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 15 毫秒
1.
In this article, we attempt to estimate the contemporary phytoplankton species pool of a particular lake, by assessing the rate of floral change over a period of 15 years. Phytoplankton time series data from Lake Stechlin, an oligo-mesotrophic lake in the Baltic Lake District (Germany) were used. Of the 254 algal species recorded during the 15-year of studies with roughly biweekly sampling, 212 species were planktonic. In the individual plankton years, the recorded total number of species changed between 97 and 122, of which the number of dominants (>1% contribution to the annual average of total biomass) was only 10–19. The 15-year cumulative number of species exhibited an almost linear increase after an initial saturation phase. This increase was attributed to two reasons: increase of sample size and immigration of species new to the flora. Based on a probabilistic model developed in this study, we estimated the number of co-existing planktonic species of the lake as some 180, and the rate of floral change as 1–2 species per year. Of these co-existing species, only few maintain the matter–energy processing ecosystem functions in any particular plankton year. Selection of these dominants is probably driven by mesoclimatic cycles, coupled with human-induced forcing, like eutrophication. All others are hiding as an ecological memory, in the sense of the capacity or experiences of past states to influence present or future responses of the community. Data analyses suggest that selection of the ‘memory species’ that show temporary abundance increases over shorter (several years) periods are largely dependent upon the dominants. These results show that interspecific interactions and the particular autecological features of the dominants, together with their effects on the whole ecosystem, act as a major organizing force. Some phytoplankton species, like Planktothrix rubescens, are efficient ecosystem engineers with cascading effects of both a top-down and bottom-up nature. Historical scientific data on Planktothrix blooms in Lake Stechlin suggest cyclic patterns in long-term development of phytoplankton which, as the legend of the Red Cock suggests, dates back much further than scientific archives.  相似文献   

2.

Background

The extraction of salt from seawater by means of coastal solar salterns is a very well-described process. Moreover, the characterization of these environments from ecological, biochemical and microbiological perspectives has become a key focus for many research groups all over the world over the last 20 years. In countries such as Spain, there are several examples of coastal solar salterns (mainly on the Mediterranean coast) and inland solar salterns, from which sodium chloride is obtained for human consumption. However, studies focused on the characterization of inland solar salterns are scarce and both the archaeal diversity and the plant communities inhabiting these environments remain poorly described.

Results

Two of the inland solar salterns (termed Redonda and Penalva), located in the Alto Vinalopó Valley (Alicante, Spain), were characterized regarding their geological and physico-chemical characteristics and their archaeal and botanical biodiversity. A preliminary eukaryotic diversity survey was also performed using saline water. The chemical characterization of the brine has revealed that the salted groundwater extracted to fill these inland solar salterns is thalassohaline. The plant communities living in this environment are dominated by Sarcocornia fruticosa (L.) A.J. Scott, Arthrocnemum macrostachyum (Moris) K. Koch, Suaeda vera Forsk. ex Gmelin (Amaranthaceae) and several species of Limonium (Mill) and Tamarix (L). Archaeal diversity was analyzed and compared by polymerase chain reaction (PCR)-based molecular phylogenetic techniques. Most of the sequences recovered from environmental DNA samples are affiliated with haloarchaeal genera such as Haloarcula, Halorubrum, Haloquadratum and Halobacterium, and with an unclassified member of the Halobacteriaceae. The eukaryote Dunaliella was also present in the samples.

Conclusions

To our knowledge, this study constitutes the first analysis centered on inland solar salterns located in the southeastern region of Spain. The results obtained revealed that the salt deposits of this region have marine origins. Plant communities typical of salt marshes are present in this ecosystem and members of the Halobacteriaceae family can be easily detected in the microbial populations of these habitats. Possible origins of the haloarchaea detected in this study are discussed.  相似文献   

3.
Breeding systems were evaluated for 51 plant species according to life form, pollination system, vegetation type, and phenology, in the coastal plain of Paraguaná Peninsula, Venezuela. Sexual systems were no associated to life form, pollination system, vegetation type, and phenology. The frequency distribution of sexual system was 82.3% hermaphroditism, 15.6% monoecy, and 1.9% dioecy. All sexual systems had a peak during the lowest rainfall. Genetic system distribution was 64.8% self-compatibility (including partially self-compatibility) and 35.2% self-incompatibility. Among self-compatible species, 45.1% were autogamous (19.6% not autogamous). The genetic systems were associated significantly to: (1) plant life form: self-compatible species tend to be herbaceous and self-incompatible plants tend to be woody species; (2) vegetation type: self-compatible species were predominant in the three vegetation types, but in the mangrove the frequency of self-compatible and self-incompatibles was similar; and (3) pollination system: most of the self-compatible species were polyphilous. Mating systems: xenogamous and autogamous species were associated only with plant life forms. Xenogamous plants were mostly woody species and autogamous plants were mostly herbaceous species. The high incidence of hermaphroditism, self-compatibility, and autogamy are related to herbaceous life form, polyphilous pollination system, and climatic conditions, together the insularity of the Paraguaná peninsula.  相似文献   

4.
5.
1. Starvation for 3 days produces a decrease in methaemoglobin-reductase and glutathione-reductase activities, but it does not alter the glucose 6-phosphate-dehydrogenase activity of the rat erythrocyte. 2. The feeding of a protein-free diet for 11 days causes greater changes in the first two enzymes and also a diminution of the third. Under this experimental condition slight decreases in protein and haemoglobin contents were noted. 3. The experimental animals did not show methaemoglobinaemia, probably because the activity of methaemoglobin diaphorase is preserved. 4. The GSH content was not affected but the stability of the tripeptide in the presence of an oxidizing agent was diminished.  相似文献   

6.
Studies of floral scent generally assume that genetic adaptation due to pollinator-mediated natural selection explains a significant amount of phenotypic variance, ignoring the potential for phenotypic plasticity in this trait. In this paper, we assess this latter possibility, looking first at previous studies of floral scent variation in relation to abiotic environmental factors. We then present data from our own research that suggests among-population floral scent variation is determined, in part, by environmental conditions and thus displays phenotypic plasticity. Such an outcome has strong ramifications for the study of floral scent variation; we conclude by presenting some fundamental questions that should lead to greater insight into our understanding of the evolution of this trait, which is important to plant-animal interactions.Key words: abiotic factors, aromatics, floral scent, GxE interaction, phenotypic plasticity, pollination, terpenoids, volatilesFloral scent is thought to function as a major non-visual attractive cue for many pollinators in a large number of plant systems1,2 and therefore most research on this plant trait has proceeded in the context of pollination ecology. Such studies have revealed the physiological and behavioral responses of pollinators to various floral volatiles (reviewed in refs. 3 and 4), convergent evolution of odor phenotypes attractive to specific pollinator classes (reviewed in refs. 5 and 6), reproductive isolation of plant species due to differences in pollinator attraction by scent,7 and instances of deception in which flowers mimic insect pheromones to effect pollination.8 Together, this body of evidence suggests that specific floral scent profiles can have important implications for the reproductive potential of many plant species.This pollinator-centered viewpoint has carried through to research on floral scent variation, including our most recent work on the insect-pollinated species Hesperis matronalis (Brassicaceae).9 Such studies usually suggest that the floral scent variation commonly found within and among individuals, populations and species (reviewed in ref. 2) is due to genetic differentiation as a result of selection by pollinators over time (reviewed in ref. 10). But an organism''s genes are only one factor determining phenotype. Both biotic (living) and abiotic (non-living) environmental conditions can profoundly affect phenotype expression, leading to significant variation. For plants, abiotic factors such as climate and soil chemistry can have particularly strong effects on phenotypes. When these environmental conditions cause changes in phenotype, we would say that a trait displays phenotypic plasticity.1113 A number of studies have uncovered phenotypic plasticity for many different plant traits.12 However, while phenotypic variation in floral scent has been well-documented1,2 and correlated with variation in biotic factors like pollinator behavior,1417 these studies were decidedly focused on natural selection, rather than phenotypic plasticity, as an organizational framework.However, in examining the scientific literature on floral scent, we found four studies in which the effects of naturally variable abiotic factors on floral scent profiles were examined, three of which were performed by the same research group (1821 (21). Moreover, these studies are decidedly not analyzed and interpreted using standard protocols for phenotypic plasticity studies.13

Table 1

A survey of previous studies examining changes in floral scent phenotype due to abiotic factors
StudySpeciesEnvironmental characteristicPlant materialStudy locationChange in volatile emissions?Direction of change
Loper and Berdel 1978Medicago sativa L.IrrigationClonesExperimental farmNon/a
CuttingClonesExperimental farmNon/a
Hansted et al. 1994Ribes nigrumTemperatureTwo varietiesGrowth chamberYes+ temperature, + ER*
Jakobsen and Olsen 1994Trifolium repens L.TemperatureCultivarGrowth chamberYes+ temperature, + ER
IrradianceCultivarGrowth chamberYes+ irradiance, + ER
Air HumidityCultivarGrowth chamberYes+ humidity, − ER
Nielsen et al. 1995Hesperis matronalis L.TemperatureWild seedsGrowth chamberYes+ temperature,
+ monoterpene ER
This study, 2009Hesperis matronalisGrowingWild plantsWild vs.YesWild—different ER,
EnvironmentCommon GardenSC between populations;
Garden—similar ER,
SC between populations
Open in a separate window*Plus signs indicate a numerical increase, minus signs indicate a decrease; ER = floral scent emission rate, SC = scent composition.Research we have conducted in conjunction with our recently published work on the floral scent of H. matronalis9 suggests that some of the natural variation in the odor of this species may be attributable to phenotypic plasticity. We reared potted H. matronalis rosettes from two populations (PA1 and PA2) in northwestern Pennsylvania in a common garden environment and upon flowering, collected scent from these individuals using dynamic headspace extractions (reviewed in ref. 9). We then compared floral scent composition and emission rates of potted plants with each other (between populations in a common garden), as well as with the floral scent profiles of plants reared in their source population (i.e., between individuals from the same population reared in different environments). The results were striking. Analysis of scent composition using non-metric multidimensional scaling and analysis of similarity (NMDS and ANOSIM, respectively: reviewed in ref. 9) suggested that the scent composition of plant populations reared in their native environments differ significantly from each other in terms of two major biosynthetic classes of volatiles—aromatics and terpenoids (Fig. 1, filled symbols only). This was especially true for the aromatic eugenol and derivatives of the terpenoid linalool (furanoid linalool oxides and linalool epoxide). In contrast, common-garden reared plants from different populations did not differ in floral scent composition, regardless of their original source population. Perhaps even more interestingly, while both populations showed changes due to rearing environment, the degree of change differed: in only one population (PA1) did scent composition change significantly between native and garden reared plants (Fig. 1).Open in a separate windowFigure 1NMDS (non-metric multidimensional scaling) plots of scent composition for purple morphs from two populations of Hesperis matronalis—(A) Aromatics and (B) Terpenoids. Filled symbols represent scent from home environment in situ plants, which are significantly different from one another as determined by analysis of similarity (ANOSIM: aromatics—p = 0.03, R = 0.22; terpenoids—p = 0.01, R = 0.25). Open symbols represent scent from plants reared in a common environment. Population PA1 is represented by triangles and population PA2 is represented by squares. Arrows indicate the direction of shift from home environment to common garden floral scent composition; black arrows represent a significant difference between groups determined by ANOSIM (Aromatics—p = 0.01, R = 0.30; Terpenoids—p = 0.06; R = 0.20) and gray arrows represent a non-significant difference.Floral scent emission rate also showed environmentally induced differences. While wild plants from our two populations differed significantly in the amount of scent emitted in situ, with PA1 emitting more total scent, total aromatics and total terpenoids,9 we found that rearing plants from these sites in a common garden environment either significantly reverses the direction of differences in emission rates seen between natural populations, with PA2 now emitting more aromatic scent (Analysis of Variance: F = 4.09; p = 0.05; Fig. 2A), or homogenizes the quantity of scent emitted (i.e., no significant differences in emission rates between populations; Fig. 2B and C).Open in a separate windowFigure 2Box plots of scent emission rates for purple Hesperis matronalis plants grown in common garden environments in terms of (A) Aromatics, (B) Terpenoids and (C) Total Scent. The edges of each box represent the range of data between the 25th percentile and the 75th percentile, while the horizontal bar indicates the median for each population. The error bars on each box extend to the 5th and 95th percentile of the data range respectively. To the right of each box plot, the mean is presented as a horizontal line, with standard error bars. Mean values not sharing letters are significantly different as determined by analysis of variance (ANOVA).Together, these results suggest that rearing environment can have a profound effect on floral scent composition and emission rate, such that plants from the same maternal environment can have radically different floral scent phenotypes in response to differential growing conditions. If our work effectively incorporates a random genetic sample from each population into each growing environment, then at least some of the phenotypic variation we describe here could be interpreted as phenotypic plasticity. This experiment does not allow us to pinpoint the exact environmental conditions associated with phenotypic differences in floral scent (although variation in nutrient or water availability between wild and common-garden settings is likely), nor does it completely conform to the traditional “reactionnorm” studies associated with plasticity research which would allow detection of genetic variation in scent plastiticy.12,13 However, our results suggest that floral scent of plants grown in wild populations may be plastic, which provides some additional insight into our recently published work uncovering significant among-population variation in floral scent.9 For researchers that study phenotypic plasticity, such an outcome is probably not a surprise, nor is our finding that populations respond differently to environmental conditions (i.e., potential GxE interaction, reflecting genetic variability in plasticity).However, if floral scent can be plastic, this raises a number of biologically relevant questions that should be addressed in floral scent research, including: (1) Is there truly a canonical floral scent blend that can be attributed to a given plant species, as is normally supposed by those studying floral scent from an evolutionary perspective? (2) Which environmental conditions exert the strongest influence on floral scent profiles in a species? (3) How do such conditions interact with genetic variation in the factors responsible for scent biosynthesis and emission? (4) Are floral scent profiles plastic within a single flowering period; if so, what impact does this have on pollinator behavior and therefore plant fitness? (5) At what scale do biotic agents such as pollinators and herbivores respond to quantitative and qualitative variation in floral scent? Studies that address these questions should lead us to a more mature understanding of the causes and consequences of natural variation in floral scent.  相似文献   

7.
Insect attack can have major consequences for plant population dynamics. We used individually based simulation models to ask how insect oviposition behaviour influences persistence and potential stability of an herbivore–plant system. We emphasised effects on system dynamics of herbivore travel costs and of two kinds of behaviour that might evolve to mitigate travel costs: insect clutch size behaviour (whether eggs are laid singly or in groups) and female aggregation behaviour (whether females prefer or avoid plants already bearing eggs). Travel costs that increase as plant populations drop lead to inverse density dependence of plant reproduction under herbivore attack. Female clutch size and aggregation behaviours also strongly affect system dynamics. When females lay eggs in large clutches or aggregate their clutches, herbivore damage varies strongly among plants, providing probabilistic refuges that permit plant reproduction and persistence. However, the population dynamics depend strongly on whether insect behaviour is fixed or responds adaptively to plant population size: when (and only when) females increase clutch size or aggregation as plants become rare, refuges from herbivory weaken at high plant density, creating inverse density dependence in plant reproduction. Both herbivore travel costs themselves, and also insect behaviour that might evolve in response to travel costs, can thus create plant density dependence—a basic requirement for regulation of plant populations by their insect herbivores.  相似文献   

8.
Many Acacia species in arid areas of eastern Australia have been severely impacted by grazing, habitat degradation and fragmentation. These factors have been at the core of proposed explanations for the reproductive failure and numerical decline of Acacia carneorum and other threatened acacias. Paradoxically, the sympatric Acacia ligulata is thriving and highly fecund. Although these species have superficially similar floral displays, differences in sexual reproductive success may reflect interactions between flower and inflorescence ontogeny and pollinator assemblages. We compared the floral biology and flower visitor assemblages of A. carneorum and A. ligulata at four sites per species. Both species displayed similar floral ontogeny and synchronicity of display, with inflorescences simultaneously hermaphroditic for 4–5 days. However, A. ligulata displayed a higher density of flowers than A. carneorum and, while both species received a range of flower visitors, A. ligulata was visited by relatively few species and was serviced primarily by the non-native honeybee Apis mellifera, which typically made many within-plant movements during foraging bouts. In contrast, A. carneorum was visited by a diverse suite of native insects that carried little pollen and made fewer within plant movements. On average, Apis mellifera carried 98.4 % A. ligulata pollen, whereas the native insect visitors of A. carneorum carried only 45 % A. carneorum pollen. Differing floral ontogeny or lack of native pollinators does not explain the reproductive failure of A. carneorum. The success of A. ligulata may reflect pollination services provided by A. mellifera and interactions with plant mating systems.  相似文献   

9.
Florigen is a mobile signal released by the leaves that reaching the shoot apical meristem (SAM), changes its developmental program from vegetative to reproductive. The protein FLOWERING LOCUS T (FT) constitutes an important element of the florigen, but other components such as sugars, have been also proposed to be part of this signal.1-5 We have studied the accumulation and composition of starch during the floral transition in Arabidopsis thaliana in order to understand the role of carbon mobilization in this process. In A. thaliana and Antirrhinum majus the gene coding for the Granule-Bound Starch Synthase (GBSS) is regulated by the circadian clock6,7 while in the green alga Chlamydomonas reinhardtii the homolog gene CrGBSS is controlled by photoperiod and circadian signals.8,9 In a recent paper10 we described the role of the central photoperiodic factor CONSTANS (CO) in the regulation of GBSS expression in Arabidopsis. This regulation is in the basis of the change in the balance between starch and free sugars observed during the floral transition. We propose that this regulation may contribute to the florigenic signal and to the increase in sugar transport required during the flowering process.  相似文献   

10.
The time course of changes in the dopamine concentration in dopaminergic neurons of the nigro-neostriatal and mesolimbic systems of the rat brain during 1 h after intraperitoneal injection of -phenylethylamine (100 mg/kg) was studied by quantitative fluorescence-histochemical analysis. The results showed that -phenylethylamine causes a marked fall in the dopamine level in neurons of dopaminergic systems of the brain. The dopamine level in the bodies of dopaminergic neurons changes more than in their axon terminals. The fall in the dopamine concentration in the dopaminergic systems of the brain during the first hour is irregular in character: in the terminals between 10 and 30 min and in the bodies between 30 and 45 min there is actually a temporary increase in the dopamine concentration. The rise in the dopamine concentration in the terminals coincides with a sharp fall in the dopamine level in the neuron bodies, and conversely, the fall in the dopamine concentration in the terminals after 30 min is accompanied by some increase in the dopamine concentration in the neuron bodies. The results suggest that the increase in motor activity described in the literature in animals after injection of -phenylethylamine is connected with its action on catecholaminergic, especially dopaminergic, brain systems.Institute of Biophysics, Academy of Sciences of the USSR, Pushchino-on-Oka. Translated from Neirofiziologiya, Vol. 11, No. 6, pp. 578–584, November–December, 1979.  相似文献   

11.
Arundina graminifolia is an early successional plant on Iriomote Island, the Ryukyus, Japan, where it is endangered. Populations flower for more than half a year, and many inflorescences bloom for one to several months. The nectarless gullet flowers, which open for up to six days, are self-compatible but cannot self-pollinate spontaneously; thus they rely on pollinating agents for capsule production. Field observations at two habitats identified at least six species of bees and wasps, primarily mate-seeking males of Megachile yaeyamaensis and Thyreus takaonis, as legitimate pollinators. Thus, this orchid is a pollinator generalist, probably owing to its long blooming period and simple flower morphology. Carpenter bees, which were previously reported to pollinate this orchid, frequently visited flowers but were too large to crawl into the labellum chamber and never pollinated the flowers. Extrafloral nectaries on inflorescences attracted approximately 40 insect taxa but were not involved with pollination. Fruit-set ratios at the population level varied spatiotemporally but were generally low (5.2–12.4 %), presumably owing to infrequent flower visits by mate-seeking pollinators and the lack of food rewards to pollinators.  相似文献   

12.
13.
The Southern Espinha?o Range consists of large areas covered by quartzitic or metaliferous tropical altitudinal fields. The Espinha?o Range ecosystems are endangered by anthropic high impacts, particularly due to mining and urbanization. We conducted a one-year inventory of the bee flora and fauna at the quartzitic Ouro Branco Mountains and a two-year survey of the metaliferous Ouro Preto fields. The samples were collected twice a month, from 8:00 am to 5:00 pm. The bees (677) belonged to 91 species, five families. The family Apidae was the richest and most abundant, followed by the Halictidae and Megachilidae. The bees visited 46 flowering plant species; the most visited plants were the Asteraceae (n = 220), the Malpighiaceae (n = 95), the Melastomataceae (n = 94), the Fabaceae (n = 78), and the Solanaceae (n = 63). Diversity was higher in Ouro Branco (H = 1.47) than in Ouro Preto (H = 1.17). The low richness and abundance of bees in our research site when compared to other Brazilian "Cerrado" areas can be due to the high altitude, low temperature, and low availability of flowers we found. "Canga" and rupestrian areas house fauna and flora species that are rare and threatened by extinction. The southern Espinha?o areas can, therefore, be given the status of permanent biodiversity preservation area.  相似文献   

14.
ω-Transaminase (TA) catalyzed asymmetric syntheses of amines were carried out in the one enzyme systems with wild-type enzymes (S)-TA from Pseudomonas aeruginosa, (S)-TA from Paracoccus denitrificans and (R)-TA from Aspergillus terreus. The scope of amine donors and aromatic carbonyl substrates was thoroughly explored. Among the range of potential amino donors, 2-propylamine, 2-butylamine and 1-phenylethylamine were found as promising candidates, which gave superior conversions in the amination reactions compared to other donors. Various prochiral aromatic ketones were accepted as substrates by the investigated enzymes. In most cases, good to excellent conversions (up to 98%) to the amine products with excellent e.e.-values (>99.9% for (S) or (R)) were obtained by the action of a single enzyme and an appropriate amino donor. (S)-TA from Paracoccus denitrificans was found to accept bulky ketones, e.g. 1-indanone, α- and β-tetralone or 2-acetonaphthone, in the asymmetric amination. In some cases the enantiomeric excesses in the amination reactions were dependent on the amino donor. Moreover, the influence of the pH, temperature and cosolvents on the outcome of reactions was additionally investigated.  相似文献   

15.
The karstic nature of the Yucatan Peninsula allows the formation of natural sink-holes from the dissolution of calcareous rock. These systems are almost the only epigean source of fresh water available in this region. In spite of their biological importance, little is known about the morphometric and limnologic characteristics of these karstic systems. We measured limnological variables in eight cenotes in central Quintana Roo during February–May, 2001. Zooplankton biomass and chlorophyll a were also measured in order to determine if the behavior of primary and secondary production was related to environmental parameters. Important short-term changes were observed in nutrients (NO3 , NO2 , PO4 3-), biomass, and chlorophyll a. The morphometrically conditioned productivity (MCP), which evaluates the cumulative effect of several morphometric variables on production (area, maximum length, shoreline development, perimeter), showed a negative correlation with respect to zooplankton biomass, as did also both pH and temperature. Conversely, NO3 and NO2 had a positive correlation with zooplankton biomass. No correlation was found for chlorophyll a. Significant differences in NO3 (F = 61.52, p<0.001), NO2 (F = 7.36, p<0.001), zooplankton biomass (F = 17.57, p<0.001), chlorophyll a (F = 62.19, p<0.001), and conductivity (F = 497.49, p<0.001) were found among the systems. These results indicate the existence of sharp differences between these karstic systems (oligotrophic, with smaller area, deep and less productive) and non-karstic ones, (eutrophic, larger area, shallow and more productive) but are similar to previous data from other karstic systems of Mexico and other parts of the world. However, understanding of these fragile tropical systems is in the initial phase. It is necessary to increase the intensity of these studies in order to allow a full explanation of their limnological behavior.  相似文献   

16.
The meiofauna from seagrass meadows in the western sector of the Gulf of Batabanó, Cuba were studied to describe the spatial and temporal variations in community structure. Replicated cores were taken in three locations (arranged in m- and km-scales) and in two seasons (dry and wet). The meiofauna (metazoans between 500 and 45 microm) were identified to major taxa. Temporal changes in the meiofaunal communities could not be detected and they are not linked to the subtle seasonal changes in the water column. A larger variation in community structure was observed in the spatial m-scale (among cores in a station) probably accredited to heterogeneity of microenvironment and biological processes. A second source of variation in the km-scale (among locations) was identified relating to physical processes affecting seagrass meadows: marine currents and anthropogenic disturbances. Distribution patterns of meiofauna across locations coincide with one study from 20 years ago in seagrass beds (i.e. higher densities in area closer to break-shelf and diminution of fauna at southern of Pinar del Rio); however, cumulative anthropogenic disturbances on seagrass meadows would most likely explain the depletion of communities observed in our survey in comparison with decades ago. Estimates of meiofaunal density and richness of major taxa from our study (and other areas from the Cuban shelf) are consistently lower than other temperate and tropical sites; possibly caused by low primary productivity due to narrow tidal amplitude and oligotrophic waters.  相似文献   

17.
BackgroundInsect-vectored Leishmania are responsible for loss of more disability-adjusted life years than any parasite besides malaria. Elucidation of the environmental factors that affect parasite transmission by vectors is essential to develop sustainable methods of parasite control that do not have off-target effects on beneficial insects or environmental health. Many phytochemicals that inhibit growth of sand fly-vectored Leishmania—which have been exhaustively studied in the search for phytochemical-based drugs—are abundant in nectars, which provide sugar-based meals to infected sand flies.Principle findingsIn a quantitative meta-analysis, we compare inhibitory phytochemical concentrations for Leishmania to concentrations present in floral nectar and pollen. We show that nectar concentrations of several flowering plant species exceed those that inhibit growth of Leishmania cell cultures, suggesting an unexplored, landscape ecology-based approach to reduce Leishmania transmission.SignificanceIf nectar compounds are as effective against parasites in the sand fly gut as predicted from experiments in vitro, strategic planting of antiparasitic phytochemical-rich floral resources or phytochemically enriched baits could reduce Leishmania loads in vectors. Such interventions could provide an environmentally friendly complement to existing means of disease control.  相似文献   

18.
Complex formation reactions of phenylboronic, phenylphosphonic, phenylarsonic and 4-aminophenyl arsonic acids with β-cyclodextrin (cycloheptaamylose, β-CD) and some simple carbohydrates (mannitol, sorbitol, glucose) have been studied using spectrophotometric, potentiometric methods and solubility measurements, supplemented with HPLC and IR analyses of the solid samples. Equilibrium constants have been determined at ionic strength of 0.2 M (NaCl) and 25 °C. β-CD forms the most stable complexes with the neutral, undissociated forms of the acids, the stability constants are as follows: phenylboronic acid: 320 ± 36, phenylphosphonic acid: 108 ± 25, phenylarsonic acid: 97 ± 4 and 4-aminophenyl arsonic acid: 107 ± 10. The stability constants for the β–CD-complexes of the ionic forms are much lower. Ternary complexes of low stability could be detected in the case of phenylphosphonic acid and sorbitol with the undissociated form and with glucose and the dianion. In more concentrated solutions phenylboronic acid forms insoluble complexes with mannitol, sorbitol and β-CD. The solid phases obtained in the ternary systems are predominantly mixtures of ester type 3:1 complexes with the carbohydrate and 1:1 inclusion complex with the β-CD. No significant interaction has been found with glucose. The phenomena can be explained by the differences in the structures of the components and by the changes in the H-bonding network of β-CD on the complex formation.  相似文献   

19.
Autogamously self-fertilizing taxa have evolved from outcrossing progenitors at least 12 times in the annual wildflower genus, Clarkia (Onagraceae). In C. xantiana, individuals of the selfing subspecies (ssp. parviflora) flower at an earlier age, produce successive flowers more rapidly, and produce flowers that complete their development more rapidly than their outcrossing counterparts (ssp. xantiana). Two hypotheses have been proposed to explain the joint evolution of these whole-plant and individual floral traits. The accelerated life cycle hypothesis proposes that selection favoring a short life cycle in environments with short growing seasons (such as those typically occupied by parviflora) has independently favored genotypes with early reproduction, synchronous flower production, and rapidly developing, self-fertilizing flowers. The correlated response to selection hypothesis similarly proposes that selection in environments with short growing seasons favors early reproduction, but that rapid floral development and increased selfing evolve as correlated responses to selection due to genetic linkage (or pleiotropy) affecting both whole-plant and floral development. We conducted a greenhouse experiment using maternal families from two field populations of each subspecies to examine covariation between floral and whole-plant traits within and among populations to seek support for either of these hypotheses. Our results are consistent with the accelerated life cycle hypothesis but not with the correlated response to selection hypothesis.  相似文献   

20.
Decreases in pollinator abundance may particularly constrain plants that lack floral rewards, since they are poor competitors for pollinators in the plant community. Here, we documented the pollination ecology of a rewardless orchid, Calanthe reflexa Maxim., and examined effects of forest understory degradation by deer browsing on pollination success of the species in the light of a change in the abundance of neighboring flowering plants in 2010 and 2011. Bombus species were the only pollinators at each site and the flowering phenology of C. reflexa did not overlap with that of other rewarding plants. Pollinator visit rates (assessed by time‐lapse photography), and pollinia removal rate were higher in the undegraded understory site than the degraded site in both years, while the fruit set ratio did not differ between the sites in 2011. Coverage by neighboring flowering plants was extremely low in the degraded site. Our results suggest that, although its flowering phenology and consequently lower interspecific competition of C. reflexa with rewarding plants for attracting bumblebees, neighboring flowering plants may play an important role for maintaining the visitation frequency of bumblebees of C. reflexa and contribute to its pollination success.  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