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.
Taber D. Allison 《Oecologia》1992,89(2):223-228
Summary Browsed Canada yew (Taxus canadensis) populations have a higher proportion of males and a lower proportion of monoecious plants than unbrowsed yew populations. The proportion of monoecious plants increases with time following protection from browsing suggesting that deer browsing causes male-biased sex expression in Canada yew. In contrast, results from comparing browsed and unbrowsed populations, exclosure studies, and browse simulation experiments indicate that strobilus ratios and phenotypic gender of browsed yews may be female-biased. In part, these results correspond to the influence of size on sex expression in Canada yew; small yews tend to be male, but if monoecious, have female-biased strobilus ratios. Large yews are monoecious, but have male-biased strobilus ratios. There is, however, no consistent relationship between size and gender in Canada yew, suggesting that in some circumstances, yews shift allocation to female function in response to browsing.  相似文献   

2.
Taber D. Allison 《Oecologia》1990,83(4):523-529
Summary Canada yew (Taxus canadensis) populations currently browsed by white-tailed deer (Odocoileus virginianus) or browsed by deer in the past had significantly lower production of male strobili, female strobili, and seeds than unbrowsed yew populations. Exclosure studies showed that protected yews produced significantly more male and female strobili than unprotected yews, but only after several years of protection. Seed production did not respond as readily to protection from deer perhaps because of reduced pollination levels in browsed yew populations. Previously unbrowsed yews were clipped at different levels of removal of available browse (control (no removal), 25%, 50%, 75%, and 100% removal) to simulate deer browsing. Reduction in male strobilus production was linearly related to clipping intensity in three years of observation. Female strobilus production was significantly reduced only at the 100% level of removal. Intermediate levels of clipping may have even stimulated production of female strobili. Analysis of covariance, with previous year's branch production as the covariate, showed no significant effect of clipping on male strobilus production except in the 100% removal group. Female strobilus production showed no such covariance with branch production. Effects of clipping on seed production could not be reliably assessed in 1984 and 1985 due to low seed production. Seed production in 1986 was significantly reduced only in the 100% removal group. Field observations of deer browsing of Canada yew indicate that 100% levels of removal are typical of natural levels of browsing.  相似文献   

3.
The dioecious species Urtica dioica harbours wide variation in sex ratio of seeds. We conducted a series of crosses to analyse the genetic basis of sex determination in this species. Dutch populations of U. dioica contain low proportions of monoecious individuals beside male and female plants. Self-pollination of monoecious plants always yielded female, male and monoecious plants, generally in a ratio of one female to three male/monoecious individuals. This motivated us to write down a simple model in which gender is determined by one major sex-determination locus with four alleles. In the model males and monoecious plants have distinct genotypes but are both heterozygous at the sex-determination locus. We first made crosses among progeny obtained after self-pollination of monoecious plants. These crosses showed that the monoecious trait generally showed Mendelian inheritance and was passed on to the next generation via both pollen and seeds. Further crosses between monoecious plants and plants from dioecious system indicated that alleles from the dioecious system are often dominant. However, many exceptions to our genetic model are observed which suggest that dominance is incomplete and/or that more genes are involved in sex determination. We discuss to what extent sex determination genes explain the strongly biased seed sex ratios and argue that additional genes, for instance genes for female choice, must also be involved.  相似文献   

4.
Gender expression was recorded for three geographically distinctpopulations and two stand densities of pinsapo fir (Abies pinsapo) in southern Spain, during 4 consecutive years (1990–1993).During this period the trees only flowered in 1991 and 1992,and thus only two cone crops could be quantified. No significantgeographical variation in flowering intensity or in cone cropproduction was found. The most extreme variations were correlatedwith stand density: flowering intensity and cone crop were alwaysgreater in low density trees.Abies pinsapo is a monoecious speciesbut all the populations studied showed a predominance of femaleplants, with a few male and monoecious individuals, thus indicatinga functionally subdioecious breeding system. No differencesin gender expression between populations were recorded. Differencesin sex expression of individuals at two different stand densitieswere found, but these were only significant in 1991. In densestands of pinsapo firs branches were limited to the top of treesand thus, they produced mainly female cones, whereas isolatedtrees had branches from ground level and as a consequence theywere predominantly male, with a higher total reproductive effort.The results indicate that relative investment in male and femalereproductive structures byA. pinsapo individuals is a responseto plant architecture, plant resource status and environmentalvariation. Abies pinsapo ; plant density; gender variation; reproductive biology; monoecy; Mediterranean fir  相似文献   

5.
Urtica dioica is a sub-dioecious plant species, i.e. males and females coexist with monoecious individuals. Under standard conditions, seed sex ratio (SSR, fraction of males) was found to vary significantly among seed samples collected from female plants originating from the same population (0.05–0.76). As a first step, we investigated the extent to which SSR and sex expression of male, female, and monoecious individuals is influenced by external factors. We performed experiments to analyse: (1) whether the environment of a parental plant affects the sex ratio (SR) of its offspring, (2) whether SSR can be affected by environmental conditions before flowering, and (3) whether sex expression of male, female and monoecious plants that have already flowered can be modified by environmental conditions or by application of phyto-hormones. Within the range of our experimental design, SSR was not influenced by external factors, and gender in male and female plants was stable. However, sex expression in monoecious plants was found to be labile: flower sex ratio (FSR, fraction of male flowers) differed considerably between clones from the same individual within treatments, and increased toward 100% maleness under benign conditions. These results provide strong evidence that monoecious individuals are inconstant males, which alter FSR according to environmental circumstances. In contrast, we consider sex expression in male and female individuals to be solely genetically based. The observed variation in SSR between maternal parents cannot be explained by sex-by-environment interactions.  相似文献   

6.
Monoecy allows high plasticity in gender expression because the production of separate female and male flowers increases the ability to respond to specific environmental circumstances. We studied variation in sexual expression and its correlates in the monoecious shrub Buxus balearica, for two years, in six populations in the Balearic Islands and four in the Iberian Peninsula. Phenotypic gender varied among populations; while island populations showed slight variations around an average gender, mainland populations showed a broad range of variation in gender among individuals, always biased towards increasing maleness compared with the other populations. Within populations, gender was not related to plant size. Between-year changes were slight and mainly consisted of an increase in relative maleness in the mast year. Reproduction did not affect gender in the next year, as assessed by either observational or experimental methods. Most variation in gender expression occurred among individuals within populations (83.6 %), followed by variation among populations (13.6 %) and years (2.8 %). Our results suggest that male-biased gender at population and plant levels was related to stressful conditions and resource limitation, because: (1) maleness was higher in mainland populations, where summer drought was stronger; (2) maleness increased with elevation; (3) fruit set was positively correlated with femaleness; (4) the percentage of male inflorescences increased over the flowering period; and (5) male inflorescences were preferentially in lower parts of the branch nodes. Higher maleness in mast years, however, could be related to increased male success under synchronic flowering.  相似文献   

7.
Myriophyllum ussuriense has been described as dioecious but monoecious plants were newly found from some populations in south-western Japan. Sex expression of monoecious plants proved labile and they sometimes bore male or female flowers alone. On the other hand, sex expression of dioecious plants was stable and seemed to be fixed genetically. M. ussuriense may be still in the course of differentiation from monoecy to dioecy. Received 13 April 2001/ Accepted in revised form 22 May 2001  相似文献   

8.
Summary Flowering individuals of dwarf ginseng may be either male or hermaphroditic. I recorded the sex expression and size of individuals in three populations for three or four years in order to 1) determine whether this bimodal distribution of sex expression was due to sex changing or genetic dimorphism, and 2) test predictions about a) the relationship between size and gender, and b) the association of size change and sex change. Twenty five to 37% of the flowering individuals in each population changed gender from one year to the next. Of the plants I followed for four years, 83% changed sex and 57% changed more than once. In each of these populations as well as two others, hermaphrodites were significantly larger than males. Gender dynamics of the three populations differed, but hermaphrodites tended to become smaller and were more likely to change gender than remain hermaphroditic the following year, whereas males tended to grow larger and were more likely to remain male than to change gender. Dwarf ginseng is clearly a diphasic (sex changing) species in which sex expression is determined primarily by size. A difference between genders in the immediate resource costs of reproduction appears to be an important determinant of sex change and gender phase ratios in populations.  相似文献   

9.
To explain the floral sex ratio strategy in wind-pollinated monoecious species, we developed four models with special reference to wind-pollination efficiency (WPE) and competitive sharing among male flowers (CSM). WPE is a function that follows a Poisson distribution and explains the frequency of seeds fertilized by an individual via wind-pollination, whereas CSM is defined by the sharing of female flowers among male flowers within the local breeding population. We argued the applicability of the results to the actual tendencies observed in wind-pollinated monoecious species and found that a game model with WPE and CSM was the most applicable. The model predicted that individuals should change their gender expression in the following order: female phase (female flowers only), male phase (male flowers only), and constant male phase (individuals constantly allocate reproductive resources to male flowers, and remaining resources to female flowers), with increasing reproductive resources. However, the trend is likely to be influenced by the variation in the reproductive investment among individuals and the degree of WPE. Thus, large variation and low pollination efficiency enable three phases to co-occur within a population. Actual trends in real populations correspond to our prediction.  相似文献   

10.
Arisaema species exhibit gender diphasy, or sex change, where individual plants produce either male, monoecious or female inflorescences depending on their size. Three basic sex-change patterns have been described in Arisaema. Type I species change between male and monoecious phases, type II species change between male, monoecious and female phases, while type III species change between male and female phases. Theoretical models suggest that sex ratios should be biased toward males, the sex with the lowest cost of reproduction. The goal of this study was to examine sex-ratio variation among Arisaema species that differ in sex-change patterns. Data from an extensive literature review, consisting of all available studies reporting Arisaema sex ratios, were combined with data from extensive field surveys of Arisaema dracontium and Arisaema triphyllum in southern Indiana, USA. This data set contains nearly 30 000 plants from 12 species. All species conformed to either the type I or type III pattern of sex change. There was little evidence for a distinct type II pattern of sex change, given that plants with monoecious inflorescences were rare relative to plants with pistillate inflorescences. The mean sex ratio in type I species (79.9% male) was significantly greater than in type III species (63.7% male). The data are consistent with the prediction that type I species are likely to have greater costs associated with female reproduction. We suggest that all Arisaema species have similar patterns of floral development, but differ in their ontogenetic patterns for male and female flowering.  相似文献   

11.
Sex allocation theory forecasts that larger plant size may modify the balance in fitness gain in both genders, leading to uneven optimal male and female allocation. This reasoning can be applied to flowers and inflorescences, because the increase in flower or inflorescence size can differentially benefit different gender functions, and thus favour preferential allocation to specific floral structures. We investigated how inflorescence size influenced sexual expression and female reproductive success in the monoecious Tussilago farfara, by measuring patterns of biomass, and N and P allocation. Inflorescences of T.?farfara showed broad variation in sex expression and, according to expectations, allocation to different sexual structures showed an allometric pattern. Unexpectedly, two studied populations had a contrasting pattern of sex allocation with an increase in inflorescence size. In a shaded site, larger inflorescences were female-biased and had disproportionately more allocation to attraction structures; while in an open site, larger inflorescences were male-biased. Female reproductive success was higher in larger, showier inflorescences. Surprisingly, male flowers positively influenced female reproductive success. These allometric patterns were not easily interpretable as a result of pollen limitation when na?vely assuming an unequivocal relationship between structure and function for the inflorescence structures. In this and other Asteraceae, where inflorescences are the pollination unit, both male and female flowers can play a role in pollinator attraction.  相似文献   

12.
Sex ratios of flowering individuals in dioecious plant populations are often close to unity, or are male biased owing to gender-specific differences in flowering or mortality. Female-biased sex ratios, although infrequent, are often reported in species with heteromorphic sex chromosomes. Two main hypotheses have been proposed to account for female bias: (1) selective fertilization resulting from differential pollen-tube growth of female- versus male-determining microgametophytes (certation); (2) differences in the performance and viability of the sexes after parental investment. Here we investigate these hypotheses in Rumex nivalis (Polygonaceae), a European alpine herb with female-biased sex ratios in which females possess XX, and males XY1Y2, sex chromosomes. Using field surveys and a glasshouse experiment we investigated the relation between sex ratios and life-history stage in 18 populations from contrasting elevations and snowbed microsites and used a male-specific SCAR-marker to determine the sex of nonflowering individuals. Female bias among flowering individuals was one of the highest reported for populations of a dioecious species (mean female frequency = 0.87), but males increased in frequency at higher elevations and in the center of snowbeds. Female bias was also evident in nonflowering individuals (mean 0.78) and in seeds from open-pollinated flowers (mean 0.59). The female bias in seeds was weakly associated with the frequency of male flowering individuals in populations in the direction predicted when certation occurs. Under glasshouse conditions, females outperformed males at several life-history stages, although male seeds were heavier than female seeds. Poor performance of Y1Y2 gametophytes and male sporophytes in R. nivalis may be a consequence of the accumulation of deleterious mutations on Y-sex chromosomes.  相似文献   

13.
Green dragon (Arisaema dracontium; Araceae) is a perennial woodland herb capable of switching gender from year to year. Small flowering plants produce only male flowers but when larger they produce male and female flowers simultaneously. Distinct male and monoecious phenotypes (referred to hereafter as plants) share a single underlying cosexual genotype. Four populations in southern Louisiana were sampled to determine frequencies and size distributions of male and monoecious plants, and to determine the relationship of plant size with male and female flower production in monoecious plants. Male plants were significantly smaller than monoecious plants and made up 34%–78% of flowering plants within populations. Flower number (average = 120) was weakly positively correlated with size. Monoecious plants produced an average of 169 flowers (90 female) and had 100% fruit set, with individual berries containing an average of 2.5 ovules and 1.3 filled seeds. Male flower number was negatively correlated, and female flower number positively correlated, with basal stem diameter. Extrapolation of regression slopes suggested that green dragon should become completely female at a size 20% larger than the largest plant observed in this study. A simple model of inflorescence development is presented to illustrate how the reproductive system of green dragon is related to that of jack-in-the-pulpit (A. tnphyllum), which exhibits a more distinct switch between male and female phenotypes.  相似文献   

14.
 Aquatic plants are well known for their high degree of phenotypic plasticity in vegetative structures, particularly leaves. Less well understood is the extent to which their sexuality can be modified by environmental conditions. Here we investigate gender plasticity in the European clonal monoecious aquatic Sagittaria sagittifolia (Alismataceae) to determine how floral sex ratios may vary with plant size and inflorescence order. We sampled two populations from aquatic habitats in East Anglia, U.K. and measured a range of plant attributes including ramet size and the number of female and male flowers per inflorescence. The two populations exhibited similar patterns of phenotypic gender, despite contrasting patterns of total allocation to female and male flower number. Plants produced male-biased floral sex ratios but female flower number increased from the first to the second inflorescence whereas male flower number decreased. Size-dependent gender modification occurred in both populations, but the patterns of allocation to female flower production differed between the two populations. Our results are consistent with the view that monoecy is a sexual strategy that enables plants to adjust female and male allocation in response to changing environmental conditions. Received September 16, 2002; accepted October 23, 2002 Published online: March 20, 2003  相似文献   

15.
Other than studies on sex-labile Arisaema species, studies of gender patterns in Araceae are scarce. The modification of phenotypic and functional gender was investigated in three populations of the monoecious Arum italicum Miller. The probability of reproduction and the number of inflorescences produced increased with plant size, and flower number (total, male, staminodes, female, pistillodes) increased with both plant and inflorescence sizes. However, plant and inflorescence sizes were poor predictors of floral sex ratio (female to male flower ratio). In contrast, change in floral sex ratio towards increasing femaleness was found among inflorescences sequentially produced by a plant. This change could not be explained by either a decrease in inflorescence size or a change in the mating environment. Differences in functional gender did not appear to be related to plant size or stage in the flowering period. Instead, different patterns of functional gender were found between plants with different number of inflorescences. Multi-inflorescence plants showed a functional gender around 0.5, while plants with one inflorescence showed a more extreme functional gender (either male, female, or functionally sterile). Sex of flowers in this species did not seem to exhibit a phenotypic trade-off.  相似文献   

16.
Summary We measured variation in gender among individuals within populations of ragweed (Ambrosia artemisiifolia) in an abandoned old-field and in the greenhouse. There was great variability in sex expression, from all-female to approximately 78% male. Plants differed significantly in gender in different locations within the field. Plants in an area abandoned from agriculture one year previously were more male than plants in a nearby area abandoned four years previously. In the greenhouse, soil moisture treatments and levels of attack by spittlebugs (Philaenus spumarius) did not affect gender. Height was positively correlated with relative maleness in both populations. Plants with greater shoot weight were relatively more male in the greenhouse, but not in the field. The gender variation we observed either has a genetic basis or is controlled by environmental variables other than those we investigated.  相似文献   

17.
Sex-allocation models predict that the evolution of self-fertilization should result in a reduced allocation to male function and pollinator attraction in plants. The evolution of sex allocation may be constrained by both functional and genetic factors, however. We studied sex allocation and genetic variation for floral sex ratio and other reproductive traits in a Costa Rica population of the monoecious, highly selfing annual Begonia semiovata. Data on biomass of floral structures, flower sex ratios, and fruit set in the source population were used to calculate the average proportion of reproductive allocation invested in male function. Genetic variation and genetic correlations for floral sex ratio and for floral traits related to male and female function were estimated from the greenhouse-grown progeny of field-collected maternal families. The proportion of reproductive biomass invested in male function was low (0.34 at flowering, and 0.07 for total reproductive allocation). Significant among-family variation was detected in the size (mass) of individual male and female flowers, in the proportion of male flowers produced, and in the proportion of total flower mass invested in male flowers. Significant among-family variation was also found in flower number per inflorescence, petal length of male and female flowers, and petal number of female flowers. Except for female petal length, we found no difference in the mean value of these characters between selfed and outcrossed progeny, indicating that, with the possible exception of female petal length, the among-family variation detected was not the result of variation among families in the level of inbreeding. Significant positive phenotypic and broad-sense genetic correlations were detected between the mass of individual male and female flowers, between male and female petal length, and between number of male and number of female flowers per inflorescence. The ratio of stamen-to-pistil mass (0.33) was low compared to published data for autogamous species with hermaphroditic flowers, suggesting that highly efficient selfing mechanisms may evolve in monoecious species. Our results indicate that the study population harbors substantial genetic variation for reproductive characters. The positive genetic correlation between investment in male and female flowers may reflect selection for maximum pollination efficiency, because in this self-pollinating species, each female flower requires a neighboring male flower to provide pollen.  相似文献   

18.
The Chinese yew (Taxus wallichiana var. mairei) is ranked in the first class of important wild endangered plants in China. Due to overexploitation, it now occurs scattered only in the forest undergrowth along the Yangtze River Valley. To improve the conservation management of the species, we applied structural indices to investigate the structural diversity of naturally regenerating yew populations that have established via ex situ conservation. The results show that most yews had larger non‐yew tree neighbors; these were 30–70% larger than their reference trees. Collectively, the average distances between the yews and the three nearest‐neighboring trees were short ( 3 m), suggesting that the yews face strong interspecific competition from neighbors. In these two forest stands, most of the pole‐sized yews were found beneath a single tall neighboring tree (height  10 m), and their growth was enhanced under a single neighboring tree but not under two, three or zero neighboring trees. Finally, we recommend simple silvicultural measures to reduce interspecific competition and improve the vitality of the yew population; specifically, the cutting or pruning of branches of large neighboring trees in tandem with the thinning of canopy trees growing next to the mother yews.  相似文献   

19.
Buchloe dactyloides (Nutt.) Engelm. is a perennial, sod-forming shortgrass of the grasslands of central North America from Montana to northern Mexico. The objectives of this research were to compare a cultivar (Sharps Improved) and a wild population from the Oklahoma panhandle, both in the field and in the greenhouse over several growing seasons, in regard to life-history strategies, sex ratios, and constancy of sex expression. Sex ratios of offspring from individual burs (usually 1–5 seeds per bur) and for all surviving plants germinated from seeds in the greenhouse were markedly different for the two populations. The wild population exhibited a 1: 1 male/female ratio with no monoecious plants and complete stability in sex expression over 3 yr of study in the greenhouse. In contrast, 13% of the cultivar plants were monoecious, and its unisexual plants showed a female bias. Burs of the two populations, in flats exposed to three separated month-long periods of favorable moisture with intervening “drought” periods, showed differential germinability of seeds within burs. The cultivar had less dormancy and produced a greater number of seedlings in the first and second periods; wild burs significantly exceeded cultivar burs in the number of seedlings during the third period and in the number of burs giving rise to seedlings in all three periods. Field and greenhouse measurements of vegetative (VRE) and sexual reproductive effort (SRE) per plant indicated that SRE was minimal (ca. 2% of aboveground biomass) in relation to asexual ramet propagation by rapidly growing stolons (VRE = ca. 80%). Mean SRE and VRE were not only similar for the populations, but also for the males and females within a population, although the values for individual plants were highly variable. It is concluded that Buchloe dactyloides is largely dioecious, with at least some wild populations showing a 1: 1 male/female ratio and constancy of sex expression, and that the bur as a dispersal unit containing 1–5 seeds is particularly appropriate for the breeding system and habitat ecology of Buchloe dactyloides.  相似文献   

20.
Summary Sex ratios of a population and of litters were sampled in muskrats in Ontario, Canada. Sex ratios of litters sampled from nests were male biased (54% male). Until weaning, no differential costs of producing and rearing male and female young were identified that could account for this greater production of males. Following weaning, however, male-biased dispersal of juveniles from their natal site and more frequent acquisition by females of these sites as breeding sites the following year suggested a greater investment by adult females in female young. Therefore, competition between female siblings for the acquisition of their natal site may be sufficient to result in the greater production of males. In addition, the simultaneous occupation of, and competition between, siblings and parents for the resources of the natal home range may not be necessary for local resource competition to result in a greater production of the dispersing sex. Greater-than-expected binomial variance in sex ratios of litters suggested that adjustment of sex-ratios occurred. However, we were unable to associate the adjustment of litter sex ratios with changes in maternal condition. The greater production of males and the predominance of monogamous associations between adults in this population may have lead to slightly greater variation in male fitness than female fitness. Therefore, a female in better-than-average condition may have benefited by producing more males. Similarly, a lower cost of producing dispersing males may allow nutritionally-stressed females to reduce their total expenditure on offspring by producing more males. Because these experiments were non-manipulative, maternal condition may not have varied sufficiently during this study to detect adjustments of litter sex ratios resulting from either of the above mechanisms acting separately, but the combined effects of small differences in matermal condition and selective pressures operating in the same direction may have resulted in the observed deviation from the binomial.  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