首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 0 毫秒
1.
Ratio estimation of the parametric mean for a characteristic measured on plants sampled by a line intercept method is presented and evaluated via simulation using different plant dispersion patterns (Poisson, regular cluster, and Poisson cluster), plant width variances, and numbers of lines. The results indicate that on average the estimates are close to the parametric mean under all three dispersion patterns. Given a fixed number of lines, variability of the estimates is similar across dispersion patterns with variability under the Poisson pattern slightly smaller than varia-bility under the cluster patterns. No variance estimates were negative under the Poisson pattern, but some estimates were negative under the cluster patterns for smaller numbers of lines. Variance estimates become closer to zero similarly for all spatial patterns as the number of lines increases. Ratio estimation of the parametric mean in line intercept sampling works better, from the viewpoint of approximate unbiasedness and variability of estimates, under the Poisson pattern with larger numbers of lines than other combinations of spatial patterns, plant width variances and numbers of lines.  相似文献   

2.
The paper deals with the problem of estimating diversity indexes for an ecological community. First the species abundances are unbiasedly and consistently estimated using designs based on n random and independent selections of plots, points or lines over the study area. The problem of sampling elusive populations is also considered. Finally, the diversity index estimates are obtained as functions of the abundance estimates. The resulting estimators turn out to be asymptotically (n large) unbiased, even if a considerable bias may occur for a small n. Accordingly, the method of jackknifing is made use of in order to reduce bias.  相似文献   

3.
A method for calibrating (localizing) detection function models in line transect sampling is proposed. The method is based on a random parameter model which supplies localized predictions of detection function parameters utilizing a few sample data points from the concerned location(s). The method has the clear advantage of being able to provide density estimates based on very few observations from a location which would be impossible through traditional methods. The method is successfully illustrated using census data on sambar (Cervus unicolor) from a set of wildlife sanctuaries in Kerala, India. The need for further research in this direction is indicated.  相似文献   

4.
For the duration of the war accident in former Yugoslavia, several industrial and military targets were burnt and damaged, resulting in a significant release of persistent organic pollutants. Locations heavily targeted in the attacks were later defined by UNEP as four “hot spots”: Kragujevac, Novi Sad, Pancevo and Bor. We analyzed concentration levels of pollutants collected in 2004 and 2005 in air samples from the city of Kragujevac, Serbia, following the war accident of 1999. Pollutants included polycyclic aromatic hydrocarbons (PAHs), hexachlorocyclohexane (HCH), dichloro-diphenyl-trichloroethane (DDT), dichloro-diphenyl-dichloroethylene (DDE), dichloro-diphenyl-dichloroethane (DDD) and polychlorinated biphenyls (PCBs). We present results obtained during air sampling campaign conducted in July 2004 by the active sampling method; and during September 2004–June 2005 by the passive sampling method. Our findings show the occurrence of residual quantities of DDT, HCH, PCBs and PAHs in air samples. High levels of PCBs are probably due to the destruction of transformers during the war accident.  相似文献   

5.
Freshwater mussels appear to be promising candidates for adaptive cluster sampling because they are benthic macroinvertebrates that cluster spatially and are frequently found at low densities. We applied adaptive cluster sampling to estimate density of freshwater mussels at 24 sites along the Cacapon River, WV, where a preliminary timed search indicated that mussels were present at low density. Adaptive cluster sampling increased yield of individual mussels and detection of uncommon species; however, it did not improve precision of density estimates. Because finding uncommon species, collecting individuals of those species, and estimating their densities are important conservation activities, additional research is warranted on application of adaptive cluster sampling to freshwater mussels. However, at this time we do not recommend routine application of adaptive cluster sampling to freshwater mussel populations. The ultimate, and currently unanswered, question is how to tell when adaptive cluster sampling should be used, i.e., when is a population sufficiently rare and clustered for adaptive cluster sampling to be efficient and practical? A cost-effective procedure needs to be developed to identify biological populations for which adaptive cluster sampling is appropriate.  相似文献   

6.
Various methods exist to model a species’ niche and geographic distribution using environmental data for the study region and occurrence localities documenting the species’ presence (typically from museums and herbaria). In presence-only modelling, geographic sampling bias and small sample sizes represent challenges for many species. Overfitting to the bias and/or noise characteristic of such datasets can seriously compromise model generality and transferability, which are critical to many current applications - including studies of invasive species, the effects of climatic change, and niche evolution. Even when transferability is not necessary, applications to many areas, including conservation biology, macroecology, and zoonotic diseases, require models that are not overfit. We evaluated these issues using a maximum entropy approach (Maxent) for the shrew Cryptotis meridensis, which is endemic to the Cordillera de Mérida in Venezuela. To simulate strong sampling bias, we divided localities into two datasets: those from a portion of the species’ range that has seen high sampling effort (for model calibration) and those from other areas of the species’ range, where less sampling has occurred (for model evaluation). Before modelling, we assessed the climatic values of localities in the two datasets to determine whether any environmental bias accompanies the geographic bias. Then, to identify optimal levels of model complexity (and minimize overfitting), we made models and tuned model settings, comparing performance with that achieved using default settings. We randomly selected localities for model calibration (sets of 5, 10, 15, and 20 localities) and varied the level of model complexity considered (linear versus both linear and quadratic features) and two aspects of the strength of protection against overfitting (regularization). Environmental bias indeed corresponded to the geographic bias between datasets, with differences in median and observed range (minima and/or maxima) for some variables. Model performance varied greatly according to the level of regularization. Intermediate regularization consistently led to the best models, with decreased performance at low and generally at high regularization. Optimal levels of regularization differed between sample-size-dependent and sample-size-independent approaches, but both reached similar levels of maximal performance. In several cases, the optimal regularization value was different from (usually higher than) the default one. Models calibrated with both linear and quadratic features outperformed those made with just linear features. Results were remarkably consistent across the examined sample sizes. Models made with few and biased localities achieved high predictive ability when appropriate regularization was employed and optimal model complexity was identified. Species-specific tuning of model settings can have great benefits over the use of default settings.  相似文献   

7.
Cleaning validation is a major challenge in multi-product pharmaceutical industries. UV spectrophotometric and HPLC methods have been developed and validated for determination of residual amount of Loratadine. Both methods were validated for linearity, range, accuracy, precision, and robustness. The limit of quantification was 1 mg L?1 by UV spectrophotometric method and 0.5 mg L?1 by HPLC method. A spike recovery study was done on a stainless steel (316 grade) plate and specific residual cleaning level (SRCL) was down to 6 μg 25.8 cm?2. Recovery was found to be more than 70%. Both methods were simple, highly sensitive, precise, and accurate, and have potential of being useful for routine quality control.  相似文献   

8.
The combined mark-recapture and line transect sampling methodology proposed by Alpizar-Jara and Pollock [Journal of Environmental and Ecological Statistics, 3(4), 311–327, 1996; In Marine Mammal Survey and Assessment Methods Symposium. G.W. Garner, S.C. Amstrup, J.L. Laake, B.F.J. Manly, L.L. McDonald, and D.C. Robertson (Eds.), A.A. Balkema, Rotterdam, Netherlands, pp. 99–114, 1999] is used to illustrate the estimation of population size for populations with prominent nesting structures (i.e., bald eagle nests). In the context of a bald eagle population, the number of nests in a list frame corresponds to a pre-marked sample of nests, and an area frame corresponds to a set of transect strips that could be regularly monitored. Unlike previous methods based on dual frame methodology using the screening estimator [Haines and Pollock (Journal of Environmental and Ecological Statistics, 5, 245–256, 1998a; Survey Methodology, 24(1), 79–88, 1998b)], we no longer need to assume that the area frame is complete (i.e., all the nests in the sampled sites do not need to be seen). One may use line transect sampling to estimate the probability of detection in a sampled area. Combining information from list and area frames provides more efficient estimators than those obtained by using data from only one frame. We derive an estimator for detection probability and generalize the screening estimator. A simulation study is carried out to compare the performance of the Chapman modification of the Lincoln–Petersen estimator to the screening estimator. Simulation results show that although the Chapman estimator is generally less precise than the screening estimator, the latter can be severely biased in presence of uncertain detection. The screening estimator outperforms the Chapman estimator in terms of mean squared error when detection probability is near 1 wheareas the Chapman estimator outperforms the screening estimator when detection probability is lower than a certain threshold value depending on particular scenarios.  相似文献   

9.
Iwao's quadratic regression or Taylor's Power Law (TPL) are commonly used to model the variance as a function of the mean for sample counts of insect populations which exhibit spatial aggregation. The modeled variance and distribution of the mean are typically used in pest management programs to decide if the population is above the action threshold in any management unit (MU) (e.g., orchard, forest compartment). For nested or multi-level sampling the usual two-stage modeling procedure first obtains the sample variance for each MU and sampling level using ANOVA and then fits a regression of variance on the mean for each level using either Iwao or TPL variance models. Here this approach is compared to the single-stage procedure of fitting a generalized linear mixed model (GLMM) directly to the count data with both approaches demonstrated using 2-level sampling. GLMMs and additive GLMMs (AGLMMs) with conditional Poisson variance function as well as the extension to the negative binomial are described. Generalization to more than two sampling levels is outlined. Formulae for calculating optimal relative sample sizes (ORSS) and the operating characteristic curve for the control decision are given for each model. The ORSS are independent of the mean in the case of the AGLMMs. The application described is estimation of the variance of the mean number of leaves per shoot occupied by immature stages of a defoliator of eucalypts, the Tasmanian Eucalyptus leaf beetle, based on a sample of trees within plots from each forest compartment. Historical population monitoring data were fitted using the above approaches.  相似文献   

10.
土壤重金属X射线荧光光谱非标样测试方法研究   总被引:3,自引:0,他引:3  
采用粉末压片制样,用X射线荧光光谱非标样测试方法测定土壤中Ti、V、Cr、Mn、Fe、Ni、Cu、Zn、Ga、As、Cd、Sn、Sb、Pb、Hg等15种重金属元素。研究了样品制备方法和元素测量条件等影响测试准确度的因素。结果表明此方法不需对固体样品进行消化处理,无需制备标准样片,快速、简便、效率高,是一种非破坏性的分析方法,方法的检出限、准确度和精密度基本能够满足土壤中有毒有害重金属的快速筛查要求。  相似文献   

11.
We investigated quantitatively the sensitivity of plant species response curves to sampling characteristics (number of plots, occurrence and frequency of species), along a simulated pH gradient. We defined 54 theoretical unimodal response curves, issued from combinations of six values for optimum (opt = 3, 4, …, 8), three values for tolerance (tol = 0.5, 1.0, and 1.5, sensu ter Braak and Looman [ter Braak, C.J.F., Looman, C.W.N., 1986. Weighted averaging, logistic regression and the Gaussian response model. Vegetatio 65, 3–11]), and three values for maximum probability of presence (pmax = 0.05, 0.20, and 0.50). For each of these 54 theoretical response curves, we built artificial binary data sets (presence/absence) to test the influence of species occurrence, frequency, or number of available plots. With real data extracted from EcoPlant, a phytoecological database for French forests [Gégout, J.-C., Coudun, Ch., Bailly, G., Jabiol, B., 2005. EcoPlant: a forest sites database linking floristic data with soil characteristics and climatic conditions. J. Veg. Sci. 16, 257–260], we compared the ecological response of 50 plant species to soil pH, based first on a small data set (100 randomly sampled plots), and then based on the whole data set available (3810 plots).  相似文献   

12.
随着肥料质量检测工作的深入开展,比色方法在许多检测项目中得到了越来越多的应用。作者在多年从事土壤肥料检测工作的基础上,介绍复混肥料化验检测中各种比色方法所使用仪器的特点,阐述了比色方法中标准曲线的配制和制作,分析比色方法中环境因素对检测结果的影响,为复混肥料的科学分析检测提供技术依据。  相似文献   

13.
Abstract: Often abundance of rare species cannot be estimated with conventional design‐based methods, so we illustrate with a population of blue whales (Balaenoptera musculus) a spatial model‐based method to estimate abundance. We analyzed data from line‐transect surveys of blue whales off the coast of Chile, where the population was hunted to low levels. Field protocols allowed deviation from planned track lines to collect identification photographs and tissue samples for genetic analyses, which resulted in an ad hoc sampling design with increased effort in areas of higher densities. Thus, we used spatial modeling methods to estimate abundance. Spatial models are increasingly being used to analyze data from surveys of marine, aquatic, and terrestrial species, but estimation of uncertainty from such models is often problematic. We developed a new, broadly applicable variance estimator that showed there were likely 303 whales (95% CI 176–625) in the study area. The survey did not span the whales' entire range, so this is a minimum estimate. We estimated current minimum abundance relative to pre‐exploitation abundance (i.e., status) with a population dynamics model that incorporated our minimum abundance estimate, likely population growth rates from a meta‐analysis of rates of increase in large baleen whales, and two alternative assumptions about historic catches. From this model, we estimated that the population was at a minimum of 9.5% (95% CI 4.9–18.0%) of pre‐exploitation levels in 1998 under one catch assumption and 7.2% (CI 3.7–13.7%) of pre‐exploitation levels under the other. Thus, although Chilean blue whales are probably still at a small fraction of pre‐exploitation abundance, even these minimum abundance estimates demonstrate that their status is better than that of Antarctic blue whales, which are still <1% of pre‐exploitation population size. We anticipate our methods will be broadly applicable in aquatic and terrestrial surveys for rarely encountered species, especially when the surveys are intended to maximize encounter rates and estimate abundance.  相似文献   

14.
Food webs are usually aggregated into a manageable size for their interpretation and analysis. The aggregation of food web components in trophic or other guilds is often at the choice of the modeler as there is little guidance in the literature as to what biases might be introduced by aggregation decisions. We examined the impacts of the choice of the a priori model on the subsequent estimation of missing flows using the inverse method and on the indices derived from ecological network analysis of both inverse method-derived flows and on the actual values of flows, using the fully determined Sylt-Rømø Bight food web model. We used the inverse method, with the least squares minimization goal function, to estimate ‘missing’ values in the food web flows on 14 aggregation schemes varying in number of compartments and in methods of aggregation. The resultant flows were compared to known values; the performance of the inverse method improved with increasing number of compartments and with aggregation based on both habitat and feeding habits rather than diet similarity. Comparison of network analysis indices of inverse method-derived flows with that of actual flows and the original value for the unaggregated food web showed that the use of both the inverse method and the aggregation scheme affected indices derived from ecological network analysis. The inverse method tended to underestimate the size and complexity of food webs, while an aggregation scheme explained as much variability in some network indices as the difference between inverse-derived and actual flows. However, topological network indices tended to be most robust to both the method of determining flows and to the inverse method. These results suggest that a goal function other than minimization of flows should be used when applying the inverse method to food web models. Comparison of food web models should be done with extreme care when different methodologies are used to estimate unknown flows and to aggregate system components. However, we propose that indices such as relative ascendency and relative redundancy are most valuable for comparing ecosystem models constructed using different methodologies for determining missing flows or for aggregating system components.  相似文献   

15.
用离子交换树脂袋法,研究了鼎湖山三种森林(马尾松林、马尾松针叶阔叶混交林和季风常绿阔叶林)土壤硝态氮对外加氮的响应特征。结果表明,土壤硝态氮显著地受森林类型、季节和氮处理的影响。整体而言,阔叶林土壤硝态氮极显著高于马尾松林和混交林,而马尾松林和混交林之间的差异则不显著。三种森林土壤硝态氮的季节变化均表现为春季和夏季极显著高于冬季和秋季,而冬季又显著高于秋季。外加氮处理提高土壤硝态氮水平,其中在马尾松林和阔叶林氮处理效应显著。所得结果与直接采集土壤或土壤水测定的硝态氮含量的结果一致,表明离子交换树脂袋法是评价土壤硝态氮水平行之有效的手段之一。  相似文献   

16.
月湖近代生物硅沉积测定与营养演化的动态过程   总被引:2,自引:2,他引:2  
在综合了国内外实验方法基础上,完善了湖泊沉积生物硅实验室测定方法.利用此方法,首次对一富营养化小型浅水湖泊--武汉市汉阳区月湖进行了2个柱状沉积物的生物硅含量测试.结果表明,生物硅可以反映月湖的营养演化的动态过程,130 a前,湖泊受到的环境压力为轻;20世纪初期,月湖生物硅沉积缓慢上升,为水体富营养化开始发生时间;20世纪30年代至50年代,月湖已成为一富营养化湖泊;20世纪60年代至80年代,月湖水质持续恶化;20世纪80年代至月湖清淤前这段时期入湖污水增多,这加速了生物硅的沉积,正是这段时期使月湖最终沦为劣五类水质的湖泊.  相似文献   

17.
Despite international waters covering over 60% of the world's oceans, understanding of how fisheries in these regions shape ecosystem processes is surprisingly poor. Seabirds forage at fishing vessels, which has potentially deleterious effects for their population, but the extent of overlap and behavior in relation to ships is poorly known. Using novel biologging devices, which detect radar emissions and record the position of boats and seabirds, we measured the true extent of the overlap between seabirds and fishing vessels and generated estimates of the intensity of fishing and distribution of vessels in international waters. During breeding, wandering albatrosses (Diomedea exulans) from the Crozet Islands patrolled an area of over 10 million km2 at distances up to 2500 km from the colony. Up to 79.5% of loggers attached to birds detected vessels. The extent of overlap between albatrosses and fisheries has widespread implications for bycatch risk in seabirds and reveals the areas of intense fishing throughout the ocean. We suggest that seabirds equipped with radar detectors are excellent monitors of the presence of vessels in the Southern Ocean and offer a new way to monitor the presence of illegal fisheries and to better understand the impact of fisheries on seabirds.  相似文献   

18.
采用"半静态法"测定了3种农药及其混剂对大型溞的24 h、48 h急性毒性,根据我国《化学农药环境安全评价实验准则》中的毒性等级标准,它们对大型溞的毒性等级如下:精甲霜灵悬浮种衣剂对大型溞的24 h、48 h-EC_(50)均大于10 a.i.mg·L~(-1),属"低毒"级,咯菌腈悬浮种衣剂对大型溞的24 h、48 h-EC_(50)分别是0.339 mg·L~(-1)、0.246 mg·L~(-1),根据0.1 a.i.mg·L~(-1)EC_(50)(48 h)≤1.0 a.i.mg·L~(-1)判断,属"高毒"级。嘧菌酯水分散粒剂对大型溞的24 h、48 h-EC_(50)分别是0.389 mg·L~(-1)、0.286 mg·L~(-1),根据0.1a.i.mg·L~(-1)EC_(50)(48 h)≤1.0 a.i.mg·L~(-1)判断,属"高毒"级。精甲霜灵·咯菌腈·嘧菌酯悬浮种衣剂对大型溞的24 h、48 h-EC_(50)分别是0.292 mg·L~(-1)、0.228 mg·L~(-1),根据0.1 a.i.mg·L~(-1)EC_(50)(48 h)≤1.0 a.i.mg·L~(-1)判断,属"高毒"级。精甲霜灵·咯菌腈·嘧菌酯悬浮种衣剂和嘧菌酯水分散粒剂都属"高毒",但比较具体数值,发现精甲霜灵·咯菌腈·嘧菌酯悬浮种衣剂毒性相对更大,原因是其中还含有"高毒"的咯菌腈。  相似文献   

19.
3种杀菌剂及其复配剂对斑马鱼的急性毒性   总被引:1,自引:0,他引:1  
采用"半静态法"测定了3种农药及其复配剂对斑马鱼的24 h、48 h、72 h和96 h急性毒性,根据我国《化学农药环境安全评价实验准则》中毒性等级划分标准来划分它们对斑马鱼的毒性等级,发现35%精甲霜灵悬浮种衣剂的24 h、48 h、72 h和96 h-LC_(50)均超过10 a.i.mg·L~(-1),属"低毒"级,而25 g·L~(-1)咯菌腈悬浮种衣剂的24 h、48 h、72 h和96 h-LC_(50)分别为0.704 mg·L~(-1)、0.514 mg·L~(-1)、0.424 mg·L~(-1)、0.262 mg·L~(-1),根据0.100 a.i.mg·L~(-1)LC_(50)(96 h)≤1.00 a.i.mg·L~(-1)划分,属"高毒"级,80%嘧菌酯水分散粒剂的24 h、48 h、72 h和96 h-LC_(50)分别为9.803 mg·L~(-1)、5.175 mg·L~(-1)、4.328 mg·L~(-1)、2.326 mg·L~(-1),根据1.00 a.i.mg·L~(-1)LC_(50)(96 h)≤10.00 a.i.mg·L~(-1)划分,属于"中毒"级。11%精甲霜灵-咯菌腈-嘧菌酯悬浮种衣剂的24 h、48 h、72 h和96 h-LC_(50)分别为2.267 mg·L~(-1)、2.073 mg·L~(-1)、1.620 mg·L~(-1)、1.280 mg·L~(-1),根据1.00 a.i.mg·L~(-1)LC_(50)(96 h)≤10.00 a.i.mg·L~(-1)划分,属于"中毒",虽然11%精甲霜灵-咯菌腈-嘧菌酯悬浮种衣剂和80%嘧菌酯水分散粒剂同属"中毒",但比较具体数值,发现11%精甲霜灵-咯菌腈-嘧菌酯悬浮种衣剂毒性相对更大,原因是其中含有"高毒"的咯菌腈。  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