Advanced Search
Article Contents

Robust Solution for Boundary Layer Height Detections with Coherent Doppler Wind Lidar


doi: 10.1007/s00376-021-1068-0

  • Although coherent Doppler wind lidar (CDWL) is promising in detecting boundary layer height (BLH), differences between BLH results are observed when different CDWL measurements are used as tracers. Here, a robust solution for BLH detections with CDWL is proposed and demonstrated: mixed layer height (MLH) is retrieved best from turbulent kinetic energy dissipation rate (TKEDR), while stable boundary layer height (SBLH) and residual layer height (RLH) can be retrieved from carrier-to-noise ratio (CNR). To study the cause of the BLH differences, an intercomparison experiment is designed with two identical CDWLs, where only one is equipped with a stability control subsystem. During the experiment, it is found that the CNR could be distorted by instrument instability because the coupling efficiency from free-space to the polarization-maintaining fiber of the telescope is sensitive to the surrounding environment. In the ML, a bias up to 2.13 km of the MLH from CNR is found, which is caused by the CNR deviation. In contrast, the MLH from TKEDR is robust as long as the accuracy of wind is guaranteed. In the SBL (RL), the CNR is found capable to retrieve SBLH and RLH simultaneously and robustly. This solution is tested during an observation period over one month. Statistical analysis shows that the root-mean-square errors (RMSE) in the MLH, SBLH, and RLH are 0.28 km, 0.23 km, and 0.24 km, respectively.
    摘要: 尽管相干多普勒测风激光雷达在边界层高度探测方面有很好的应用前景,但当采用不同的观测数据时,边界层高度的探测结果存在差异。本文提出了一种基于相干多普勒测风激光雷达探测边界层高度的稳健方法:用湍流动能耗散率确定混合层高度,用载噪比确定稳定边界层高度和残留层高度。为了研究边界层探测高度差异的成因,本文采用了两台相同机制的雷达作对比,并为其中一台配备了温控装置。实验发现,由于自由空间到望远镜保偏光纤的耦合效率对环境敏感,仪器不稳定性将导致载噪比误差,进而导致基于载噪比确定的混合层高度出现高达2.13 km的偏移。相比之下,只要风场观测数据是准确的,基于湍流动能耗散率就可以稳健地确定混合层高度。实验还发现载噪比可以稳健地确定稳定边界层高度和残留层高度。为了验证该方法的有效性,本文进行了为期一个月的实验,统计结果表明:混合层高度,稳定边界层高度和残留层高度的均方根误差分别为0.28 km,0.23 km和0.24 km。
  • 加载中
  • Figure 1.  The two CDWLs used in the experiment.

    Figure 2.  The 7-min average profiles centered at 0853 LST 27 September 2019 of CDWL1: (a) Range-corrected CNR after normalization and its (b) HWCT result, (c) original CNR, (d) lg(TKEDR). Centered at 1311 LST 27 September: (e)–(h). The black dots denote the BLHs retrieved by the HWCT method. The blue dots are the BLHs retrieved by the CNR threshold method. The red squares are the BLHs retrieved by the TKEDR threshold method.

    Figure 3.  The linear regression of the BLH retrieved by using different CNR thresholds as a function of that by the HWCT method: (a) Determination coefficient (R2) (black dashed line) and root-mean-square error (RMSE) (red line) for SBLHCNR/MLHCNR. (b) Similar to (a) but for RLHCNR.

    Figure 4.  Lidar observation results during 27–28 September 2019: (a) CNR, (b) horizontal wind speed, (c) horizontal wind direction, (d) vertical wind speed, (e) lg(TKEDR) of CDWL1. (f)–(j) Results of CDWL2. The horizontal wind direction is defined as 0° for northerly wind, rotating clockwise. Negative vertical wind speed denotes rising motion.

    Figure 5.  (a)–(j) Radial CNR profiles at different measurement moments on 27 September. (k)–(t) The corresponding radial velocities and their standard deviation profiles. The △t values denote the moment difference between CDWL1 (blue line) and CDWL2 (red dashed line).

    Figure 6.  Statistical analysis of the wind differences between the two lidars on 27 September: Histogram distributions of (a) horizontal wind speed difference, (b) horizontal wind direction difference, and (c) vertical wind speed difference. The dashed lines denote Gaussian fits.

    Figure 7.  BLH retrieval results during 27–28 September: (a) Results of CDWL1. (b) Results of CDWL2. SBLH/MLHCNR (blue dots) and RLHCNR (black open circles) are retrieved by the CNR method. MLHTKEDR (red squares) are retrieved by the TKEDR method.

    Figure 8.  Scatter diagrams of the BLHs derived from the two lidars: (a) MLHCNR and (b) MLHTKEDR in the ML, (c) SBLHCNR and (d) RLHCNR in the SBL (RL). The color-shaded dots denote normalized density. The dashed lines denote 1:1 lines. The y–x is inserted in the bottom right corner of each panel.

    Table 1.  Key parameters of the CDWLs.

    ParameterValue
    LaserWavelength (nm)1548
    Pulse energy (μJ)100
    Repetition frequency (kHz)10
    TelescopeDiameter (mm)80
    ScannerZenith scanning range (°)0−90
    Azimuth scanning range (°)0−360
    Radial temporal resolution (s)2
    DownLoad: CSV

    Table 2.  Linear regression of the BLH derived from CDWL2 as a function of that from CDWL1.

    BLHR2RMSE (km)
    MLMLHCNR−2.501.08
    MLHTKEDR0.900.28
    SBL (RL)SBLHCNR0.870.23
    RLHCNR0.950.24
    R2: determination coefficient between dots and 1:1 line; RMSE: root-mean-square error.
    DownLoad: CSV
  • Banakh, V. A., A. Brewer, E. L. Pichugina, and I. N. Smalikho, 2010: Measurements of wind velocity and direction with coherent Doppler lidar in conditions of a weak echo signal. Atmospheric and Oceanic Optics, 23, 381−388, https://doi.org/10.1134/s1024856010050076.
    Banakh, V. A., I. N. Smalikho, and A. V. Falits, 2017: Estimation of the turbulence energy dissipation rate in the atmospheric boundary layer from measurements of the radial wind velocity by micropulse coherent Doppler lidar. Optics Express, 25, 22 679−22 692, https://doi.org/10.1364/oe.25.022679.
    Banakh, V. A., I. N. Smalikho, and A. V. Falits, 2020: Estimation of the height of turbulent mixing layer from data of Doppler lidar measurements using conical scanning by a probe beam. Atmospheric Measurement Techniques Discussions, https://doi.org/10.5194/amt-2020-259.
    Brooks, I. M., 2003: Finding boundary layer top: Application of a wavelet covariance transform to lidar backscatter profiles. J. Atmos. Oceanic Technol., 20, 1092−1105, https://doi.org/10.1175/1520-0426(2003)020<1092:fbltao>2.0.co;2.
    Chambers, D. M., 1997: Modeling heterodyne efficiency for coherent laser radar in the presence of aberrations. Optics Express, 1, 60−67, https://doi.org/10.1364/oe.1.000060.
    Coen, M. C., C. Praz, A. Haefele, D. Ruffieux, P. Kaufmann, and B. Calpini, 2014: Determination and climatology of the planetary boundary layer height above the Swiss plateau by in situ and remote sensing measurements as well as by the COSMO-2 model. Atmospheric Chemistry and Physics, 14, 13 205−13 221, https://doi.org/10.5194/acp-14-13205-2014.
    Cohn, S. A., and W. M. Angevine, 2000: Boundary layer height and entrainment zone thickness measured by lidars and wind-profiling radars. J. Appl. Meteorol. Climatol., 39, 1233−1247, https://doi.org/10.1175/1520-0450(2000)039<1233:blhaez>2.0.co;2.
    Emeis, S., C. Jahn, C. Münkel, C. Münsterer, and K. Schäfer, 2007: Multiple atmospheric layering and mixing-layer height in the Inn valley observed by remote sensing. Meteor. Z., 16, 415−424, https://doi.org/10.1127/0941-2948/2007/0203.
    Emeis, S., K. Schäfer, and C. Münkel, 2008: Surface-based remote sensing of the mixing-layer height−A review. Meteor. Z., 17, 621−630, https://doi.org/10.1127/0941-2948/2008/0312.
    Flamant, C., J. Pelon, P. H. Flamant, and P. Durand, 1997: Lidar determination of the entrainment zone thickness at the top of the unstable marine atmospheric boundary layer. Bound. -Layer Meteorol., 83, 247−284, https://doi.org/10.1023/a:1000258318944.
    Guo, J. P., and Coauthors, 2016: The climatology of planetary boundary layer height in China derived from radiosonde and reanalysis data. Atmospheric Chemistry and Physics, 16, 13 309−13 319, https://doi.org/10.5194/acp-16-13309-2016.
    Hooper, W. P., and E. W. Eloranta, 1986: Lidar measurements of wind in the planetary boundary layer: The method, accuracy and results from joint measurements with radiosonde and kytoon. J. Appl. Meteorol. Climatol., 25, 990−1001, https://doi.org/10.1175/1520-0450(1986)025<0990:lmowit>2.0.co;2.
    Huang, M., and Coauthors, 2017: Estimate of boundary-layer depth over Beijing, China, using Doppler lidar data during SURF-2015. Bound. -Layer Meteorol., 162, 503−522, https://doi.org/10.1007/s10546-016-0205-2.
    Kaimal, J. C., and J. J. Finnigan, 1994: Atmospheric Boundary Layer Flows: Their Structure and Measurement. Oxford University Press, 7−9.
    Leung, M. Y. T., W. Zhou, C. M. Shun, and P. W. Chan, 2018: Large-scale circulation control of the occurrence of low-level turbulence at Hong Kong international airport. Adv. Atmos. Sci., 35, 435−444, https://doi.org/10.1007/s00376-017-7118-y.
    Lewis, J. R., E. J. Welton, A. M. Molod, and E. Joseph, 2013: Improved boundary layer depth retrievals from MPLNET. J. Geophys. Res., 118, 9870−9879, https://doi.org/10.1002/jgrd.50570.
    Li, H., Y. Yang, X. M. Hu, Z. W. Huang, G. Y. Wang, B. D. Zhang, and T. J. Zhang, 2017: Evaluation of retrieval methods of daytime convective boundary layer height based on lidar data. J. Geophys. Res., 122, 4578−4593, https://doi.org/10.1002/2016jd025620.
    Li, Z. Q., and Coauthors, 2016: Remote sensing of atmospheric particulate mass of dry PM2.5 near the ground: Method validation using ground-based measurements. Remote Sensing of Environment, 173, 59−68, https://doi.org/10.1016/j.rse.2015.11.019.
    Luo, T., Z. E. Wang, D. M. Zhang, and B. Chen, 2016: Marine boundary layer structure as observed by A-train satellites. Atmospheric Chemistry and Physics, 16, 5891−5903, https://doi.org/10.5194/acp-16-5891-2016.
    Manninen, A., T. Marke, M. Tuononen, and E. J. O'Connor, 2018: Atmospheric boundary layer classification with Doppler lidar. J. Geophys. Res., 123, 8172−8189, https://doi.org/10.1029/2017jd028169.
    Melfi, S. H., J. D. Spinhirne, S. H. Chou, and S. P. Palm, 1985: Lidar observations of vertically organized convection in the planetary boundary layer over the ocean. J. Appl. Meteorol. Climatol., 24, 806−821, https://doi.org/10.1175/1520-0450(1985)024<0806:loovoc>2.0.co;2.
    O’Connor, E. J., A. J. Illingworth, I. M. Brooks, C. D. Westbrook, R. J. Hogan, F. Davies, and B. J. Brooks, 2010: A method for estimating the turbulent kinetic energy dissipation rate from a vertically pointing Doppler lidar, and independent evaluation from balloon-borne in situ measurements. J. Atmos. Oceanic Technol., 27, 1652−1664, https://doi.org/10.1175/2010jtecha1455.1.
    Peña, A., S. E. Gryning, and A. N. Hahmann, 2013: Observations of the atmospheric boundary layer height under marine upstream flow conditions at a coastal site. J. Geophys. Res., 118, 1924−1940, https://doi.org/10.1002/jgrd.50175.
    Sathe, A., and J. Mann, 2013: A review of turbulence measurements using ground-based wind lidars. Atmospheric Measurement Techniques, 6, 3147−3167, https://doi.org/10.5194/amt-6-3147-2013.
    Schween, J. H., A. Hirsikko, U. Löhnert, and S. Crewell, 2014: Mixing-layer height retrieval with ceilometer and Doppler lidar: From case studies to long-term assessment. Atmospheric Measurement Techniques, 7, 3685−3704, https://doi.org/10.5194/amt-7-3685-2014.
    Seibert, P., F. Beyrich, S.-E. Gryning, S. Joffre, A. Rasmussen, and P. Tercier, 2000: Review and intercomparison of operational methods for the determination of the mixing height. Atmos. Environ., 34, 1001−1027, https://doi.org/10.1016/s1352-2310(99)00349-0.
    Seidel, D. J., C. O. Ao, and K. Li, 2010: Estimating climatological planetary boundary layer heights from radiosonde observations: Comparison of methods and uncertainty analysis. J. Geophys. Res., 115, D16113, https://doi.org/10.1029/2009jd013680.
    Smalikho, I. N., and V. A. Banakh, 2017: Measurements of wind turbulence parameters by a conically scanning coherent Doppler lidar in the atmospheric boundary layer. Atmospheric Measurement Techniques, 10, 4191−4208, https://doi.org/10.5194/amt-10-4191-2017.
    Steyn, D. G., M. Baldi, and R. M. Hoff, 1999: The detection of mixed layer depth and entrainment zone thickness from lidar backscatter profiles. J. Atmos. Oceanic Technol., 16, 953−959, https://doi.org/10.1175/1520-0426(1999)016<0953:tdomld>2.0.co;2.
    Stull, R. B., 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic Publishers, 9−16.
    Su, T. N., Z. Q. Li, and R. Kahn, 2020: A new method to retrieve the diurnal variability of planetary boundary layer height from lidar under different thermodynamic stability conditions. Remote Sens. Environ., 237, 111519, https://doi.org/10.1016/j.rse.2019.111519.
    Vakkari, V., E. J. O'Connor, A. Nisantzi, R. E. Mamouri, and D. G. Hadjimitsis, 2015: Low-level mixing height detection in coastal locations with a scanning Doppler lidar. Atmospheric Measurement Techniques, 8, 1875−1885, https://doi.org/10.5194/amt-8-1875-2015.
    Wang, C., and Coauthors, 2017: 1.5 μm polarization coherent lidar incorporating time-division multiplexing. Optics Express, 25, 20 663−20 674, https://doi.org/10.1364/oe.25.020663.
    Wang, C., H. Y. Xia, Y. P. Liu, S. F. Lin, and X. K. Dou, 2018: Spatial resolution enhancement of coherent Doppler wind lidar using joint time–frequency analysis. Optics Communications, 424, 48−53, https://doi.org/10.1016/j.optcom.2018.04.042.
    Wang, C., and Coauthors, 2019: Relationship analysis of PM2.5 and boundary layer height using an aerosol and turbulence detection lidar. Atmospheric Measurement Techniques, 12, 3303−3315, https://doi.org/10.5194/amt-12-3303-2019.
    Wei, T. W., and Coauthors, 2019: Simultaneous wind and rainfall detection by power spectrum analysis using a VAD scanning coherent Doppler lidar. Optics Express, 27, 31 235−31 245, https://doi.org/10.1364/oe.27.031235.
    Yang, Y. J., and Coauthors, 2020: Diurnal evolution of the wintertime boundary layer in urban Beijing, China: Insights from doppler lidar and a 325-m meteorological tower. Remote Sensing, 12, 3935, https://doi.org/10.3390/rs12233935.
    Yuan, J. L., H. Y. Xia, T. W. Wei, L. Wang, B. Yue, and Y. B. Wu, 2020: Identifying cloud, precipitation, windshear, and turbulence by deep analysis of the power spectrum of coherent Doppler wind lidar. Optics Express, 28, 37 406−37 418, https://doi.org/10.1364/oe.412809.
  • [1] Zexu Luo, Xiaoquan Song, Jiaping Yin, Zhichao Bu, Yubao Chen, Yongtao Yu, Zhenlu Zhang, 2024: Comparison and Verification of Coherent Doppler Wind Lidar and Radiosonde in Beijing Urban Area, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-024-3240-9
    [2] Zhao Ming, Xu Yinzi, Wu Rongsheng, 1989: The Wind Structure in Planetary Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 6, 365-376.  doi: 10.1007/BF02661542
    [3] WANG Gaili, LIU Liping, LIU Zhishen, LU Bo, MU Rong, 2011: The Application of Sea-Surface Wind Detection with Doppler Lidar in Olympic Sailing, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 1471-1480.  doi: 10.1007/s00376-011-9189-5
    [4] HU Yinqiao, ZUO Hongchao, 2003: The Influence of Convergence Movement on Turbulent Transportation in the Atmospheric Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 794-798.  doi: 10.1007/BF02915404
    [5] SUN Jianning, JIANG Weimei, CHEN Ziyun, YUAN Renmin, 2005: A Laboratory Study of the Turbulent Velocity Characteristics in the Convective Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 770-780.  doi: 10.1007/BF02918721
    [6] Liu Shikuo, Huang Wei, Rong Pingping, 1992: Effects of Turbulent Dispersion of Atmospheric Balance Motions of Planetary Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 147-156.  doi: 10.1007/BF02657505
    [7] Tan Zhemin, Wang Yuan, 2002: Wind Structure in an Intermediate Boundary Layer Model Based on Ekman Momentum Approximation, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 266-278.  doi: 10.1007/s00376-002-0021-0
    [8] TANG Xiaodong, TAN Zhemin, 2006: Boundary-Layer Wind Structure in a Landfalling Tropical Cyclone, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 737-749.  doi: 10.1007/s00376-006-0737-3
    [9] Jun ZOU, Jianning SUN, Aijun DING, Minghuai WANG, Weidong GUO, Congbin FU, 2017: Observation-based Estimation of Aerosol-induced Reduction of Planetary Boundary Layer Height, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1057-1068.  doi: 10.1007/s00376-016-6259-8
    [10] Yidan SI, Shenshen LI, Liangfu CHEN, Chao YU, Zifeng WANG, Yang WANG, Hongmei WANG, 2018: Validation and Spatiotemporal Distribution of GEOS-5-Based Planetary Boundary Layer Height and Relative Humidity in China, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 479-492.  doi: 10.1007/s00376-017-6275-3
    [11] Shen YAN, Jie XIANG, Huadong DU, 2019: Determining Atmospheric Boundary Layer Height with the Numerical Differentiation Method Using Bending Angle Data from COSMIC, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 303-312.  doi: 10.1007/s00376-018-7308-2
    [12] Liu Shikuo, Peng Weihong, Huang Feng, Chi Dongyan, 2002: Effects of Turbulent Dispersion on the Wind Speed Profile in the Surface Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 794-806.  doi: 10.1007/s00376-002-0045-5
    [13] Sun-Hee SHIN, Kyung-Ja HA, 2009: Implementation of Turbulent Mixing over a Stratocumulus-Topped Boundary Layer and Its Impact in a GCM, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 995-1004.  doi: 10.1007/s00376-009-8145-0
    [14] Yu SHI, Qingcun ZENG, Fei HU, Weichen DING, Zhe ZHANG, Kang ZHANG, Lei LIU, 2023: Different Turbulent Regimes and Vertical Turbulence Structures of the Urban Nocturnal Stable Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1089-1103.  doi: 10.1007/s00376-022-2198-8
    [15] Zhu Ping, Xu Xiaojin, Li Xingsheng, 1992: A Numerical Study of Second-Order Turbulent Moments in the Stably Stratified Nocturnal Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 201-212.  doi: 10.1007/BF02657510
    [16] Xu Yinzi, 1988: THE WIND IN THE BAROCLINIC BOUNDARY LAYER WITH THREE SUBLAYERS INCORPORATING THE WEAK NON-LINEAR EFFECT, ADVANCES IN ATMOSPHERIC SCIENCES, 5, 483-497.  doi: 10.1007/BF02656793
    [17] YUE Ping, ZHANG Qiang, WANG Runyuan, LI Yaohui, WANG Sheng, 2015: Turbulence Intensity and Turbulent Kinetic Energy Parameters over a Heterogeneous Terrain of Loess Plateau, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1291-1302.  doi: 10.1007/s00376-015-4258-9
    [18] Li Xin, Hu Fei, Pu Yifen, M.H.Al-Jiboori, Hu Zhaoxia, Hong Zhongxiang, 2002: Identification of Coherent Structures of Turbulence at the Atmospheric Surface Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 687-698.  doi: 10.1007/s00376-002-0008-x
    [19] Zhao Yanzeng, Hu Yuliang, Zhao Hongjie, 1984: INTEGRATION METHOD AND RATIO METHOD FOR RETRIEVING EXTINCTION COEFFICIENT FROM LIDAR SIGNALS, ADVANCES IN ATMOSPHERIC SCIENCES, 1, 53-75.  doi: 10.1007/BF03187616
    [20] Zhou Xiuji, Qiu Jinhuan, 1984: CHARACTERISTICS OF ATMOSPHERIC EXTINCTION-TO-BACKSCATTERING RATIO IN RUBY LIDAR MEASUREMENTS, ADVANCES IN ATMOSPHERIC SCIENCES, 1, 179-198.  doi: 10.1007/BF02678130

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 10 February 2021
Manuscript revised: 21 April 2021
Manuscript accepted: 01 June 2021
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Robust Solution for Boundary Layer Height Detections with Coherent Doppler Wind Lidar

    Corresponding author: Haiyun XIA, hsia@ustc.edu.cn
  • 1. School of Earth and Space Science, University of Science and Technology of China, Hefei 230026, China
  • 2. Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei 230026, China
  • 3. CAS Center for Excellence in Comparative Planetology, University of Science and Technology of China, Hefei 230026, China

Abstract: Although coherent Doppler wind lidar (CDWL) is promising in detecting boundary layer height (BLH), differences between BLH results are observed when different CDWL measurements are used as tracers. Here, a robust solution for BLH detections with CDWL is proposed and demonstrated: mixed layer height (MLH) is retrieved best from turbulent kinetic energy dissipation rate (TKEDR), while stable boundary layer height (SBLH) and residual layer height (RLH) can be retrieved from carrier-to-noise ratio (CNR). To study the cause of the BLH differences, an intercomparison experiment is designed with two identical CDWLs, where only one is equipped with a stability control subsystem. During the experiment, it is found that the CNR could be distorted by instrument instability because the coupling efficiency from free-space to the polarization-maintaining fiber of the telescope is sensitive to the surrounding environment. In the ML, a bias up to 2.13 km of the MLH from CNR is found, which is caused by the CNR deviation. In contrast, the MLH from TKEDR is robust as long as the accuracy of wind is guaranteed. In the SBL (RL), the CNR is found capable to retrieve SBLH and RLH simultaneously and robustly. This solution is tested during an observation period over one month. Statistical analysis shows that the root-mean-square errors (RMSE) in the MLH, SBLH, and RLH are 0.28 km, 0.23 km, and 0.24 km, respectively.

摘要: 尽管相干多普勒测风激光雷达在边界层高度探测方面有很好的应用前景,但当采用不同的观测数据时,边界层高度的探测结果存在差异。本文提出了一种基于相干多普勒测风激光雷达探测边界层高度的稳健方法:用湍流动能耗散率确定混合层高度,用载噪比确定稳定边界层高度和残留层高度。为了研究边界层探测高度差异的成因,本文采用了两台相同机制的雷达作对比,并为其中一台配备了温控装置。实验发现,由于自由空间到望远镜保偏光纤的耦合效率对环境敏感,仪器不稳定性将导致载噪比误差,进而导致基于载噪比确定的混合层高度出现高达2.13 km的偏移。相比之下,只要风场观测数据是准确的,基于湍流动能耗散率就可以稳健地确定混合层高度。实验还发现载噪比可以稳健地确定稳定边界层高度和残留层高度。为了验证该方法的有效性,本文进行了为期一个月的实验,统计结果表明:混合层高度,稳定边界层高度和残留层高度的均方根误差分别为0.28 km,0.23 km和0.24 km。

1.   Introduction
  • The atmospheric boundary layer (ABL) is the lowest layer of the troposphere that is directly influenced by the earth’s surface (Stull, 1988). In the case of fair-weather days, the ABL has a well-defined structure that evolves with a diurnal cycle, consisting of a mixed layer (ML) in the daytime, and a stable boundary layer (SBL) which is capped by a residual layer (RL) in the nighttime (Coen et al., 2014). The boundary layer height (BLH) is of prime importance to boundary layer parameterization (Li et al., 2017) in weather, climate, and air quality models (Seidel et al., 2010; Li et al., 2016).

    The BLH is difficult to be directly measured. However, at a well-mixed layer top, there are usually significant increases in potential temperature, depolarization ratio, and horizontal wind speed or decreases in relative humidity, aerosol concentration, and turbulence intensity (Li et al., 2017). Based on these characteristics, various BLH definitions and retrieval methods have been developed, such as ideal profile fitting (Steyn et al., 1999), maximum variance (Hooper and Eloranta, 1986), first gradient (Flamant et al., 1997), threshold (Melfi et al., 1985), wavelet transform (Cohn and Angevine, 2000), and the combination method of wavelet transform and image processing (Lewis et al., 2013).

    Numerous instruments have been applied in detecting the BLH, including in situ instruments such as radiosondes mounted on masts or balloons (Guo et al., 2016) and remote sensing instruments such as microwave radiometers (Coen et al., 2014), wind profilers (Cohn and Angevine, 2000), sodars (Emeis et al., 2007), lidars (Yang et al., 2020), and satellites (Luo et al., 2016). The advantages and shortcomings of these instruments are summarized by Seibert et al. (2000). Among these, coherent Doppler wind lidar (CDWL) is widely applied in detecting wind and rain (Wei et al., 2019), turbulence (O’Connor et al., 2010; Banakh et al., 2017; Leung et al., 2018), and ABL classification (Manninen et al., 2018; Yuan et al., 2020). Its high spatial and temporal resolutions make the CDWL one of the most promising instruments in detecting BLH (Wang et al., 2018).

    BLH can be detected by using one of the CDWL measurements as a tracer: for example, horizontal wind speed (Emeis et al., 2008), the aerosol-related parameter such as aerosol backscatter coefficient or carrier-to-noise ratio (CNR) (Peña et al., 2013), or the turbulence-related parameter such as vertical wind speed variance or turbulent kinetic energy dissipation rate (TKEDR) (Vakkari et al., 2015; Huang et al., 2017; Banakh et al., 2020). The aerosol-based BLH is found to be higher than the turbulence-based BLH with an overestimation of several hundred meters both in the morning and the late afternoon due to different BLH definitions (Schween et al., 2014). Besides, differences of up to one kilometer between the two kinds of BLHs are observed, which raises great concern. A reasonable assumption is that the CNR is distorted, leading to an incorrect BLH value. In theory, CNR is an indicator of aerosol concentration and thus can be used to retrieve BLH. In practice, CNR could be distorted by instrument instability. For a single CDWL, it is hard to prove whether the CNR change is a natural variation introduced by aerosol concentration variation or an abnormal deviation caused by instrument instability. Thus, an intercomparison experiment is designed with two identical CDWLs, where only one lidar is equipped with a stability control subsystem. The lidar with stability control can be used as a reference. Note that the instrument stability is difficult to be controlled in a harsh environment or for long-term observations. Even minor instrument instability will result in a BLH bias. Thus, a robust solution is expected to be found.

    In this work, an intercomparison experiment is carried out to analyze the effect of instrument instability on BLH bias. A robust solution for detecting BLH with CDWL is provided. The experimental site, instruments, and data are introduced in section 2. Section 3 describes the BLH retrieval methods. Section 4 presents the lidar observation results and BLH retrieval results. A statistical comparison of the BLHs derived from the two lidars is also performed. Finally, a conclusion is given in section 5. If not specified, LST (Local Standard Time, LST = UTC + 8) is used.

2.   Site, instruments, and data
  • The experiment is conducted in Xilingol, Inner Mongolia, China (43°54′N, 115°58′E) from 28 August to 1 October 2019.

    Two all-fiber micro-pulse CDWLs are located as shown in Fig. 1. The lidars have full hemispheric scanning capability with a rotatable platform. As two D-shaped aspheric lenses are glued together with aligned optical axes, the overlap distance and the blind zone are 1 km and 30 m, respectively. Below 2.2 km, the vertical resolution is set to 26 m for studying the fine structure. Above 2.2 km, the aerosol concentration is usually low, and the vertical resolution is set to 52 m for improving the detection probability. The key parameters are listed in Table 1. The detailed specifications and the error analysis of radial velocity have been described by Wang et al. (2017). Only CDWL1 is equipped with a temperature stabilizing module. It is composed of two thermoelectric coolers (TEC) and a bottle of cryogen which is connected with the heat-conducting pipes wrapping around the telescope. A proportional-integral-differential algorithm is developed to control the TEC and the valve of the cryogen. A package made of a heat-insulating material is used to guarantee the temperature of telescope. The telescope temperature is controlled at 25°C with a fluctuation of 0.5°C.

    Figure 1.  The two CDWLs used in the experiment.

    ParameterValue
    LaserWavelength (nm)1548
    Pulse energy (μJ)100
    Repetition frequency (kHz)10
    TelescopeDiameter (mm)80
    ScannerZenith scanning range (°)0−90
    Azimuth scanning range (°)0−360
    Radial temporal resolution (s)2

    Table 1.  Key parameters of the CDWLs.

    The two lidars operate simultaneously for 12 days and separately for the remaining days of the observation period. During the experiment, CNR, wind vector, and TKEDR are simultaneously measured with a velocity azimuth display scanning mode, which is defined as conical scanning by a laser beam around the vertical axis with a fixed zenith angle of 30° (Sathe and Mann, 2013). The azimuth angle resolution is 5° and the period of one scan is about 144 s. For every radial measurement during one scan, the atmospheric backscatter signal received by the telescope is mixed with a local oscillator, resulting in a radio frequency beating signal. After Fourier transformation, the Doppler shift caused by the radial velocity can be retrieved. The ratio of the beating signal power to noise power over the entire spectral bandwidth is the radial CNR (Wang et al., 2017). The CNR is an average of radial CNRs over one scan. Then, the wind vector is determined from the sine dependence of radial velocity versus azimuth angle. After filtering out the invalid radial velocities, the vector is obtained by achieving the maximum of the filtered sine wave fitting function (Banakh et al., 2010). Finally, TKEDR can be estimated by fitting the azimuth structure function of radial velocity to a model prediction. The algorithm, including error analysis, is demonstrated by Smalikho and Banakh (2017). Note that the accuracy of wind and TKEDR mainly depends on CNR.

3.   BLH retrieval methods
  • The aerosol concentration in the ABL is much higher than that in the free atmosphere, resulting in a decrease in aerosol backscatter signal at the boundary layer top. CNR can be used as a measure of aerosol backscatter. One method for retrieving the BLH from CNR is the Haar wavelet covariance transform (HWCT) method. The Haar function is defined as (Brooks, 2003):

    where $z$ is height, $b$ is the center position of the function, and $a$ is the dilation of the function. The covariance transform of the function is defined as:

    where $f(z)$ is the range-corrected CNR, and ${z_{\rm{b}}}$ and ${z_{\rm{t}}}$ are the bottom and top heights of a selected range. For a given dilation, the local maximum in ${W_f}\left({a,b} \right)$ determines the BLH. Considering the spatial resolution and several tested values, a dilation of 250 m is used.

    Firstly, the CNR profile is range-corrected via multiplying by the square of distance. Secondly, a 7-min block average and a 250-m spatial smoothing are applied to mitigate the interferences from noise and small-scale events, respectively. Figure 2a shows a nighttime range-corrected CNR profile, which is normalized in the range of 0 to 4 km. Figure 2b shows the profile’s HWCT result, in which two local maximum positions (LMP) are found. Multiple LMPs demonstrate the complex aerosol structure, consisting of a lower SBL formed by radiative cooling from the ground and a higher RL containing former mixed layer air. There are usually decreases in aerosol backscatter at the tops of these layers. Based on the positions of the two LMPs, the SBL height (SBLH) of 1.82 km and the RL height (RLH) of 2.86 km are obtained. In the daytime, as shown in Fig. 2f, there is usually a single LMP, by which the ML height (MLH) of 2.81 km is obtained. Thirdly, a Different Thermo-Dynamic Stabilities algorithm is applied to make the BLH results more uniform under the full range of atmospheric stability conditions, especially under SBL (RL) conditions. The physical basis of this algorithm is to check the temporal continuity and vertical coherence of the BLH (Su et al., 2020).

    Figure 2.  The 7-min average profiles centered at 0853 LST 27 September 2019 of CDWL1: (a) Range-corrected CNR after normalization and its (b) HWCT result, (c) original CNR, (d) lg(TKEDR). Centered at 1311 LST 27 September: (e)–(h). The black dots denote the BLHs retrieved by the HWCT method. The blue dots are the BLHs retrieved by the CNR threshold method. The red squares are the BLHs retrieved by the TKEDR threshold method.

    The performance of the HWCT method could be affected by a poorly defined boundary layer top or multiple aerosol layers. For example, RLH could be misclassified as SBLH. Thus, the other CNR threshold method, which is fast and has low uncertainty, is adopted in this study. This method is applied to the original CNR. The values of the thresholds are determined by using the retrieval results of the HWCT method as references. It can be seen from Fig. 3 that the optimal thresholds for SBLH/MLH and RLH are –25 dB and –32 dB, respectively. By testing, the values are also found to be suitable for the other days during this observation period. If not specified, the CNR method discussed in the following sections denotes the CNR threshold method.

    Figure 3.  The linear regression of the BLH retrieved by using different CNR thresholds as a function of that by the HWCT method: (a) Determination coefficient (R2) (black dashed line) and root-mean-square error (RMSE) (red line) for SBLHCNR/MLHCNR. (b) Similar to (a) but for RLHCNR.

  • At a well-mixed layer top, there is an entrainment zone between the ML and the free atmosphere. In this zone, materials are not fully mixed, resulting from a decrease in turbulence intensity, which can be characterized by TKEDR. Thus, for a given averaging time and an appropriate threshold, the TKEDR above the ML top is less than the threshold and vice versa. A threshold of 10−4 m2 s−3 is applied here (Banakh et al., 2020). For example, Fig. 2d gives a TKEDR profile after a 7-min block average, and the MLH result is 0.31 km.

    As shown in Fig. 2h, the profile fluctuation from unexpected noise causes an outlier point, indicating that the retrieval suffers from occasional interference. To exclude the outlier point, a median algorithm is adopted. First, the median ${z_{\rm{m}}}$ of all heights for which TKEDR is less than the threshold is found. Then, the MLH is determined by finding the maximum of all heights that are below ${z_{\rm{m}}}$ and for which TKEDR achieves the threshold. As a result, the MLH of 2.78 km is obtained. The TKEDR method discussed in the following sections refers to the TKEDR threshold method.

4.   Results and discussion
  • The continuous observation results during 27–28 September are shown in Fig. 4, including CNR, horizontal wind speed and direction, vertical wind speed, and TKEDR. The left and right columns are the results of CDWL1 and CDWL2, respectively. The CNRs of the two lidars are quite different. To show the details, the radial profiles are plotted in Figs. 5aj. The radial CNR of CDWL2 is much weaker than that of CDWL1 in the afternoon, and the worst case is found in Fig. 5e at 1400 LST. This phenomenon repeats each day, indicating association with diurnal atmospheric temperature variations. One reasonable explanation is that the optics of the telescope suffer aberrations due to the ambient temperature change, resulting in a loss in heterodyne efficiency (Chambers, 1997). In the CDWL, the atmospheric backscatter signal is coupled from free space into a polarization-maintaining fiber with a core diameter of 9 micrometers. The temperature variation may also change the focusing point due to the thermal expansion of the mechanical structures in the telescope.

    Figure 4.  Lidar observation results during 27–28 September 2019: (a) CNR, (b) horizontal wind speed, (c) horizontal wind direction, (d) vertical wind speed, (e) lg(TKEDR) of CDWL1. (f)–(j) Results of CDWL2. The horizontal wind direction is defined as 0° for northerly wind, rotating clockwise. Negative vertical wind speed denotes rising motion.

    Figure 5.  (a)–(j) Radial CNR profiles at different measurement moments on 27 September. (k)–(t) The corresponding radial velocities and their standard deviation profiles. The △t values denote the moment difference between CDWL1 (blue line) and CDWL2 (red dashed line).

    The corresponding radial velocities and their standard deviation profiles are plotted in Figs. 5kt. Since the measurement moments of the two lidars are not the same (see Figs. 5aj), small radial velocity differences are observed. As the standard deviation mainly depends on the radial CNR, to guarantee the accuracy of wind, the radial velocities with radial CNRs below –35 dB are abandoned (Wang et al., 2017). The accuracy of the wind measurements of CDWL1 has been validated by comparison with a radiosonde, and the mean differences in horizontal wind speed and direction are 0.3 m s−1 and 1.1°, respectively (Wei et al., 2019). Although the radial CNR of CDWL2 is weaker in the afternoon, most values are still valid for measuring the wind within the ABL. As shown in Fig. 6, the mean differences between the two lidars in horizontal wind speed, horizontal wind direction, and vertical wind speed are 0.08 m s−1, –0.16°, and –0.02 m s−1, respectively, indicating the accuracy of the wind measured by CDWL2.

    Figure 6.  Statistical analysis of the wind differences between the two lidars on 27 September: Histogram distributions of (a) horizontal wind speed difference, (b) horizontal wind direction difference, and (c) vertical wind speed difference. The dashed lines denote Gaussian fits.

  • The BLH retrieval results during 27–28 September are shown in Fig. 7. The results are smoothed by finding the median with a 21-min window. The ABL can be classified by the gap between the SBLH from CNR and the MLH from TKEDR. In the morning, the ABL changes from the SBL (RL) into the ML when the gap is less than a specified value. In the late afternoon, the MLH departs from the SBLH, and the ML turns into the SBL (RL) again when the gap is larger than the value (Wang et al., 2019). For example, a specified value of 500 m is used on 27 September as the ML top is about 3 km.

    Figure 7.  BLH retrieval results during 27–28 September: (a) Results of CDWL1. (b) Results of CDWL2. SBLH/MLHCNR (blue dots) and RLHCNR (black open circles) are retrieved by the CNR method. MLHTKEDR (red squares) are retrieved by the TKEDR method.

    The results of CDWL1 are plotted in Fig. 7a. The ABL experiences two obvious diurnal cycles, which evolve as illustrated in Coen et al. (2014). The sunrise and sunset times are marked by upward and downward arrows, respectively. Before sunrise on 27 September, the SBL is found to be capped by the RL. The SBLH and the RLH retrieved from CNR are around 1.5 km and 2.5 km, respectively. After sunrise, the ML develops gradually due to solar radiation and is well-mixed by turbulence. The SBL (RL) is finally merged into the developing ML at about 1200 LST. The MLH can be retrieved from both CNR and TKEDR. In principle, the MLH will keep steady if the temporal gradient of surface temperature is around zero in the afternoon. The fluctuation of the MLH from TKEDR is a result of the turbulence intensity variation related to the cloud-top radiative cooling. The ML departs from the SBL (RL) rapidly before sunset. After sunset, the SBL (RL) develops again. On 28 September, the BLH evolves as it did on 27 September, but the ML top is a little higher than on 27 September. This sequence of BLH evolution is typical for land surfaces in the midlatitudes (Kaimal and Finnigan, 1994). Note that at about 0200 LST 28 September, the sudden rise of the RLH from CNR is caused by an injection of aerosols, and a coinciding wind direction shear from the south to the northwest is observed (see Fig. 4c). This indicates that the retrieval of the BLH based on the vertical distribution of aerosols could be affected by multiple aerosol layers.

    The accuracy of BLH detections of CDWL1 has been validated by comparison with a direct detection lidar (Wang et al., 2019). Thus, the results of CDWL1 are used as references in this study. Figure 7b plots the results of CDWL2. In the daytime on 27–28 September, the MLH from CNR of CDWL2 is much lower than that of CDWL1. This suggests that the performance of the CNR method is affected by the CNR deviation of CDWL2, resulting in the MLH bias. In contrast, since the wind measured by CDWL2 is accurate, as discussed in section 4.1, the accuracy of TKEDR estimated from wind is guaranteed. As a result, robust MLH from TKEDR of CDWL2 is observed.

  • For quantitative analysis, a 12-day statistical comparison of the BLHs derived from the two lidars is performed.

    The scatter diagrams of the MLHs retrieved from CNR and TKEDR are shown in Figs. 8ab. The results of CDWL2 are plotted versus those of CDWL1. The histogram distributions of the MLH bias are inserted in the bottom right corner of each panel. The negative bias means that the result of CDWL2 is lower than that of CDWL1. For the MLH from CNR, there are two gathering regions where normalized densities are larger than 0.5. The low gathering region that lies on the 1:1 line is a result of the calm days during this observation period, because weak convection leads to low ML tops; the high gathering region that deviates from the line results from the turbulent days. Table 2 presents the linear regression. The fit of these points to the 1:1 line shows the worst performance with the smallest determination coefficient (R2) of –2.50 and the largest root-mean-square error (RMSE) of 1.08 km. For the MLH from TKEDR, most points lie on the 1:1 line. Good performance with an R2 of 0.90 and an RMSE of 0.28 km is obtained, indicating the robustness of TKEDR in retrieving MLH.

    Figure 8.  Scatter diagrams of the BLHs derived from the two lidars: (a) MLHCNR and (b) MLHTKEDR in the ML, (c) SBLHCNR and (d) RLHCNR in the SBL (RL). The color-shaded dots denote normalized density. The dashed lines denote 1:1 lines. The y–x is inserted in the bottom right corner of each panel.

    BLHR2RMSE (km)
    MLMLHCNR−2.501.08
    MLHTKEDR0.900.28
    SBL (RL)SBLHCNR0.870.23
    RLHCNR0.950.24
    R2: determination coefficient between dots and 1:1 line; RMSE: root-mean-square error.

    Table 2.  Linear regression of the BLH derived from CDWL2 as a function of that from CDWL1.

    In the SBL (RL), since the CNR deviation of CDWL2 shows a diurnal cycle related to the atmospheric temperature variation, the performance of the CNR method should be less affected in the nighttime. Figures 8cd show the scatter diagrams of the SBLH and RLH retrieved from CNR. A linear regression is also performed as presented in Table 2. As expected, the fit of SBLH to the 1:1 line shows good performance with an R2 of 0.87, while the value for RLH is 0.95. The corresponding RMSEs are 0.23 km and 0.24 km, which are both in the acceptable range.

5.   Conclusion
  • An intercomparison experiment was conducted to analyze the effect of instrument instability on the retrieval of BLH. A robust solution for BLH detections with CDWL was found to overcome the instability: MLH was retrieved from TKEDR, while SBLH and RLH were retrieved from CNR. In the experiment, a diurnal CNR deviation caused by instrument instability was found, which results in a diurnal bias of the MLH from CNR. Although the CNR deviation exists, TKEDR is accurate as long as the accuracy of wind is guaranteed, and thus the MLH from TKEDR is robust. Such an improvement in MLH detections is expected to improve the accuracy of weather and air quality forecast systems. Furthermore, the principal BLH features, including MLH, SBLH, and RLH, can be detected by CDWL robustly, which is significant for model validation and parameterization development.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return