while they became uneven at higher surface pressures and increased to up to 7.1 nm, presumably due to the compression of interfacial whey protein networks by adsorbed caseinate. The rigidity of oil droplets coated with protein adsorbed to their surfaces was then evaluated based on the slope of approximately linear force-distance curves obtained by pressing an oil droplet against another. The adsorption of whey protein to oil droplet surfaces increased droplets’ rigidity. The subsequent addition of caseinate to the bulk solution surrounding oil droplets coated with pre-adsorbed whey protein further increased droplets’ rigidity. The present results suggest that caseinate adsorbed to an interface to which whey protein had adsorbed in advance did not completely expel pre-adsorbed whey protein molecules into the aqueous phase but caused a compaction of interfacial whey protein networks and thereby strengthened the interfacial film.  相似文献   

15.
Nano-engineered superhydrophobic surfaces have been investigated for potential fouling resistance properties. Integrating hydrophobic materials with nanoscale roughness generates surfaces with superhydrophobicity that have water contact angles (θ) >150° and concomitant low hysteresis (<10°). Three superhydrophobic coatings (SHCs) differing in their chemical composition and architecture were tested against major fouling species (Amphora sp., Ulva rigida, Polysiphonia sphaerocarpa, Bugula neritina, Amphibalanus amphitrite) in settlement assays. The SHC which had nanoscale roughness alone (SHC 3) deterred the settlement of all the tested fouling organisms, compared to selective settlement on the SHCs with nano- and micro-scale architectures. The presence of air incursions or nanobubbles at the interface of the SHCs when immersed was characterized using small angle X-ray scattering, a technique sensitive to local changes in electron density contrast resulting from partial or complete wetting of a rough interface. The coating with broad spectrum antifouling properties (SHC 3) had a noticeably larger amount of unwetted interface when immersed, likely due to the comparatively high work of adhesion (60.77 mJ m?2 for SHC 3 compared to 5.78 mJ m?2 for the other two SHCs) required for creating solid/liquid interface from the solid/vapour interface. This is the first example of a non-toxic, fouling resistant surface against a broad spectrum of fouling organisms ranging from plant cells and non-motile spores, to complex invertebrate larvae with highly selective sensory mechanisms. The only physical property differentiating the immersed surfaces is the nano-architectured roughness which supports longer standing air incursions providing a novel non-toxic broad spectrum mechanism for the prevention of biofouling.  相似文献   

16.
Initial microbial adhesion to surfaces is a complicated process that is affected by a number of factors. An important property of a solution that may influence adhesion is pH. The surface properties of the cedar wood were characterized by the sessile drop technique. Moreover, the interfacial free energy of surface adhesion to the cedar wood was determined under pH values (2, 3, 5, 7, 9 and 11). The results showed that cedar wood examined at different pH levels could be considered hydrophobic ranged from Giwi = ?13.1 mJ/m2 to Giwi = ?75 mJ/m2. We noted that the electron-donor character of cedar wood was important at both basic and limit acidic conditions (pH 11 and pH 3) and it decreased at intermediate pH (pH 5). The cedar wood substratum presents a weak electron acceptor under various pH’s. In addition, the adhesion of conidia from Penicilllium expansum to the cedar wood surfaces at different pH values (2, 3, 5, 7, 9 and 11) was investigated using Environmental Scanning Electron Microscopy and image analysis was assessed with the Mathlab® program. The data analysis showed that the conidia from P. expansum were strongly influenced by the pH. The maximum adhesion occurs in the pH 11 and pH 3 and decreased to 24% at pH 5.  相似文献   

17.
Sea ice microalgae in McMurdo Sound, Antarctica were examined for photosynthesis-irradiance relationships and for the extent and time course of their photoadaptation to a reduction in in situ irradiance. Algae were collected from the bottom centimeter of coarse-grained congelation ice in an area free of natural snow cover. Photosynthetic rate was determined in short term (1 h) incubations at ?2° C over a range of irradiance from 0 to 286 μE·m?2·s?1. Assimilation numbers were consistently below 0.1 mg C·mg chl a?1·h?1. The Ik's3 averaged only 7 μE·m?2·s?1, and photosynthesis was inhibited at irradiances above 25 μE·m?2·s?1. Photosynthetic parameters of the ice algal community were examined over a nine day period following the addition of 4 cm of surface snow while a control area remained snow-free. A reduction of 40% in PmB relative to the control occurred after two days of snow cover; α, β, Ik, and Im were not significantly altered. Low assimilation numbers and constant standing crop size, however, suggested that the algal bloom may have already reached stationary growth phase, possibly minimizing their photoadaptive response.  相似文献   

18.
Abstract According to computer energy balance simulations of horizontal thin leaves, the quantitative effects of stomatal distribution patterns (top vs. bottom surfaces) on transpiration (E) were maximal for sunlit leaves with high stomatal conductances (gs) and experiencing low windspeeds (free or mixed convection regimes). E of these leaves decreased at windspeeds > 50 cm s?1, despite increases in the leaf-to-air vapour density deficit. At 50 cm s?1 wind-speed, rapidly transpiring leaves had greater E when one-half of the stomata were on each leaf surface (amphistomaty; 10.16 mmol H2O m?2 s?1) than when all stomata were on either the top (hyperstomaty; 9.34 mmol m?2s?1) or bottom (hypostomaty; 7.02 mmol m?2s?1) surface because water loss occurred in parallel from both surfaces. Hyperstomatous leaves had larger E than hypostomatous leaves because free convection was greater on the top than on the bottom surface. Transpiration of leaves with large g, was greatest at windspeeds near zero when ~60–75% of the stomata were on the top surface, while at high windspeeds E was greatest with, 50% of the stomata on top. For leaves with low gs, stomatal distribution exerted little influence on simulated E values. Laboratory measurements of water loss from simulated hypo-, hyper-, and amphistomatous leaf models qualitatively supported these predictions.  相似文献   

19.
The seasonal abundance of epilithic algae was correlated with major physico-chemical parameters in a first-order, heavily shaded stream in northern Arizona. Diatoms made up over 85%, by numerical abundance, of the epilithon community Light energy, water temperature, and stream discharge were most highly correlated with seasonal abundance of epilithic diatom taxa when analyzed with stepwise multiple regression. None of the chemical variables measured in the study (NO3-N, O-PO4, SiO2, including PH) was found to be significantly correlated with the seasonal community structure of epilithic diatoms. Total diatom cell densities showed a significant negative correlation to stream bed light energy. Likewise, total diatom cell densities along a transect in the stream bed showed a negative correlation to current velocity during those months when base flow was low and stable, and current velocity was ≤25 cm·sec-1. Most diatom taxa had highest cell densities at temperatures < 16°C and at daily mean stream bed light levels < 400 μE·m?2·s?1. Highest cell densities of green algae occurred at temperatures between 6–16°C and at daily mean stream bed light levels of > 400 μE·m?2·s?1. Blue-green algae (cyanobacteria) grew best at the highest recorded water temperatures and daily mean stream bed light energy (16–20°C and 900–1200 μE·m?2·s?1). Abrupt increases in NO3-N coincided with a brief pulse of Nostoc pruniforme colonies during June, and leaf drop from Alnus oblongifolia during October.  相似文献   

20.
The responses of the early development of Laminaria japonica collected from Kiaochow Bay in China to enhanced ultraviolet-B radiation (UV-B, 280–320 nm) were studied in the laboratory. The low UV-B radiations (11.7–23.4 J·m−2·d−1) had no significant effects on zoospores attachment, but when the UV-B dose > 35.1 J·m−2·d−1 the attachment decreased significantly compared with the control. Germination of embryospores was >93% under the low (11.7–35.1 J·m−2·d−1) doses, and in the range of 78.5%–88.5% under the high (46.8–70.2 J·m−2·d−1) UV-B doses, indicating a significant radiation effect. Under the higher UV-B exposure (35.1–70.2 J·m−2·d−1), all of the few gametophytes formed from embryospores died 120 h post-release. After exposure to the low UV-B radiation (11.7–23.4 J·m−2·d−1), the formation of sporophytes decreased and the female gametophyte clones increased compared with the control. However, the sex ratio and the relative growth of female gametophytes/sporophytes had not significantly changed. According to the results, enhanced UV-B radiation has a significant effect on the early development of L. japonica under laboratory conditions, suggesting that the UV-B radiation could not be overlooked as one of the important environmental factors influencing the ontogeny of macroalgae living in marine ecosystems. Supported by the Program for New Century Excellent Talents in University (Grant No. NCET-05-0597) and National Natural Science Foundation of China (Grant No. 30270258)  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