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 125 毫秒
1.
基于激光雷达资料的小波变换法反演边界层高度的方法   总被引:1,自引:0,他引:1  
利用小波变换法反演边界层高度时,不同小波母函数的选取可能得到不同的边界层高度。因此,对构造的白天及夜间激光雷达后向散射信号理想廓线进行Haar小波协方差变换,并对后向散射信号梯度廓线进行Morlet与Mexican Hat小波变换反演边界层高度。结果表明,宜采用Haar函数与Mexican Hat函数作为小波母函数,其中Haar函数准确性优于Mexican Hat函数,而Mexican Hat函数更易稳定。同时为了进一步检验3种小波变换法的反演结果对小波振幅的敏感性,通过改变小波母函数的小波振幅,发现无论是理想廓线还是叠加扰动的廓线,较大的小波振幅易得到比较稳定准确的白天边界层高度与夜间混合层高度。  相似文献   

2.
陆面过程与大气边界层之间耦合关系是理解青藏高原热力效应的关键环节和难点之一。本文基于那曲高寒气候环境观测研究站2019年5月、 7月和10月地面及探空观测数据分析了青藏高原那曲地区地表能量收支及大气温湿垂直廓线的日变化和季节差异,探讨了该地区干湿季大气边界层高度的演变规律。结果表明,在5月观测期间内受日间净辐射强度变化的影响,对流边界层在晴天较高,为2842 m;阴天较低,为1481 m,强对流天气也可能使其在低层转变成稳定边界层。同时,位于近地层大气的感热和潜热交换为大气边界层的维持和发展提供了能量支持,位温和比湿垂直廓线能够正确反映出那曲地区大气边界层高度的季节性差异,对流边界层高度在5月最高、 10月次之、 7月最低,而稳定边界层在7月最高、 5月次之、 10月最低。  相似文献   

3.
根据2017、2019年7月塔克拉玛干沙漠腹地GPS探空和地面观测数据,利用位温廓线法等方法,对比分析了沙漠腹地夏季晴天和沙尘暴天气大气边界层结构变化特征。结果表明:晴天和沙尘暴天气大气边界层结构特征显著不同。晴天大气边界层各气象要素垂直分布较为均一,白天对流边界层深厚,高度接近5 km,夜间稳定边界层一般在500 m左右。沙尘暴天气边界层内位温和比湿垂直变化较小,风速较大,可达24.0 m/s,其白天对流边界层在1.5 km左右,夜间稳定边界层在1 km左右。晴天辐射强烈,地表升温迅速,湍流旺盛,是形成晴天深厚对流边界层的主要因素。大尺度天气系统冷平流的动力条件,以及云和沙尘减弱了到达地表的辐射强度是形成沙尘暴天气独特的大气边界层结构的主要因素。  相似文献   

4.
王倩茹  范广洲  赖欣  张永莉  朱伊 《气象》2018,44(3):396-407
本文利用探空气球加密观测资料和欧洲中心ERA-Interim 0.125°×0.125°再分析资料,对2016年8月29日午后降霰过程进行大气边界层特征分析,与同年8月26日典型晴天个例对比分析,结果表明:降霰过程前,温度0℃线随时间增加而升高,温度递减率分层现象显著,逆温层不明显,边界层多为对流不稳定层结;位温随高度增加而增加,随时间增加呈现5K·(2h)~(-1)的增加趋势;比湿随高度增加而减小,水汽含量较晴天更大;风速随高度呈多层次变化,近地层风速大于晴天同高度风速,边界层顶风速小于晴天边界层顶风速,风向始终以西风为主,随高度不存在大波动;降霰过程前云覆盖量大,云层厚度达4000m,存在复杂垂直运动,近地层为下沉运动,云层内为上升运动。综合以上可以看出那曲29日降霰过程前,08时边界层内存在明显过冷水,边界层顶波动极大,08时存在最大高度(3780m),10时为最低高度(850m)。位温随时间增加而上升,持续积累能量达6h,比湿大于晴天,边界层内风速大于晴天,且随高度变化不大,风向始终以西风为主,存在深厚的云系提供水汽,云内的上升运动和云下的下沉运动是促发霰过程的主要动力机制。  相似文献   

5.
利用2013年重庆多普勒天气雷达(SA)和风廓线雷达(TWP8-L)观测的垂直风廓线数据,对晴空、弱降水、一般性降水和强降水四种不同天气条件下垂直风廓线特征及其演变情况进行了分析。结果表明:(1)风廓线雷达的探测高度随降水增加逐渐增加;(2)晴空天气条件下,边界层(1 km以下)风向存在明显的日变化,夜间以偏东气流为主,白天以偏南气流为主,高空(3 km以上)为一致的偏西气流,风速较小;(3)弱降水天气条件下,边界层风向以偏东气流为主,相对较为杂乱,高空与晴空一致,中高层(1~3 km)以偏南气流为主;(4)一般性降水天气条件下,低层与弱降水较一致,而高空出现较一致的西南气流,有利于水汽输送,同时垂直切变具有较好的单一方向性,较有利于对流的发展和维持;(5)强降水天气条件下,风廓线雷达和多普勒雷达观测的垂直风廓线较为一致。降水前期风向随高度的增加逐渐由偏东气流转为偏西气流,有利于对流的触发;降水期间风切变具有很好的单一方向性并在中低层出现低空急流区,有利于对流系统的维持,同时西南气流厚度加深,也有利于水汽的输送;降水结束期风速减小,中低层风向也逐渐转为偏北气流,对流系统逐渐消亡。  相似文献   

6.
利用宜昌2007年12月10-25日的加密观测资料,分析了两次低值系统经过宜昌时大气边界层的温湿风廓线结构及其日变化特征。结果表明:位温廓线具有明显的日变化特征,对流边界层在白天出现和发展,其高度可达600m,而稳定边界层在夜间出现和发展,其高度可达300m,降水会抑制对流边界层和稳定边界层的发展;湿度廓线结构及其日变化与对流边界层的发展有关,总体上湿度随高度减小,贴近地面的薄层湿度随高度减小较快,而混合层内湿度随高度变化较小,出现降水时,近地层的湿度有明显增加,大气边界层内湿度随高度快速平稳减小;风速廓线结构比较复杂,总体上风速随高度增大,在大气边界层低层有时会出现一个风速极大值,风速廓线没有明显的日变化特征,大气边界层内风向变化较大,但以偏东风为主。  相似文献   

7.
赵玉春  王叶红 《大气科学》2020,44(2):371-389
利用2009~2017年7~9月福建省逐小时地面加密自动站资料和2015~2017年7~9月厦门站的探空资料,通过K均值聚类法和中尺度数值模式(WRF3.9.1.1版本)理想数值模拟,分析了我国东南沿岸及复杂山地(福建)后汛期降水日变化特征,揭示了地形热力环流以及海陆风环流在热对流降水日变化形成中的作用,探讨了环境温湿廓线及风垂直廓线对热对流降水日峰值强度和日峰值出现时间的影响。结果发现:我国东南沿岸复杂山地(福建)后汛期降水日变化受地形热力环流和海陆风环流的影响和调制,白天辐射加热在复杂山地形成的局地热力环流激发出对流降雨带,午后受海风环流的影响,对流降雨带组织发展达到峰值,之后随着地形热力环流和海风环流减弱雨带逐渐减弱。武夷山及周边复杂山地的降水日变化主要受地形热力环流的影响,在午后对流降水达到峰值,夜间减弱几近消失。理想数值试验进一步证实了我国东南沿岸复杂山地地形热力环流对对流降雨的触发以及海陆风环流在山地对流雨带组织发展中的作用,环境温湿廓线以及风垂直廓线对热对流降水日峰值强度以及日峰值出现的时间具有重要影响,其中环境温湿廓线的大气抬升凝结高度、大气可降水量、大气的对流不稳定度以及大气中低层湿度分布的不同,会影响热对流降水日峰值强度,并通过影响山地热力对流触发时间,改变热对流降水日峰值时间,而环境风垂直廓线的低层气流强度和方向、中低层垂直风切变的不同,会影响地形热力对流系统的启动、组织发展和移动等特征,进而影响热对流降水日峰值强度以及热对流降水日峰值时间。  相似文献   

8.
基于风廓线仪的华南地区夏季边界层湍流统计特征研究   总被引:1,自引:0,他引:1  
采用双权重算法,使用2015年6—8月我国东南部业务风廓线雷达资料,通过湍流脉动垂直速度方差和偏度的计算和分析,对晴空和低云主导情况下的边界层湍流特征以及中小尺度局地环流对于边界层湍流的影响进行研究。主要结论如下:(1)晴天情况下垂直速度标准差和垂直速度偏度都具有明显的日变化特征,湍流主要由下垫面加热驱动发展;(2)在低云主导情况下,湍流明显弱于晴天对流边界层的湍流强度,边界层内湍流的发展不仅受地面加热的影响,而且在边界层上部存在明显的自上而下发展的湍流,这主要是由于边界层顶云辐射冷却造成的;(3)除了上述两种情况,边界层湍流发展同时受到局地中小尺度环流或者天气系统的影响,因而呈现出更多的复杂性。   相似文献   

9.
青藏高原纳木错湖区大气边界层结构分析   总被引:8,自引:3,他引:5  
利用2007年8月8~19日期间系留气球低探空和GPS无线电探空资料,分析了纳木错湖区大气边界层高度、风、温、湿等要素的垂直结构。结果表明:纳木错湖的冷湖效应推迟了边界层湍流混合及对流边界层出现的时间,边界层高度日变化非常明显,对流边界层高度最高可达1750 m;在晴天条件下,边界层内湖陆风日变化非常明显,湖陆风控制范围常超过边界层高度,可达对流层中部;边界层内比湿变化呈V型变化,白天减小,夜间增大,早晨08:00出现峰值。  相似文献   

10.
胡宁  汪会 《热带气象学报》2019,35(5):681-693
2014年5月22日华南地区出现了一次大范围强对流天气过程,该过程中出现了两个中尺度对流系统(Mesoscale Convective System,MCS)MCS-A和MCS-B,两个MCS表现出迥异的形态特征,产生了不同的强对流天气。利用多源观测资料以及高分辨率数值模式分析了环境条件对于MCS形态特征的影响,结果表明:(1)广西夜间到凌晨边界层顶附近强盛的低空急流,使得MCS-A在北部山区出现后向建立(BB, back building)的形态特征,有利于大量级的短时强降水的出现;(2) MCS-A进入广西平原地区以后,强盛的边界层以上的低空急流使得能量垂直廓线的极大值在边界层高度以上,且风垂直切变特征不利于冷池前方的垂直运动发展,冷池前方无法连续触发对流,MCS-A逐渐演化成线状对流/层云伴随(TL/AS, Training Line/Adjoining stratiform)的形态特征,而后消亡;(3)在广东,能量极大值出现在大气底层,环境风廓线有利于冷池前方垂直运动发展,进而触发新的对流,新生成的MCS-B呈现典型的层云拖曳型(TS,trailing stratiform)形态,最终形成飑线,造成雷暴大风天气。   相似文献   

11.
Mixing depth structure and its evolution have been diagnosed from radar wind profiler data in the Chamonix and the Maurienne valleys (France) during summer 2003. The behaviour of refractive index structure parameter C n 2 peaks coupled with the vertical velocity variance σ w 2 was used to estimate the height of the mixed layer. Tethersonde vertical profiles were carried out to investigate the lower layers of the atmosphere in the range of approximately 400–500 m above ground level. The tethersonde device was especially useful to study the reversal of the valley wind system during the morning transition period. Specific features such as wind reversal and the convective mixed layer up to approximately the altitude of the surrounding mountains were documented. The wind reversal was observed to be much more sudden in the Maurienne valley than in the Chamonix valley  相似文献   

12.
We utilized a Doppler lidar to measure integral scale and coherence of vertical velocity w in the daytime convective boundary layer (CBL). The high resolution 2 μm wavelength Doppler lidar developed by the NOAA Environmental Technology Laboratory was used to detect the mean radial velocity of aerosol particles. It operated continuously in the zenith-pointing mode for several days in the summer 1996 during the “Lidars in Flat Terrain” experiment over level farmland in central Illinois. We calculated profiles of w integral scales in both the alongwind and vertical directions from about 390 m height to the CBL top. In the middle of the mixed layer we found, from the ratio of the w integral scale in the vertical to that in the horizontal direction, that the w eddies are squashed by a factor of about 0.65 as compared to what would be the case for isotropic turbulence. Furthermore, there is a significant decrease of the vertical integral scale with height. The integral scale profiles and vertical coherence show that vertical velocity fluctuations in the CBL have a predictable anisotropic structure. We found no significant tilt of the thermal structures with height in the middle part of the CBL.The National Center for Atmospheric Research is sponsored by the National Science Foundation.  相似文献   

13.
Nine profiles of the temperature structure parameter C T 2 and the standard deviation of vertical velocity fluctuations ( w) in the convective boundary layer (CBL) were obtained with a monostatic Doppler sodar during the second intensive field campaign of the First ISLSCP Field Experiment in 1987. The results were analyzed by using local similarity theory. Local similarity curves depend on four parameters: the height of the mixed layer (z i ), the depth of the interfacial layer (), and the temperature fluxes at the top of the mixed layer (Q i ) and the surface (Q o). Values of these parameters were inferred from sodar data by using the similarity curve for C T 2 and observations at three points in its profile. The effects of entrainment processes on the profiles of C T 2 and wnear the top of the CBL appeared to be described well by local similarity theory. Inferred estimates of surface temperature flux, however, were underestimated in comparison to fluxes measured by eddy correlation. The measured values of wappeared to be slightly smaller than estimates based on available parmeterizations. These discrepancies might have been caused by experimental error or, more likely, by the distortion of turbulence structure above the site by flow over the nonuniform terrain at the observation site.  相似文献   

14.
We question the correlation between vertical velocity (w) on the one hand and the occurrence of convective plumes in lidar reflectivity (i.e. range corrected backscatter signal Pz 2) and depolarization ratio (Δ) on the other hand in the convective boundary layer (CBL). Thermal vertical motion is directly investigated using vertical velocities measured by a ground-based Doppler lidar operating at 2 μm. This lidar provides also simultaneous measurements of lidar reflectivity. In addition, a second lidar 200 m away provides reflectivities at 0.53 and 1 μm and depolarization ratio at 0.53 μm. The time series from the two lidars are analyzed in terms of linear correlation coefficient (ρ). The main result is that the plume-like structures provided by lidar reflectivity within the CBL as well as the CBL height are not a clear signature of updrafts. It is shown that the lidar reflectivity within the CBL is frequently anti-correlated (ρ (w, Pz 2 )) with the vertical velocity. On the contrary, the correlation coefficient between the depolarization ratio and the vertical velocity ρ (w, Δ ) is always positive, showing that the depolarization ratio is a fair tracer of updrafts. The importance of relative humidity on the correlation coefficient is discussed. An erratum to this article can be found at  相似文献   

15.
The morning development of the daytime convective boundary layer (CBL) during fine weather has been observed with an acoustic Doppler sodar of the C.R.P.E. In particular, the vertical profile of the vertical velocity third-order statistic W* 3 has been obtained. This quantity is a maximum near 0.3z I where z I, is the height of the CBL. The histogram of vertical velocity in the CBL shows a relationship between W 3 and the convective velocity W * and is useful for convective plume determination.  相似文献   

16.
The friction velocity, the surface heat flux and the height of the Atmospheric Boundary Layer (ABL) are important parameters. In this work, vertical velocity variance ( w 2 ) and wind velocity structure parameter (C v 2 ) profiles estimated by acoustic sounder measurements are used, along with similarity relations, to estimate these parameters in the unstable Atmospheric Boundary Layer and the friction velocity in the stable one. The data were collected by two acoustic sounders with different height range and resolution under various atmospheric conditions (stability) and at two experimental sites in different terrain. The C v 2 profiles are estimated using gate difference of the vertical velocity measurements and the assumption of local isotropy. The vertical velocity data are corrected for the significant effects of noisy measurements and sampling volume averaging on the w 2 and C v 2 estimations using original techniques that are presented in this work. The results of the similarity method using acoustic sounder data are compared against estimates of the corresponding atmospheric parameters obtained from direct measurements. The comparison confirms the ability of the method to provide reasonably accurate estimates of these parameters especially in the middle of the day.  相似文献   

17.
We have analyzed measurements of vertical velocity w statistics with the NOAA high resolution Doppler lidar (HRDL) from about 390 m above the surface to the top of the convective boundary layer (CBL) over a relatively flat and uniform agricultural surface during the Lidars-in-Flat-Terrain (LIFT) experiment in 1996. The temporal resolution of the zenith-pointing lidar was about 1 s, and the range-gate resolution about 30 m. Vertical cross-sections of w were used to calculate second- to fourth-moment statistics of w as a function of height throughout most of the CBL. We compare the results with large-eddy simulations (LES) of the CBL and with in situ aircraft measurements. A major cause of the observed case-to-case variability in the vertical profiles of the higher moments is differences in stability. For example, for the most convective cases, the skewness from both LES and observations changes more with height than for cases with more shear, with the observations changing more with stability than the LES. We also found a decrease in skewness, particularly in the upper part of the CBL, with an increase in LES grid resolution.  相似文献   

18.
Temperature variance and temperature power spectra in the unstable surface layer have always presented a problem to the standard Monin-Obukhov similarity model. Recently that problem has intensified with the demonstration by Smedman et al. (2007, Q J Roy Meteorol Soc 133: 37–51) that temperature spectra and heat-flux cospectra can have two distinct peaks in slightly unstable conditions, and by McNaughton et al. (2007, Nonlinear Process Geophys 14: 257–271) who showed that the wavenumber of the peak of temperature spectra in a convective boundary layer (CBL), closely above the surface friction layer (SFL), can be sensitive to the CBL depth, z i. Neither the two-peak form at slight instability nor the dependence of peak position on z i at large instability is compatible with the Monin-Obukhov model. Here we examine the properties of temperature spectra and heat-flux cospectra from between these extremes, i.e. from within the unstable SFL, in two experiments. The analysis is based on McNaughton’s model of the turbulence structure in the SFL. According to this model, heat is transported through most of the SFL by sheet plumes, created by the action of impinging outer eddies. The smallest and most effective of these outer eddies have sizes that scale on SFL depth, z s. The z s-scale eddies and plumes are organised within the overall convection pattern in the CBL, and in turn they organise the motion of smaller eddies within the SFL, whose sizes scale on height, z. The main experimental results are: (1) the peak amplitudes of the temperature spectra in the SFL are collapsed with a scaling factor (zsz)1/3eo2/3{(z_{\rm s}z)^{1/3}\varepsilon_{\rm o}^{2/3}} divided by the square of the surface temperature flux, where eo{\varepsilon_{\rm o}} is the dissipation rate of turbulent energy in the outer CBL (above the SFL); (2) the peak wavenumbers of the temperature spectra are collapsed with the mixed length scale (z i z s)1/2; (3) the peak wavenumbers of the heat-flux cospectra are collapsed with the doubly-mixed length scale (z i z s)1/4 z 1/2; (4) for z/z s < 0.03, the peak in the cospectrum is replaced by another peak at a wavenumber about a magnitude larger. This peak’s position scales on z; (5) all these findings are consistent with the observations of Smedman et al.  相似文献   

19.
Some aspects of determining the stable boundary layer depth from sodar data   总被引:3,自引:2,他引:1  
The question of estimating the height of the stable boundary layer (SBL) based on digitalized vertical profiles of sodar signal intensity has been re-examined. A simple one-dimensional numerical boundary-layer model is used to compute vertical profiles of the temperature structure parameterC T 2 . It is shown that especially at the beginning of the night (when stratification is weak) one can not expect any significant profile structure in the upper part of the SBL if its depth is determined in terms of common turbulent height scales. From this it is concluded that the SBL-height will be underestimated early in the night when derived from the maximum gradient in the signal intensity profiles. Later in the night in contrast, the computations often show elevated maxima or even zones with reduced, and above them enhanced, vertical gradients ofC T 2 , from which a SBL-height can be deduced that compares well with other common height scales. The computed profiles ofC T 2 are shown to be in qualitative agreement with observed profiles of sodar signal intensity for several analysed cases from the HAPEX-MOBILHY experiment.Comparing different SBL-depth scales with sodar observations, it is demonstrated that most of them are often closely related to a sodar-derived SBL-height only during certain phases of the night. Thus the sodar-SBL-height can, after a transition period, be perhaps associated with the lower turbulent layer of the growing surface inversion during the first part and with the height of the low-level wind maximum during the second part of the night.  相似文献   

20.
We investigated the flux footprints of receptors at different heights in the convective boundary layer (CBL). The footprints were derived using a forward Lagrangian stochastic (LS) method coupled with the turbulent fields from a large-eddy simulation model. Crosswind-integrated flux footprints shown as a function of upstream distances and sensor heights in the CBL were derived and compared using two LS particle simulation methods: an instantaneous area release and a crosswind linear continuous release. We found that for almost all sensor heights in the CBL, a major positive flux footprint zone was located close to the sensor upstream, while a weak negative footprint zone was located further upstream, with the transition band in non-dimensional upwind distances −X between approximately 1.5 and 2.0. Two-dimensional (2D) flux footprints for a point sensor were also simulated. For a sensor height of 0.158 z i, where z i is the CBL depth, we found that a major positive flux footprint zone followed a weak negative zone in the upstream direction. Two even weaker positive zones were also present on either side of the footprint axis, where the latter was rotated slightly from the geostrophic wind direction. Using CBL scaling, the 2D footprint result was normalized to show the source areas and was applied to real parameters obtained using aircraft-based measurements. With a mean wind speed in the CBL of U = 5.1 m s−1, convective velocity of w * = 1.37 m s−1, CBL depth of z i = 1,000 m, and flight track height of 159 m above the surface, the total flux footprint contribution zone was estimated to range from about 0.1 to 4.5 km upstream, in the case where the wind was perpendicular to the flight track. When the wind was parallel to the flight track, the total footprint contribution zone covered approximately 0.5 km on one side and 0.8 km on the other side of the flight track.  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