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 484 毫秒
1.
我国珍稀濒危植物见血封喉在海口地区集中分布于自然村落周围。为初步探讨人类活动对见血封喉种群数量特征和空间分布格局的影响,在海口地区共设置6个样地,从种群的大小级结构、静态生命表、存活曲线、空间分布格局及其动态5个方面分析了海口地区种群的数量特征和空间分布格局。结果表明:①种子的萌发率高,幼苗个体丰富,但是幼树个体缺乏,中龄期个体相对丰富,老龄期个体数量较少;纺锤型的大小级结构表明种群属于衰退型。②现存的Ⅱ、Ⅲ级个体数少,导致静态生命表中Ⅱ、Ⅲ级个体死亡率出现负值,由于接近实际寿命,在第Ⅵ级死亡率达到最高。③虽然见血封喉种子成活率较高,但由于其幼树的存活率较低,其存活曲线接近DeeveyⅢ型;但是如果剔除掉人为干扰在幼树阶段的影响,其生存曲线属于Ⅰ型,呈凸型,这表明该海口地区的环境条件较适宜见血封喉种群的生长。④种群空间分布格局总体为聚集分布,这与大多数珍稀植物种群一致;人为干扰和自然环境影响其分布格局,使种群由聚集分布转变成随机分布,后又向聚集分布发展。  相似文献   

2.
Betula ermanni population was divided into three groups: the upper population (2 000–2 200 m), the middle population (1 700–2 000 m), and the down population (1 400–1 700 m) in Changbai Mountain. The dynamics ofBetula ermanni populations in subalpine vegetation are studied and the population life table, fecundity schedule, survival curves, age structure, and fecundity curves were established. The results showed that the, middle population is obviously, the transition from the upper population to the down population. This project is supported by Chinese Academy of Science Responsible editor  相似文献   

3.
色木槭次生林种群结构动态分析   总被引:1,自引:0,他引:1  
以种群生命表及生存分析理论为基础,将林木依胸径大小分级,以林木径级结构代表年龄结构,编制了色木槭种群的特定时间生命表,绘制了存活曲线、死亡率曲线、损失度曲线和生存函数曲线,分析种群数量特征。结果表明:色木槭种群幼年个体丰富,中老年个体较少,种群在第Ⅲ和第Ⅷ径级出现死亡高峰,种群存活曲线属于Deevey-Ⅱ型,色木槭种群...  相似文献   

4.
In order to understand the relationship between population succession and its genetic behavior, random amplified polymorphic DNA (RAPD) technique was used to analyze the genetic diversity of Quercu glandulifera var. brevipetiolata populations in three forest communities with different succession stages (coniferous forest, coniferous and broad-leaved mixed forest, evergreen broad-leaved forest). The results showed that 145 repetitive loci were produced in 60 individuals of Q. glandulifera using 11 primers, among which 120 loci were polymorphic, and the total percentage of polymorphic loci was 82.76% with an average of 64.14%. Estimated by the Shannon information index, the total genetic diversity of the three populations was 0.4747, with an average of 0.3642, while it was 0.3234, with an average of 0.2484, judged from the Nei index. Judged from percentage of polymorphic loci, Shannon inform at ion index and Nei index, the genetic diversity followed a decreasing order: coniferous forest > broad-leaved mixed forest > evergreen broad-leaved forest. Analysis of molecular variance (AMOVA) showed that 69.73% of the genetic variance existed within populations and 30.27% of the genetic variance existed among populations. The coefficient of gene differentiation (GST) was 0.2319 and the gene flow (N m) was 1.6539. The mean of genetic identity among populations of Q. glandulifera was 0.8501 and the mean of genetic distance was 0.1626. The genetic identity between the Q. glandulifera population in the coniferous forest and that in the coniferous and broadleaved mixed forest was the highest. UPGMA cluster analysis based on Nei’s genetic distance showed that the population in the coniferous forest gathered with that in the coniferous and broad-leaved mixed forest firstly, then with that in the evergreen broad-leaved forest. The genetic structure of Q. glandulifera was not only characteristic of the biological characteristics of this species, but was also influenced by the microenvironment in different communities. __________ Translated from Journal of Northwest Forestry University, 2008, 23(1): 18–22 [译自: 西北林学院学报]  相似文献   

5.
Comparing with an affinity and widespread speciesAdenophora potaninii, the age structures of A.lobophylla population, an endangered plant species were studied. 29 sample plots were investigated in the centered distribution area, Jinchuan county from altitude 2 300 m to 3 400 m. The main factors which influence the population age structures of A.lobophylla were analyzed. The principal characteristics of A.lobophylla populations are that the number of seedling and the density of population are much less than these of A.potaninii population. Below altitude of 2 700 m, the age structures of most A.lobophylla populations show the declining status, only above altitude of 2 700 m they are stable, while age structures of all populations of A.potaninii populations at corresponding altitude perform stable and developing features. The age structure patterns, relations between age (x) and number of individual (y) of stable populations of A.lobophylla can be expressed by equation: y =e(a-bx), and the longest life span is 25a. Whereas A.potaninii populations can be expressed by the equation: y =ax-b, and the longest life span of individual is 21a. The external factors, which constrain the extension of A.lobophylla population, are extreme environmental conditions such as serious drought, external disturbance and low temperature. The project was supported by the National Nurture Science Foundation of P.R. China (No.393915007). (Responsible Editor: Zhu Hong)  相似文献   

6.
领春木(Euptelea pleiospermum)是东亚特有珍稀濒危植物,在北川小寨子沟自然保护区主要沿河岸分布.本文采用样方法,从种群的大小级结构、静态生命表、存活曲线、空间分布格局4个方面分析了该自然保护区领春木种群的数量特征与空间分布格局.结果表明:1)幼龄期个体很少,中龄期个体相对较多,老龄期个体数量也很稀少.由此可知其大小级结构呈纺锤形,因此种群属于衰退型;2)Ⅰ、Ⅱ级个体数少,致使其在静态生命表中死亡率出现负值,个体在第Ⅳ级有一个死亡率高峰,而在第Ⅶ级死亡率达到了最高;3)虽然幼苗的存活率很低,但由于幼树的存活率较高,所以存活曲线接近Deevey Ⅰ型,表明该地区沿河岸带的环境条件比较适合领春木的生长;4)种群空间分布格局为聚集分布.为了保护和维持种群,应对群落进行适度的干扰,开辟林窗,促进领春木种子萌发,幼苗和幼树正常生长.  相似文献   

7.
8.
Understanding population structure provides basic ecological data related to species and ecosystems. Our objective was to understand the mechanisms involved in the maintenance of Quercus aquifolioides populations. Using a 1 ha permanent sample plot data for Q. aquifolioides on Sejila Mountain, Tibet Autonomous Region (Tibet), China, we analyzed the population structure of Q. aquifolioides by combining data for diameter class, static life table and survival curve. Simultaneously, the spatial distribution of Q. aquifolioides was studied using Ripley’s L Function in point pattern analysis. The results showed: (1) Individuals in Q. aquifolioides populations were mainly aggregated in the youngest age classes, that accounted for 94.3% of the individuals; the older age classes had much smaller populations. Although the youngest age classes (Classes I and II) had fewer individuals than Class III, the total number of individuals in classes I and II was also greater than in classes IV to IX. In terms of tree height, few saplings, more medium-sized saplings and few large-sized trees were found. The diameter class structure of Q. aquifolioides populations formed an atypical ‘pyramid’ type; the population was expanding, but growth was limited, tending toward a stable population. (2) Mortality of Q. aquifolioides increased continuously with age; life expectancy decreased over time, and the survivorship curve was close to a Deevey I curve. (3) The spatial distribution pattern of Q. aquifolioides varied widely across different developmental stages. Saplings and medium-sized tree showed aggregated distributions at the scales of 0–33 m and 0–29 m, respectively. The aggregation intensities of saplings and medium-sized trees at small scales were significantly stronger than that of large-sized trees. However, large-sized trees showed a random distribution at most scales. (4) No correlation was observed among saplings, medium- and large-sized trees at small scales, while a significant and negative association was observed as the scale increased. Strong competition was found among saplings, medium- and large-sized trees, while no significant association was observed between medium- and large-sized trees at all scales. Biotic interactions and local ecological characteristics influenced the spatial distribution pattern of Q. aquifolioides populations most strongly.  相似文献   

9.
This study on the allelopathic effects and chemical components of the essential oi l from Eucalyptus grandis × E. urophylla shows that the leaf oil emulsion of E. grandis × E. urophylla can inhibit the proliferation of pathogenic fungi Fusarium oxysporum, Pyriculerie grisea, Glorosprium musa rum and Phytophthora capsici. Pupation and feeding of the pest insects Spodopteralitura Fabricius and Helicoverpa armigera Hubner are shown to be affected with restraining effects which increase with the increasing levels of oil concentration. A GC/MS analysis of the leaf oil indicated that the main components, with a relative content of ⩾3%, were alloocimene (43.22%), α-pinene (13.63%), γ-terpinene (5.49%), (E)-3,7-dimethyl-2,6-octadien-1-ol (3.58%), β-fenchyl alcohol (4.58%), and 2-amino-3,5-dicyano-6-(4-methoxyphenoxy)-pyridine (3.67%). Terpenes played an important role in the inhibitory effects of E. grandis × E. urophylla essential oil on pathogenic fungi and pest insects. Poor biodiversity of eucalyptus plantations is a function of allelopathy. __________ Translated from Chinese Journal of Ecology, 2007, 26(6): 835–839 [译自: 生态学杂志]  相似文献   

10.
We evaluated the survival and growth of Abies homolepis and Picea jezoensis var. hondoensis seedlings on Mt. Ohdaigahara, where the population of sika deer (Cervus nippon) is high and an experimental fence has been in place for 13 years. No significant differences were detected in the survival of small seedlings between fenced and unfenced plots. The growth of A. homolepis was significantly higher in the fenced plot, but growth of P. jezoensis var. hondoensis did not show significant differences between fenced and unfenced plots. Seedlings of height ≤5 cm in the forest floor vegetation of the unfenced plot were probably too small for deer to find and browse, so they survived.  相似文献   

11.
沙地樟子松天然纯林的结构特征   总被引:5,自引:2,他引:3       下载免费PDF全文
[目的]为了解红花尔基地区沙地樟子松天然纯林的结构特征,指导沙地樟子松的保护与经营。[方法]在红花尔基地区设置2块100 m×100 m的樟子松天然纯林固定样地,利用样地内每木定位调查数据和分析统计软件进行一元分布及二元分布特征分析。[结果](1)樟子松天然林纯林直径分布为单峰或多峰山状分布,垂直结构简单,只有乔木层和草本层。(2)樟子松天然纯林的林木分布格局为均匀分布,接近随机分布,林木分布格局类型与林分密度无关;林分中樟子松个体竞争激烈,多数单元中林木呈较密集状态。(3)2块样地中随机分布状态下的林木多数为中等密集或比较密集,不同分布状态下的林木优劣性差异较小;低密度樟子松天然纯林中多数密集状态的林木为绝对优势木或优势木,而高密度林分中林木密集度分布与林木大小无关。[结论]红花尔基沙地樟子松天然林结构不合理,应选择病腐木及聚集分布的个体作为潜在调整对象,进行密度调整和结构优化。  相似文献   

12.
Starch-gel electrophoresis was used to resolve nine polymorphic enzyme loci from leaf tissue collected from 20 Korean populations ofEurya japonica in order to determine differences in allele frequencies between male and female trees. In addition, 84 adults were sampled and mapped in a population located on Naenaro Island in Korea to examine spatial genetic structure using spatial autocorrelation analysis. Allele frequencies between males and females gave few contribution to the genetic structuring within populations. Only nine (5%) of 180 cases were significantly different from both sexes in allele frequencies. On the other hand, Moran'sI was significantly different from the expected value in 31 (23.5%) of 132 cases. In the shortest distance (0<5m),I was significantly positive in 10 (22.7%) of 44 cases. The results indicate that a significant small scale genetic structure was detected in the population and patch widths were inferred to be approximately 5–7 m. A nonrandom distribution of genotypes may be indicative of restricted gene flowvia seed and pollen dispersal, and patchy establishments of genetically distinct individuals. These factors are responsible for shaping population genetic structure ofE. japonica.  相似文献   

13.
[目的]探讨南亚热带西南桦和尾巨桉人工纯林的凋落叶分解动态及其与土壤化学性质之间的相关关系.[方法]采用原位分解袋法研究凋落叶的分解过程.[结果]表明:西南桦、尾巨桉人工林凋落叶分解系数分别为0.96 a-1和0.88 a-1.在为期12个月的分解试验中,2种凋落叶有机C含量在整个分解过程中呈逐渐下降趋势;全K含量和C/N比在分解前期迅速下降,之后趋于平缓;全N含量和全P含量在整个分解过程中呈逐渐上升趋势;2种凋落叶N/P比则呈先升高后下降的趋势.无论是分解前期还是分解后期,凋落叶质量损失与N含量均呈显著正相关(前期R=0.877;后期R=0.855),与C/N均呈显著负相关(前期R=-0.735;后期R=-0.697).与尾巨桉林地土壤性质相比,西南桦凋落叶分解提高了林地0~10、10~20 cm土壤的有机C、全N、全P、全K、N/P,对2030 cm土壤有机C、全K、pH值、C/N、N/P则未产生显著影响.相关分析表明:凋落叶初始有机C含量与土壤有机C、全N、全P、全K、N/P显著相关;凋落叶初始全N含量与土壤全N、pH值、C/N显著相关.[结论]凋落叶的养分含量与土壤养分的关系紧密;与尾巨桉相比,西南桦凋落叶的养分含量明显较高,分解速率更快,释放到土壤中的养分也更多.  相似文献   

14.
2007年4月通过对云南省德宏州陇川县内40块样地的调查,分析了中缅木莲的年龄结构、高度结构、空间分布格局,编制了特定时间生命表,绘制了存活曲线。结果表明:中缅木莲成年个体数较多,幼年植物个体较少,居群属衰退型居群;个体高度发育,较为连续;其空间分布格局表现为随机分布;从幼苗生长到树木死亡分4个阶段,其幼树、成树阶段死亡率较低,幼苗、老树阶段死亡率较高。  相似文献   

15.
Based on the data collected from 27 plots of the Pinus armandii community in Qinling Mountains, we studied the spatial distribution pattern, scale, and gap characteristics of the P. armandii population. The results showed that the population had a clumped distribution before age 50. At the age range from 15 to 25, though the population tended to be distributed randomly, the distribution was still clumped. The population distribution at the age range from 40 to 50 was at the transitional stage from clumped to random. After age 50, the population started to be senesced, the distribution pattern turning from clumped to random. The distribution pattern scale of P. armandii always changes with the development stage of the population, being 100 m2 in general. The gap size of P. armandii population was similar to its distribution pattern scale, and the gaps of 80–130 m2 occupied 59% of the total. Because of the better light and nutrient condition in the gap, P. armandii seedlings grew well, which helped the population keep its stability through “mobile mosaic circling”. __________ Translated from Chinese Journal of Ecology, 2006, 25(6): 652–656 [译自: 生态学杂志]  相似文献   

16.
The potential contribution of agroforestry systems to the management and genetic resources conservation in iroko (Milicia excelsa), an important and valuable timber tree species in sub-Saharan Africa, is addressed in this paper. The structure and dynamics of traditional agroforestry systems and the ecological structure of Milicia excelsa populations in farmlands were studied through a survey carried out in 100 farmlands covering the natural range of iroko in Benin. Forty-five species belonging to 24 plant families were recorded in traditional agroforestry systems. Average tree density varied from 1 to 7 stems ha−1 with diversity index ranging from 2.6 to 2.9. Milicia excelsa occurred sparsely in agroforestry systems in all regions, with density ranging from 1 to 4 stems ha−1; stand basal area varying from 33.10−4 to 129.10−4 m2 ha−1, and negligible seedling regeneration. However, male and female trees were apparently evenly distributed on farmlands in all regions (F/M > 0). Iroko trees produced viable seeds with moderate germination rate and early growth (germination rate 22% and height 7.29 cm after 3 months). Suggestions are made regarding optimal densities for iroko conservation in farmlands, according to farmers’ socioeconomic conditions in different regions, in order to improve traditional agroforestry systems and their use as biological corridors in conservation of Milicia excelsa genetic resources.  相似文献   

17.
Doubts exist about the effectiveness of establishing trees near saline discharge areas on farmland to manage dryland salinity. These centre on low rates of water uptake from saline water tables, salt accumulation in tree root zones and the consequent poor growth and survival of trees. Despite this, trees still survive in many plantations established adjacent to saline discharge areas and land-holders often favour such locations, as they do not compete for arable land such as that occurs with plantings in recharge areas. Tree performance and salt accumulation were assessed in three experimental plantations established adjacent to saline discharge areas 20–25 years ago. These were all in the 400–600 mm rainfall zone of south-western Australia. Mean soil salinity, within 1 m of the surface, ranged from 220 to 630 mS m−1, while permanent ground-waters occurred within 2–5 m of the surface and had electrical conductivities ranging from 175 to 4150 mS m−1. The study confirmed the low growth rates expected for trees established over shallow, saline water tables in a relatively low rainfall environment, with estimated wood volumes in Eucalyptus cladocalyx, E. spathulata, E. sargentii, E. occidentalis and E. wandoo of between 0.5 and 1.5 m3 ha−1 yr−1. Values of up to 3 m3 ha−1 yr−1 were obtained on soils with low salinity (<200 mS m−1). The excellent survival (>70%) of several Eucalyptus species confirms that discharge plantations species can persist, despite increasing soil salinity. However, the long-term sustainability of such plantings (50–100 years) without broader landscape treatment of the present hydrological imbalance must be questioned.  相似文献   

18.
India is the largest grower of Eucalyptus with an area of 3.943 million hectares under plantations and E. tereticornis is the predominant species in the plains of southern India with an average productivity of 12–25 m3 ha−1 year−1. With the aim to establish seedling seed orchards of the species, seed lots of fifteen provenances were imported from Australia and a trial was laid. In the present study the genetic diversity existing in the seed orchard was estimated using ISSR–PCR. Seven ISSR primers amplified 663 amplicons in the size ranging from 255 to 2,711 bp. The total number of polymorphic bands varied from 59 to 123 with 100% polymorphic banding profiles. The average gene diversity (Hj) of all the provenances ranged from 0.0589 to 0.1109 and the total gene diversity estimated was low (H T = 0.130) when compared to the earlier reports from other eucalypts species. Analysis of Molecular Variance partitioned the ISSR variation into inter and intra population components. The inter population component accounted for 55.2% of the variation and the intra population component accounted for 46.3% (P < 0.001). A neighbour-joining analysis was done using the dissimilarity matrix to determine the aggregation of the individuals into clusters. Existence of population structure among the populations was revealed in STRUCTURE analysis but geographical region based clustering was not observed. The assessment of intra and inter genetic variation documented in the present study suggests that, along with the phenotypic traits, knowledge about genetic diversity measured at the DNA level in individuals of seed orchards provide an objective guide for selective culling of trees for maintaining optimal diversity for enhanced genetic gains.  相似文献   

19.
The random amplified polymorphic DNA (RAPD) technique was used to evaluate the genetic diversity and population structure of 91 genets from four wild populations of Betula luminifera at different elevations in the National Nature Reserve of theWuyi Mountain, Fujian Province, China. Eighteen random primers (from 139 primers) produced a total of 199 scorable amplified fragments, of which 174 (87.44%) were polymorphic across all individuals. The genetic diversities of B. luminifera at the population level and species level were PPL = 60.05%, h = 0.2242, I = 0.3181 and PPL = 87.44%, h = 0.3442, I = 0.4899, respectively. The value of differentiation (G st= 0.3486) and analysis of molecular variance (AMOVA) indicated that there was a relatively high genetic differentiation among populations, and about one-third of the genetic variation occurred among populations. Pearson correlation analysis further revealed that the genetic diversity within populations had significant or very significant correlation with the elevation, climatic factors (annual average temperature and annual precipitation) and soil nutrient factors (total nitrogen, C/N ratio and organic matter). Mantel tests show that there was a significant correlation between the genetic distances among populations and the distance of elevation, and the divergence of soil nutrient factors. The results of the present study suggested that the relatively high genetic differentiation among populations of B. luminifera at different elevations might be caused by ecological factors and gene flow. __________ Translated from Scientia Silvae Sinicae, 2008, 44(3): 50–55 [译自: 林业科学]  相似文献   

20.
Fine root lifespan and turnover play an important role in carbon allocation and nutrient cycling in forest ecosystems. Fine roots are typically defined as less than 1 or 2 mm in diameter. However, when categorizing roots by this diameter size, the position of an individual root on the complex lateral branching pattern has often been ignored, and our knowledge about relationships between branching order and root function thus remains limited. More recently, studies on root survivals found that longevity was remarkably different in the same branching level due to diameter variations. The objectives of this study were: (1) To examine variations of fine root diameter from the first-to fifth-orders in Fraxinus mandshurica Rupr and Larix gmelinii Rupr roots; and (2) To reveal how the season, soil nutrient, and water availability affect root diameter in different branch order in two species. This study was conducted at Maoershan Forest Research Station (45°21′–45°25′N, 127°30′–127°34′E) owned by Northeast Forestry University in Harbin, northeast China. Both F. mandshurica and L. gmelinii were planted in 1986. In each plantation, fine roots of two species by sampling up to five fine root branch orders three times during the 2003 growing season from two soil depths (i.e., 0–10 and 10–20 cm) were obtained. The results showed that average diameters of fine roots were significantly different among the five branch orders. The first-order had the thinner roots and the fifth order had the thickest roots, the diameter increasing regularly with the ascending branch orders in both species. If the diameter of fine roots was defined as being smaller than 0.5 mm, the first three orders of F. mandshurica roots and the first two orders of L. gmelinii roots would be included in the fine root population. The diameter ranges of the fine roots from first-order to fifth-order were 0.15–0.58, 0.18–0.70, 0.26–1.05, 0.36–1.43, and 0.71–2.96 mm for F. mandshurica, and 0.17–0.76, 0.23–1.02, 0.26–1.10, 0.38–1.77, and 0.84–2.80 mm for L. gmelinii. The average coefficient of variation in first-order roots was less than 10%, second-and third-order was 10–20%, and fourth-and fifth-order was 20–30%. Thus, variation in root diameter also increased with the ascending root order. These results suggest that “fine roots”, which are traditionally defined as an arbitrary diameter class (i.e., <2 mm in diameter) may be too large a size class when compared with the finest roots. The finest roots have much shorter lifespan than larger diameter roots; however, the larger roots are still considered a component of the fine root system. Differences in the lifespan between root diameter and root order affect estimates of root turnover. Therefore, based on this study, it has been concluded that both diameter and branch order should be considered in the estimation of root lifespan and turnover. __________ Translated from Acta Phytoecologica Sinica, 2005, 29(6): 871–877 [译自: 植物生态学报]  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