首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Abstract

The effects of small‐scale surface inhomogeneities on the turbulence structure in the convective boundary layer are investigated using a high‐resolution large‐eddy simulation model. Surface heat flux variations are sinusoidal and two‐dimensional, dividing the total domain into a checkerboard‐like pattern of surface hot spots with a 500‐m wavelength in the x and y directions, or 1/4 of the domain size. The selected wind speeds were 1 and 4 m s‐l, respectively. As a comparison, a simulation of the turbulence structure was performed over a homogeneous surface.

When the wind speed is light, surface heat flux variations influence the horizontally averaged turbulence statistics, including the higher moments despite the small characteristic length of the surface perturbation. Stronger mean wind speeds weaken the effects of inhomogeneous surface conditions on the turbulence structure in the convective boundary layer.

Results from conditional sampling show that when the mean wind speed is small, weak mean circulations occur, with updraft branches above the high heat flux regions and down‐draft branches above the low heat flux regions. The inhomogeneous surface induces significant differences in the turbulence statistics between the high and low heat flux regions. However, the effect of the surface perturbations weaken rapidly when the mean wind speed increases. This research has implications in the explanation of the large‐scale variability commonly encountered in aircraft observations of atmospheric turbulence.  相似文献   

2.
3.
Turbulence structures in the katabatic flow in the stable boundary layer (SBL) over the ice sheet are studied for two case studies with high wind speeds during the aircraft-based experiment KABEG (Katabatic wind and boundary layer front experiment around Greenland) in the area of southern Greenland. The aircraft data allow the direct determination of turbulence structures in the katabatic flow. For the first time, this allows the study of the turbulence structure in the katabatic wind system over the whole boundary layer and over a horizontal scale of 80 km.The katabatic flow is associated with a low-level jet (LLJ), with maximum wind speeds up to 25 m s-1. Turbulent kinetic energy (TKE) and the magnitude of the turbulent fluxes show a strong decrease below the LLJ. Sensible heat fluxes at the lowest level have values down to -25 W m-2. Latent heat fluxes are small in general, but evaporation values of up to +13 W m-2 are also measured. Turbulence spectra show a well-defined inertial subrange and a clear spectral gap around 250-m wavelength. While turbulence intensity decreases monotonously with height above the LLJ for the upper part of the slope, high spectral intensities are also present at upper levels close to the ice edge. Normalized fluxes and variances generally follow power-law profiles in the SBL.Terms of the TKE budget are computed from the aircraft data. The TKE destruction by the negative buoyancy is found to be very small, and the dissipation rate exceeds the dynamical production.  相似文献   

4.
Daily observations of wind speed at 12 stations in the Greater Beijing Area during 1960–2008 were homogenized using the Multiple Analysis of Series for Homogenization method. The linear trends in the regional mean annual and seasonal (winter, spring, summer and autumn) wind speed series were-0.26,-0.39,-0.30,-0.12 and-0.22 m s-1 (10 yr)-1 , respectively. Winter showed the greatest magnitude in declining wind speed, followed by spring, autumn and summer. The annual and seasonal frequencies of wind speed extremes (days) also decreased, more prominently for winter than for the other seasons. The declining trends in wind speed and extremes were formed mainly by some rapid declines during the 1970s and 1980s. The maximum declining trend in wind speed occurred at Chaoyang (CY), a station within the central business district (CBD) of Beijing with the highest level of urbanization. The declining trends were in general smaller in magnitude away from the city center, except for the winter case in which the maximum declining trend shifted northeastward to rural Miyun (MY). The influence of urbanization on the annual wind speed was estimated to be about-0.05 m s-1 (10 yr)-1 during 1960–2008, accounting for around one fifth of the regional mean declining trend. The annual and seasonal geostrophic wind speeds around Beijing, based on daily mean sea level pressure (MSLP) from the ERA-40 reanalysis dataset, also exhibited decreasing trends, coincident with the results from site observations. A comparative analysis of the MSLP fields between 1966–1975 and 1992–2001 suggested that the influences of both the winter and summer monsoons on Beijing were weaker in the more recent of the two decades. It is suggested that the bulk of wind in Beijing is influenced considerably by urbanization, while changes in strong winds or wind speed extremes are prone to large-scale climate change in the region.  相似文献   

5.
Summary A study of mean wind speeds and directions has been completed in the Snowy Range of Southern Wyoming, U.S.A. It was conducted in a subalpine ecosystem at an altitude of 3 200 m to 3 400 m above sea level during the summers of 1988 and 1989. Indexes of deformation and axes of asymmetry due to wind shaping of Engelmann spruce (Picea engelmannii) and subalpine fir (Abies lasiocarpa) are related to wind speeds and directions on a 100 m × 100 m grid spacing over the 300 ha research site. Isotach and airflow patterns are drawn to represent climatological near-ground-level winds.A statistical analysis of the wind data and deformation indexes indicates that the indexes estimated independently by three of the authors were not significantly different at the F0.025 level. Two methods of calculating wind speeds were applied. At lower mean wind speeds in Engelmann spruce, results from the Wade-Hewson method were not significantly different from the Griggs-Putnam method at the F0.025 level. In slightly higher wind speeds in subalpine fir, the Wade-Hewson method produced significantly lower wind speeds than the Griggs-Putnam method.With 3 Figures  相似文献   

6.
Abstract

We compare the water‐parcel‐following and position‐reporting performance of two recent drifter designs. One (Tristar‐II ) uses the ARGOS satellite navigation system, has a very large drogue: float frontal area ratio and a small velocity error resulting from wind drag and surface wave forces. The second design uses an internal Loran receiver and transmits its position every half hour by radio. Its drogue: float frontal area ratio is much smaller than the TRISTAR's; we wished to know if this difference caused a significant change in drifter trajectory. Over a 78‐h joint deployment, the two designs had nearly identical trajectories. Measured drift speeds ranged from 10 to 30 cms?1. Wind speeds (an important potential biasing factor) ranged up to 15 ms?l. Net velocity difference over the full deployment was only 0.1 cms?1. Frontal trapping of both drifters during the latter part of the comparison period may have helped to minimize this difference. But we also observed little or no downwind slippage (the net velocity difference was less than 2 cm s?1 and was nearly normal to the wind and wave directions) during the first third of the deployment, when winds were strongest and the constancy of the water properties measured at drogue depth indicated no frontal trapping. Because the trajectories of the drifters were so similar, the most important differences in their performance were due to their position‐reporting and deployment characteristics. Initial deployment of the tristar was easier, and its positioning method and lower power requirement allow much longer untended deployments. The Loran design gave more frequent and precise positioning information and, therefore, better resolution of short‐time‐scale velocity fluctuations. It was also easier to find at sea, and to recover and redeploy.  相似文献   

7.
Abstract

The wind climate of the mountainous terrain in the southern Yukon is simulated using the Wind Energy Simulation Toolkit (WEST) developed by the Recherche en Prévision Numérique (RPN) group of Environment Canada and is compared to measurements in the field. WEST combines two models that operate at different spatial scales. The Mesoscale Compressible Community (MC2) model is a mesoscale numerical weather prediction model that produces simulations over large domains of the order of a thousand kilometres. The MC2 model uses long‐term synoptic scale wind climate data from the analysis of radiosonde and other observations to simulate mean wind fields at tens of metres above the ground using a horizontal resolution of a few kilometres. The mesoscale results are used as input to MS‐Micro/3 (Mason and Sykes (1979) version of the Jackson and Hunt (1975) model version for microcomputers/3‐dimensional; MS‐Micro hereafter), a more computer‐efficient, microscale model with simpler linearized momentum equations and a domain restricted to a few tens of kilometres with horizontal grid sizes of tens or hundreds of metres. MS‐Micro provides wind field results at specific wind generator hub heights (typically 30 to 50 m above ground level (AGL)) which are of interest to researchers and developers of wind farms.

WEST shows relatively strong correlations between its simulated long‐term mean wind speed and the measurements from ten wind energy monitoring stations. However, in the mountainous terrain of the Yukon, WEST tends to predict wind speeds which are about 40% too high. The model also produces erroneous wind directions and some were perpendicular to valley orientations. The most likely cause of the wind speed and direction errors is the substantially modified 5‐km grid‐spaced mesoscale terrain used in MC2. The WEST simulation was also found to double the wind speeds observed at airport stations and there was poor correlation between the simulated and observed wind speeds.

The bias in the model could be attributed to a number of factors, including the use of smoothed topography by the model, the discrepancy between the neutral atmosphere assumed in MS‐Micro and the normally observed stable atmosphere, the application of MS‐Micro to every third grid point of the MC2 output, abnormally high sea level wind speeds in the input climate data for MC2, and a certain degree of disagreement between the land surface characteristics used in the model and those found in the field.

At comparatively low computer cost, WEST predicts a wind climate map that compares favourably to the wind measurements made in several locations in the Yukon. However, the problem of the modified terrain in the mountainous regions is the most pressing problem and needs to be addressed before WEST is used in the mountainous regions of Canada.  相似文献   

8.
Data on the relationship of the surface wind to the geostrophic wind at Porton Down, Salisbury Plain, are presented for various stability conditions and analysed in the light of the Rossbynumber similarity theory. For near-neutral conditions, the geostrophic drag coefficients for geostrophic wind speeds 5 to 15 m s-1 are close to those found by other workers but at higher speeds the values are low. Comparisons of geostrophic and radar wind speeds for ⋍900-m height, suggest that undetectably small mean cyclonic curvatures of the trajectories of the air are responsible for this departure. A value of the geostrophic drag coefficient for the open sea at wind speeds around 8 m s-1 (neutral conditions) is deduced from recent observations of the drag in relation to the surface wind, combined with the ratios of 900-mb radar wind to surface wind obtained from the North Atlantic weather ship data tabulations of Findlater et al. (1966).  相似文献   

9.
Summary  Simulations alternatively assuming a real landscape with and without open-pit mines and grown settlements were performed with a non-hydrostatic meteorological model of the meso-β-scale to elucidate whether the atmospheric response to such land-use changes is sensitive to the direction and magnitude of geostrophic wind. The results of simulations with the same geostrophic wind conditions substantiate that the daily domain-averages of the variables of state hardly differ for the different landscape realizations, except for cloud- and precipitating particles. However, land-use changes may significantly affect the local conditions over and downwind of the altered surfaces. The significant differences in the cloud- and precipitating particles, however, are not bound to the surroundings of land-use changes. The vertical component of wind vector, which is modified by the different heating of converted land-use, strongly affects cloud- and precipitation formation by the interaction cloud microphysics-dynamics. The magnitude of atmospheric response changes under the various directions and speeds of geostrophic wind for most of the field quantities and fluxes. Received February 10, 1999  相似文献   

10.
近50年我国风向变化特征   总被引:8,自引:0,他引:8       下载免费PDF全文
利用我国基本和基准气象台站1956—2005年的一日4次风向和风速资料, 对近50年我国风向变化做了尝试性分析。分析发现:我国大部分地区年最大风向频率呈减小趋势, 其中西北、华南和西南地区最大风向频率减小趋势最为显著, 只有西部个别地区略有增加; 全国大部分地区年最大风向频率对应的风速均呈明显的减小趋势。同时, 年最大风向频率对应的风速减小趋势比年平均风速的减小趋势更为显著, 最大风向频率对应的风速减小是平均风速减小的主导因素; 我国冬季主要盛行的偏北风和夏季主要盛行的偏南风都呈明显的减小趋势。偏北风(冬季)和偏南风(夏季)的减小主要是亚洲冬季风和夏季风减弱造成的。  相似文献   

11.
浙江省几种灾害性大风近地面阵风系数特征   总被引:1,自引:1,他引:0       下载免费PDF全文
阵风特性研究是大风预报和服务的基础。基于2011-2013年浙江省自动气象站逐日逐10 min测风资料,分析了浙江省陆地和近海海面冷空气、热带气旋和强对流大风的阵风系数特征。结果表明:冷空气和热带气旋大风阵风系数空间分布基本相同,大风主要发生在近海海面和沿海地区,海面阵风系数一般小于1.5,等值线平行于海岸线且自西向东逐渐减小,陆地阵风系数一般大于2.0,山区可超过3.0,表现出地形对阵风系数的增强作用。强对流大风阵风系数明显高于业务规范平均值,发生地点遍及浙江省各地,但发生概率超过10%的站点主要位于沿海地区和近海海面。风向基本不影响阵风系数空间分布。冷空气和热带气旋站点阵风系数与海拔高度有较高正相关性。模糊聚类分析发现:浙江省400 m以上山区站与70 m以下的低海拔站点在阵风系数特征上分属不同空间类型;基于逐步回归建立站点阵风系数预报模型,检验表明:模糊聚类可帮助提高模型阵风系数预报能力。  相似文献   

12.
Abstract

A simple equation is developed for determining the ratio of wind speed above a rough vegetative canopy, such as a forest, to the wind speed over a relatively smooth surface, such as an airport, when both sites are subject to the same geostrophic wind. The usual assumptions for simplifying flow in the planetary boundary layer are followed and appropriate canopy parameters are introduced.

Values for the ratio of forest to airport wind speeds range from 0.4 to 0.7, with a typical value of 0.5.  相似文献   

13.
Summary In the present paper, an attempt is made for generalized the atmospheric diffusion operator. This can be accomplished by employing the realizability procedure, to identify a surface operator, that ensures self-adjointness’ of the atmospheric diffusion operator. The dispersion modeling in low wind speeds assumes importance because of the high frequency of occurrence and episodic nature of these poor diffusion conditions. A steady-state mathematical model for hermitized model has been calculated for the dispersion of air pollutants in low winds by taking into account the diffusion in the three coordinate directions and advection along the mean wind. The eddy diffusivities have been parameterized in terms of downwind distance for near source dispersion (Arya, 1995). The constants involved in this parameterization are the squares of intensities of turbulence. An analytical solution for resulting advection-diffusion equation with the physically relevant boundary conditions has been obtained. The solution has been used to simulate the field tracer data collected at IIT Delhi in low wind convective conditions.  相似文献   

14.
Abstract

Airborne measurements of mean wind velocity and turbulence in the atmospheric boundary layer under wintertime conditions of cold offshore advection suggest that at a height of 50 m the mean wind speed increases with offshore distance by roughly 20% over a horizontal scale of order 10 km. Similarly, the vertical gust velocity and turbulent kinetic energy decay on scales of order 3.5 km by factors of 1.5 and 3.2, respectively. The scale of cross‐shore variations in the vertical fluxes of heat and downwind momentum is also 10 km, and the momentum flux is found to be roughly constant to 300 m, whereas the heat flux decreases with height. The stability parameter, z/L (where z = 50 m and L is the local Monin‐Obukhov length), is generally small over land but may reach order one over the warm ocean. The magnitude and horizontal length scales associated with the offshore variations in wind speed and turbulence are reasonably consistent with model results for a simple roughness change, but a more sophisticated model is required to interpret the combined effects of surface roughness and heat flux contrasts between land and sea.

Comparisons between aircraft and profile‐adjusted surface measurements of wind speed indicate that Doppler biases of 1–2 m s?1 in the aircraft data caused by surface motions must be accounted for. In addition, the wind direction measurements of the Minimet anemometer buoy deployed in CASP are found to be in error by 25 ± 5°, possibly due to a misalignment of the anemometer vane. The vertical fluxes of heat and momentum show reasonably good agreement with surface estimates based on the Minimet data.  相似文献   

15.
Abstract

Low‐level cloud motion vectors determined by the National Oceanic and Atmospheric Administration (NOAA) and the University of Wisconsin in a 5‐degree square centred at 0°, 152°30’ W were intercompared with moored buoy wind measurements made at 0°, 152°W during April 1979‐February 1980. At the site of the intercomparison test the prevailing wind direction was easterly; monthly mean values of the meridional wind speed were less than 20% of the zonal component. The surface winds measured during the observational period were similar to the climatological‐mean wind conditions. Although the satellite wind speeds were larger than buoy wind speeds, as a priori expected, because of the vertical separation between the measurements, the comparison indicated that in the case of the zonal component there was a maximum usable frequency (muf) below which cloud drifts and surface wind vectors were 95% significantly correlated and the correlation was greater than the 50% noise level. The muf were 0.17 and 0.3 cpd for the NOAA and Wisconsin cloud motion vectors, respectively. At frequencies below the muf, an algorithm describing the frequency‐dependent differences of the rms zonal wind speed amplitudes was developed. Coherences involving the meridional wind speeds were too low for the estimation of a muf.  相似文献   

16.
利用2019年6月至12月威宁县边界层风廓线雷达数据和威宁探空数据,预设二者风速偏差<=3m/s、风向偏差<=20°为有效数据样本,研究两者在不同风速、风向范围和不同高度、时次、降水条件下风向、风速数据对比及相关性分析。结果表明:(1)风廓线雷达和探空的风速、风向均具有较好的正向相关性;(2)在不同高度下,且无论有无降水或任意时次,风速有效样本比率大体上高于风向有效样本比率,500米左右高度以下有效样本比率总是最小(不足50%),而中高层较大;(3)不同时次对风速、风向数据有效性影响不大;(4)有降水时风速、风向有效样本比率比无降水时偏小且变化剧烈;(5)除东北(20°-40°)和西南(200°-260°)风向外,其他方位风向数据一致性较差;(6)除大于24m/s的风速外其他大小风速均具有较好的一致性。  相似文献   

17.
Abstract

Changes to the Beaufort Sea shoreline occur due to the impact of storms and rising relative sea level. During the open‐water season (June to October), storm winds predominantly from the north‐west generate waves and storm surges which are effective in eroding thawing ice‐rich cliffs and causing overwash of gravel beaches. Climate change is expected to be enhanced in Arctic regions relative to the global mean and include accelerated sea‐level rise, more frequent extreme storm winds, more frequent and extreme storm surge flooding, decreased sea‐ice extent, more frequent and higher waves, and increased temperatures. We investigate historical records of wind speeds and directions, water levels, sea‐ice extent and temperature to identify variability in past forcing and use the Canadian Global Coupled Model ensembles 1 and 2 (CGCM1 and CGCM2) climate modelling results to develop a scenario forcing future change of Beaufort Sea shorelines. This scenario and future return periods of peak storm wind speeds and water levels likely indicate increased forcing of coastal change during the next century resulting in increased rates of cliff erosion and beach migration, and more extreme flooding.  相似文献   

18.
This study investigated multi-decadal variability in the wind resource over the Republic of Korea using the Weather Research and Forecasting (WRF) mesoscale meteorological model. Mesoscale simulations were performed for the period from November 1981 to November 2010. The typical wind climatology over the Korean Peninsula, which is influenced by both continental and oceanic features, was represented by the physics-based mesoscale simulations. Winter had windier conditions with northwesterly flows, whereas less windy with southwesterly flows appeared in summer. The annual mean wind speeds over the Republic of Korea were approximately 2 m s?1 with strong wind in mountainous areas, coastal areas, and islands. The multi-decadal variability in wind speed during the study period was characterized by significant increases (positive trend) over many parts of the study area, even though the various local trends appeared depending on the station locations. The longterm trend in the spatially averaged wind speed was approximately 0.002 m s?1 yr?1. The annual frequency of daily mean wind speeds over 5 m s?1 at the turbine hub height also increased during the study period throughout the Republic of Korea. The present study demonstrates that multi-decadal mesoscale simulations can be useful for climatological assessment of wind energy potential.  相似文献   

19.
王天义  朱克云  张杰  刘煦 《气象科技》2014,42(2):231-239
利用成都地区2010年8月和北京沙河地区2011年7—8月风廓线雷达以及多普勒天气雷达的风廓线探测资料,结合对应时段的天气现象相关记录,通过对比分析得到以下结论:①弱降水条件下,在300~2100m高度内,风廓线雷达与多普勒天气雷达探测具有很好的相关性,风向相关系数平均值为0.596,风速相关系数平均值为0.736,在做预报时两者可以同时应用,互为补充;②强降水天气条件下,风廓线雷达与多普勒天气雷达探测的风向、风速变化趋势基本一致,特别是在300~2100m之间各个高度上风向、风速相关性较好,风向相关系数平均值为0.573,风速相关系数为0.508,且风廓线雷达比多普勒天气雷达探测到的各层风向、风速变化更为详细、直观;③阴天条件下风廓线雷达与多普勒天气雷达的风向、风速相关性低层比高层好;④晴天条件下,风廓线雷达更适合用于预报和监测天气。  相似文献   

20.
The diurnal variation of surface winds off the coast of Oregon is described and compared with a recent analysis of winds off the coast of Peru. The Oregon wind speeds have a distinct 24-h periodicity, while the Peru wind speeds were reported to have an irregular 12-h variation. The long-and trans-shore components of both winds exhibit 24-h periodicities; the ratio of the long-shore to trans-shore diurnal amplitudes off Oregon is 2.8, twice the ratio found off Peru. Although meteorological conditions off Oregon were quasi-stationary during the period investigated, there were considerable day-to-day variations in diurnal amplitudes and phases. Diurnal amplitudes were found to be correlated with the daily mean long-shore winds.  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