Doppler Lidar Measurements of Turbulent Structure Function over an Urban Area

Similar documents
Estimation of Velocity Error for Doppler Lidar Measurements

SCIENTIFIC REPORT. Universität zu Köln, Germany. Institut für Geophysik und Meteorologie, Universität zu Köln, Germany

Submitted to J. Atmos. Ocean. Technol. 18 January 2010

ACTION: STSM: TOPIC: VENUE: PERIOD:

Lecture 11: Doppler wind lidar

WAVE PROPAGATION AND SCATTERING IN RANDOM MEDIA

18A.6 RAMAN-SHIFTED EYE-SAFE AEROSOL LIDAR (REAL) OBSERVATIONS DURING THE CANOPY HORIZONTAL ARRAY TURBULENCE STUDY (CHATS) Shane D.

Lecture 32. Lidar Error and Sensitivity Analysis

Lecture 20. Wind Lidar (2) Vector Wind Determination

Satellite coherent Doppler wind lidar performance simulation based on CALIOP measurements

Precipitation estimate of a heavy rain event using a C-band solid-state polarimetric radar

Turbulent Dissipation Estimates from Pulse Coherent Doppler Instruments

18B.2 USING THE TLS TO IMPROVE THE UNDERSTANDING OF ATMOSPHERIC TURBULENT PROCESSES

Turbulence Measurements. Turbulence Measurements In Low Signal-to-Noise. Larry Cornman National Center For Atmospheric Research

Windcube TM Pulsed lidar wind profiler Overview of more than 2 years of field experience J.P.Cariou, R. Parmentier, M. Boquet, L.

LIDAR OBSERVATIONS OF FINE-SCALE ATMOSPHERIC GRAVITY WAVES IN THE NOCTURNAL BOUNDARY LAYER ABOVE AN ORCHARD CANOPY

VALIDATION OF GOMOS HIGH RESOLUTION TEMPERATURE DATA WITH THE U. BONN LIDAR AT THE ESRANGE DURING JANUARY AND FEBRUARY 2003

1.5 DOPPLER LIDAR MEASUREMENTS OF VERTICAL VELOCITY SPECTRA, LENGTH SCALES, AND COHERENCE IN THE CONVECTIVE PLANETARY BOUNDARY LAYER

Remote sensing of meteorological conditions at airports for air quality issues

Lecture 23. Lidar Error and Sensitivity Analysis (2)

Cross-Spectral Phase Method for Distinguishing Waves from Turbulence in Single-Point Boundary Layer Flow Measurements

VALIDATION OF MIPAS TEMPERATURE DATA WITH THE U. BONN LIDAR AT THE ESRANGE DURING JULY AND AUGUST 2002

Physics of the Convective Boundary Layer based on Radar/Lidar Profiler measurements and simulation

2.1 OBSERVATIONS AND THE PARAMETERISATION OF AIR-SEA FLUXES DURING DIAMET

Doppler lidar observations of Russian forest fire plumes over Helsinki

A Note on the Estimation of Eddy Diffusivity and Dissipation Length in Low Winds over a Tropical Urban Terrain

ESCI 485 Air/sea Interaction Lesson 2 Turbulence Dr. DeCaria

Long term analysis of convective storm tracks based on C-band radar reflectivity measurements

CLOUD RADAR - INITIAL MEASUREMENTS FROM THE 94GHZ FMCW RADAR

Let us consider a typical Michelson interferometer, where a broadband source is used for illumination (Fig. 1a).

INTRODUCTION TO MICROWAVE REMOTE SENSING. Dr. A. Bhattacharya

11B.3 ADAPTIVE TECHNIQUE TO EXTRACT INTRINSIC INSECTS BACKSCATTER DIFFERENTIAL PHASE FROM POLARIMETRIC SPECTRA

The Shape Slope Relation in Observed Gamma Raindrop Size Distributions: Statistical Error or Useful Information?

A new lidar for water vapor and temperature measurements in the Atmospheric Boundary Layer

Retrieval of the vertical temperature profile of atmosphere from MST radar backscattered signal

Lecture 2. Turbulent Flow

October 1991 J. Wang and Y. Mitsuta 587 NOTES AND CORRESPONDENCE. Turbulence Structure and Transfer Characteristics

Mixing height over London: spatio-temporal characteristics observed by Ceilometer networks

Lei Feng Ben Jong-Dao Jou * T. D. Keenan

CH4 IPDA LIDAR MISSION DATA SIMULATOR AND PROCESSOR FOR MERLIN:

WLS70: A NEW COMPACT DOPPLER WIND LIDAR FOR BOUNDARY LAYER DYNAMIC STUDIES.

EFFECTS OF ALONG-TRACK INTEGRATION ON DOPPLER VELOCITY BIAS WITH A SPACEBORNE CLOUD-PROFILING RADAR

Study of wind variability over Moscow city by sodar

1. Introduction. 2. The data. P13.15 The effect of a wet radome on dualpol data quality

Turbulence and flow in the Large Plasma Device

A ZDR Calibration Check using Hydrometeors in the Ice Phase. Abstract

THE DETECTABILITY OF TORNADIC SIGNATURES WITH DOPPLER RADAR: A RADAR EMULATOR STUDY

Rayleigh-Brillouin Scattering Experiment with Atmospheric Lidar from a Mountain Observatory

Preliminary Study of Single Particle Lidar for Wing Wake Survey. M.Valla, B. Augère, D. Bailly, A. Dolfi-Bouteyre, E. Garnier, M.

DIAGNOSING OBSERVATION ERROR STATISTICS FOR NUMERICAL WEATHER PREDICTION

Active microwave systems (2) Satellite Altimetry * the movie * applications

Preliminary Results from the ATHENA-OAWL Venture Tech Airborne Mission

BIAS ON THE WIND-SPEED GUST CAUSED BY QUANTIZATION AND DISJUNCT SAMPLING OF THE CUP-ANEMOMETER SIGNAL

Is Spectral Processing Important for Future WSR-88D Radar?

Exploring the Atmosphere with Lidars

results, azimuthal over-sampling and rapid update Y. Umemoto 1, 2, T. Lei 3, T. -Y. Yu 1, 2 and M. Xue 3, 4

Note the diverse scales of eddy motion and self-similar appearance at different lengthscales of the turbulence in this water jet. Only eddies of size

Investigation of the Air-Wave-Sea Interaction Modes Using an Airborne Doppler Wind Lidar: Analyses of the HRDL data taken during DYNAMO

5.2 NCAR INTEGRATED SOUNDING SYSTEM OBSERVATIONS FOR VTMX

Large eddy simulation studies on convective atmospheric boundary layer

ESA AO/1-5467/07/NL/HE A SPONTANEOUS RAYLEIGH-BRILLOUIN SCATTERING EXPERIMENT FOR THE CHARACTERIZATION OF ATMOSPHERIC LIDAR BACKSCATTER

EFFECTS OF HORIZONTAL SHEAR OF VERTICAL VELOCITY IN DBS AND SA MEAN WIND ESTIMATES REVEALED BY A COMBINATION OF LES AND VIRTUAL RADAR

Ship-Based Measurements of Cloud Microphysics and PBL Properties in Precipitating Trade Cumuli During RICO

Anisotropic Error Distributions in a Bistatic Doppler Radar System

Ship-Based Measurements of Cloud Microphysics and PBL Properties in Precipitating Trade Cumuli During RICO

Highly Efficient and Anomalous Charge Transfer in van der Waals Trilayer Semiconductors

Aerosol and cloud related products by ESA s Aeolus mission

Measurements and Simulations of Wakes in Onshore Wind Farms Julie K. Lundquist 1,2

Assimilation of Doppler radar observations for high-resolution numerical weather prediction

January 19, 1989 NCAR/TN- 328+STR NCAR TECHNICAL NOTE. Evaluation of Collins WXR-700C-G Radar Performance During the MIST Project

Other examples of signatures of mountain waves in radiosonde observations

Akira Ishimaru, Sermsak Jaruwatanadilok and Yasuo Kuga

Comparison of 3D turbulence measurements using three staring wind lidars and a sonic anemometer

ASCAT B OCEAN CALIBRATION AND WIND PRODUCT RESULTS

A Path to Shed Light on the Windshear. Enhancing the windshear alerting systems in airports by integrating a LiDAR-based system.

Meteorological Measurements made during RxCADRE

Atmospheric Lidar The Atmospheric Lidar (ATLID) is a high-spectral resolution lidar and will be the first of its type to be flown in space.

Chapter 4 Nadir looking UV measurement. Part-I: Theory and algorithm

Investigation of the Air-Wave-Sea Interaction Modes Using an Airborne Doppler Wind Lidar: Analyses of the HRDL data taken during DYNAMO

Boundary Layer Meteorology. Chapter 2

The first tropospheric wind profiler observations of a severe typhoon over a coastal area in South China

In Situ Comparisons with the Cloud Radar Retrievals of Stratus Cloud Effective Radius

1306 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 15

S. L. Durden and S. Tanelli Jet Propulsion Laboratory California Institute of Technology 4800 Oak Grove Dr., Pasadena, CA 91109, USA

Median Filter Effects on Radar Wind Profiler Spectral Noise Statistics

AUTOMATIC MONITORING OF BOUNDARY LAYER STRUCTURES WITH CEILOMETER ABSTRACT

Winds on Horizontal Scans from Doppler Lidar during T-REX

EXPERIMENTAL ASSIMILATION OF SPACE-BORNE CLOUD RADAR AND LIDAR OBSERVATIONS AT ECMWF

10. OPTICAL COHERENCE TOMOGRAPHY

Measurements of a network of mobile radars during the experimental campaign of the HydroRad project

SCIENTIFIC REPORT. Host: Sven-Erik Gryning (DTU Wind Energy, Denmark) Applicant: Lucie Rottner (Météo-France, CNRM, France)

Simultaneous Velocity and Concentration Measurements of a Turbulent Jet Mixing Flow

Active rain-gauge concept for liquid clouds using W-band and S-band Doppler radars

Simultaneous Observations of E-Region Coherent Backscatter and Electric Field Amplitude at F-Region Heights with the Millstone Hill UHF Radar

Large-Eddy Simulations of Tropical Convective Systems, the Boundary Layer, and Upper Ocean Coupling

Flood Forecasting with Radar

6.9 ASSESSMENT OF ACOUSTIC TRAVEL-TIME TOMOGRAPHY OF THE ATMOSPHERIC SURFACE LAYER

P1.17 SENSITIVITY OF THE RETRIEVAL OF STRATOCUMULUS CLOUD LIQUID WATER AND PRECIPITATION FLUX TO DOPPLER RADAR PARAMETERS

Doppler radar spectral width broadening due to beamwidth and wind shear

A HIGH RESOLUTION HYDROMETEOR PHASE CLASSIFIER BASED ON ANALYSIS OF CLOUD RADAR DOPLLER SPECTRA. Edward Luke 1 and Pavlos Kollias 2

Transcription:

MAY 004 DAVIES ET AL. 753 Doppler Lidar Measurements of Turbulent Structure Function over an Urban Area F. DAVIES AND C. G. COLLIER TIES Research Institute, University of Salford, Greater Manchester, United Kingdom G. N. PEARSON QinetiQ, Malvern, Worcestershire, United Kingdom K. E. BOZIER TIES Research Institute, University of Salford, Greater Manchester, United Kingdom (Manuscript received 1 July 003, in final form 8 November 003) ABSTRACT Analysis of radial wind velocity data from the Salford pulsed Doppler infrared lidar is used to calculate turbulent spectral statistics over the city of Salford in the United Kingdom. The results presented here, first, outline the error estimation procedure used to correct the radial wind velocity measurements from the Salford lidar system; second, they correct the data for the spatial averaging effects of the Salford lidar pulse; and finally, they use the corrected data to calculate turbulent spectral statistics. Using lidar data collected from the Salford Urban Meteorological Experiment (SALFEX), carried out in May 00, kinetic energy dissipation rates, radial velocity variance, and integral length scales are calculated for the boundary layer above an urban canopy. The estimates of the kinetic energy dissipation rate from this method are compared to calculations using more traditional spectral methods. The estimates of the kinetic energy dissipation rate for the two methods are correlated and both show an increase in dissipation rate through the day. The procedure followed for the correction of the spatial averaging effects of the lidar pulse shape actually uses the Salford lidar pulse shape profile. 1. Introduction Given current worries over the dispersion of pollutants in urban regions there is a great deal of interest in the structure of the boundary layer in and above these metropolitan areas. The effects of wind and turbulence are critical in dispersing or maintaining the high levels of pollutants produced in our inner city areas. However, traditional methods of measuring these parameters are extremely difficult due to both the practical difficulties associated with the deployment of instruments in urban areas and also due to the large- and small-scale heterogeneity found within these areas. This heterogeneity causes problems with generalizing results taken over limited areas. The eyesafe nature of the Salford Doppler lidar system enables the use of the system in a region of the boundary layer that would otherwise be difficult to monitor. Also, the range and scanning capabilities of the system can be used to probe the atmosphere over very Corresponding author address: Dr. Fay Davies, School of Environment and Life Sciences, Peel Bldg., University of Salford, Greater Manchester M5 4WT, United Kingdom. E-mail: F.Davies@Salford.ac.uk large areas and throughout the depth of the boundary layer. The scanning capability enables the laser beam to be directed, and the system can be altered to examine either the temporal nature of the turbulence by using a fixed beam over a time period or the spatial nature of the turbulence by scanning over a large volume. Measurements of wind velocity taken using a Doppler lidar system differ from traditional point measurements in that they are a spatial average over the volume of the laser pulse, which is a pencil-shaped region with a length of 10s to 100s of meters and a diameter of approximately 0.5 m. Procedures must therefore be carried out to correct the wind velocity data for this spatial averaging. Further corrections must also be made to correct for system noise and systematic biases, as is done with any measurement system. This paper aims to provide measurements for the noise and bias for the Salford lidar system, correct for the inherent spatial averaging of the system (following well-documented procedures), and finally produce turbulence measurements over an urban boundary layer.. The Salford lidar system The first development of the Salford Doppler lidar is discussed in Pearson and Collier (1999). In 00, the 004 American Meteorological Society

754 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 1 TABLE 1. Characteristics of the Salford lidar system. Transmitter Operating wavelength Energy per pulse Pulse repetition frequency Range resolution Min range Max range CO laser 10.6 m 70 mj Up to 50 Hz 11 m 700 m 9000 m system was upgraded to increase the output power to 70 mj by the inclusion of a transverse excitation atmospheric (TEA) pressure laser. This enabled the maximum range to be increased to approximately 10 km, but increased the system minimum range to approximately 900 m. The system pulse repetition frequency (PRF), which was previously 10 Hz, was also reduced, and, although the system for the trials reported in this paper was set at a PRF of approximately 10 Hz, the system is adjustable and can theoretically work up to 50 Hz. Details are given in Table 1. The signal processing, which is discussed fully in Pearson and Collier (1999), used an offline scheme based on correlogram accumulation. Using this procedure the autocorrelation function (ACF) for the return signal for each range gate is calculated, the ACFs are accumulated over the appropriate number of records, and the spectrum is then calculated. The peak of the spectrum is then assigned a velocity. The performance of this type of simple discrete spectral peak estimator is discussed in detail in Rye and Hardesty (1993). In the original system the processing integrated the return signal over 10 shots to obtain atmospheric data at a rate of 1 Hz. The transmitted pulse shape in the old system, as shown in Pearson and Collier (1999, their Fig. 4), was close to Gaussian (having very little of its power in the tail of the pulse). The pulse did not exhibit chirp (Dabas et al. 1998), and the system consequently did not produce a systematic velocity bias. The new system uses the same processing procedure, but now carries out the calculations in real time. The number of pulses over which the return signal is accumulated can be set as a system input. The pulse shape is no longer Gaussian and has a significant amount of power in its tail, as is shown in Fig. 1. The Gaussian model fit, also shown in Fig. 1, contains 85% of the energy of the laser pulse. The pulse exhibits chirp and therefore this has to be considered in the signal processing. 3. Chirp and measurement bias Dabas et al. (1998) showed that a 10-m heterodyne Doppler lidar can produce a biased velocity measurement even when operating under stationary atmospheric conditions. The velocity bias is due to the combination of the atmospheric speckle effect and the asymmetry of the periodogram of the transmitted laser pulse. Dabas FIG. 1. (a) Lidar pulse shape (line) and its associated Gaussian model fit (dashed line) and (b) a typical profile of the frequency chirp. et al. (000) showed that the velocity bias can therefore be considered to be due to a combination of a speckle effect and a constant effect. The constant effect is due to the asymmetry in the pulse spectrum. For the Salford lidar pulse the frequency chirp in Fig. 1b shows a slight downsweep followed by an upsweep. The total magnitude of this frequency sweep is 1.5 MHz. The signal-processing procedure involves the range gating of the returned atmospheric signal. Calculating the convolution of the range gating (i.e., a rectangular window function) with the asymmetric pulse spectrum gives a constant bias due to the frequency chirp of 0.5 m s 1. In addition, bias can be caused by imperfect compensation of pulse-to-pulse variations in the offset frequency. This effect is, however, monitored and corrected for on a shot-to-shot basis. All data considered in this analysis are corrected for the constant bias effect. The atmospheric speckle over the pulse volume causes the return signal to also exhibit a velocity speckle bias, which, in the system described by Dabas et al. (1998), can reach 0.5 m s 1 for a single shot measurement. Dabas et al. (000) note that using a shot accumulation procedure of n shots can divide the speckle bias by a factor of n, when the pulses over the integration are identical. Since the return signal of the Salford lidar system is always integrated over 10s of shots, the speckle bias is only of the order of 0.01 m s 1 under normal atmospheric conditions. This bias is therefore very small compared to the velocity estimation error, which is approximately 0.4 m s 1 as discussed in section 5. The speckle bias is therefore not considered further. Rye and Hardesty (1993) did show, however, that the amount of accumulation for the correlogram accumulation procedure is an important parameter in producing

MAY 004 DAVIES ET AL. 755 TABLE. Details of datasets taken on May 00 using the Salford lidar system. FIG.. Cross section showing the position of the Salford lidar system and lidar beam compared to the underlying average surface characteristics. The mean wind direction is going into the paper. a good velocity estimate. In atmospheric investigations the requirement is for a data measurement rate that is as fast as possible. Point measurements taken by sonic anamometers are typically at the 10s of hertzs rate. Rye and Hardesty (1993) showed that for a broadband signal model, using correlogram accumulation, the number of accumulations (or integrations) should, for optimal velocity retrieval, be of the order of at least a few 10s, with the velocity estimation improving up to their 100- integration limit. The theory discussed by Dabas et al. (000) is limited to situations where the chirp of the laser pulse is up to 1.5 MHz and the signal-to-noise ratio (SNR) is above 0 db. However, due to the nature of wind shear and turbulence, the magnitude of the bias can be increased by the impact of these parameters. In the simulations carried out by Dabas et al. (1998, 000) the authors note, however, that the procedure works well for normal atmospheric conditions [with turbulent kinetic energy (TKE) dissipation rates in the region 10 3 10 m s 3 ], but that the situation with wind shear is more complicated since the biases are larger when the true frequencies increase with range, that is, when the wind strength is increasing with range. 4. Characteristics of the trial data For the initial error analysis, reported upon later in this paper, a series of three sets of data were taken at a fixed elevation angle over the center of Salford, Greater Manchester, in northwest England. The data were taken over a 45-min period starting at 1300 UTC and from an elevated position 0 m above the surface. The elevation angle used, 4.7, was the minimum possible given the local morphology. The data used were therefore in the height range 110 390 m above ground level, and over a horizontal distance of 3400 m. The ground level in this region, which includes both the city center and the River Irwell flood plane, varies in height by only 3 m. The city morphology is a mixture of low-rise ( stories) and high-rise (up to 0 stories) buildings with an industrial area to the south. The cross section shown in Fig. outlines the main characteristics of the site, but details of the average building height were not included since the minimum measurement of the lidar was well above the roughness sublayer, that is, above the effects of the individual buildings (Roth 000). The lidar beam was pointed into the direction of the city center, Dataset May 0-1 May 0- May 0-3 No. of shots accumulated 80 50 50 PRF (Hz) 9.09 9.09 1.5 and the mean wind direction was perpendicular to the beam direction. The turbulence was found to be well mixed within the height range analyzed and the urban boundary layer was considered to have sufficient fetch (several kilometers) to be adapted to the surface conditions on the large scale. The data were taken with different system inputs as detailed in Table. The system SNR was in the range 1 to 6 db for the measurements used in this analysis. Since a certain amount of averaging has been done within the signal processing, we need to understand how much spatial averaging this entails. A coherent Doppler lidar measures the return signal from a pencil-shaped region of random scatterers (aerosol particles). The lidar beam is collimated, and the Salford pulsed Doppler lidar (PDL) has a transverse spatial dimension of only 1.0 m at a range of 4.5 km. For a fixed instant in time, the spatial extent of the lidar pulse along the transmit axis, r, given by the full-width half-maximum (FWHM) of the Gaussian profile in Fig. 1a, is 75 m. For a radial velocity estimate from one lidar pulse the illuminated aerosol region travels a distance, p, of 11 m. Therefore, the estimated velocity for a single pulse is a spatial average of the instantaneous radial velocity (r, t) over the sensing volume, where r is the range along the transmit axis and t is the time of measurement. For multiple pulse measurements the total measurement time (dwell time), following Frehlich and Cornman (00), is given by N p T, (1) PRF where N p is the number of accumulated pulses and PRF is the pulse repetition rate of the system. If the mean wind velocity is moving through the beam in a transverse direction, then the lidar samples a parallelogram of the atmosphere in a plane defined by the fixed laser beam and the fixed mean velocity vector. The transverse distance sampled by the lidar for each velocity estimate will be h V H T, where V H is the mean transverse velocity. The velocity estimates should behave like the single-pulse estimates when the horizontal distance h is much less than the range gate length p (Frehlich and Cornman 00). With the Salford lidar system this is the case, since as can be seen from Table 3, h/p is at a maximum of 0.4. For the data in this analysis the transverse velocity, V H,was3ms 1 giving values for h and h/p as shown in Table 3. Frehlich and Cornman (00) used simulated PDL

756 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 1 TABLE 3. Transverse distance sampled by the lidar h, and the ratio of the transverse distance to the range gate length h/p. Dataset h h/p May 0-1 May 0- May 0-3 6.4 16.5 1.0 0.4 0.14 0.11 data to estimate the effects of spatial averaging on derived atmospheric quantities. Using calculations of the spatial structure function, they estimated the energy dissipation rate,, and the integral length scale, L i, for approximately 1000 independant realizations of a simulated wind field. This procedure will be followed to analyze data from the Salford pulsed Doppler lidar in the next section. Probability density functions (PDFs) were plotted (Frehlich and Cornman 00) to show the variability of the estimated and L i from the simulated data. These calculations were carried out for different values of the parameter h/p to show the effects of different amounts of spatial averaging. The data simulation procedure was for data from a -m PDL where the number of accumulated pulses, N p, was 0. Frehlich and Cornman found that for the calculations of the estimates were within 5% of the input value for situations where h/p was in the range 0.1 0.5 (for the velocity differencing error correcting procedure to be discussed in the following section). However, for the calculations of L i, the PDFs were shown to be skewed to large values. For an input L i of 100 m they showed that for h/p in the region 0. 0.3 the 1 (calculated from the median of the PDF) was approximately 30 m. For lower values of h/p in the region of 0.1, the standard deviation was much larger (100 m for 1). To summarize, Frehlich and Cornman note that the spatial averaging of the velocity field in the horizontal direction of the lidar beam axis produces a negligible bias when h/p 0.. Using the Salford PDL data and following Frehlich and Cornman (00), we can calculate the structure function for the May 00 data and subsequently produce estimates of and L i. 5. Error analysis The estimation of the retrieved radial velocities from the PDL are a function of the lidar parameters, the atmospheric conditions, and the velocity estimation procedure. It is therefore important to know what is the combined effect of the lidar parameters and the velocity estimation algorithm when attempting to measure the atmospheric conditions with a PDL system. Following Drobinski et al. (000) the wind velocity V r (r, t), at range r and time t, measurement by a PDL can be written as the sum of an effective wind velocity, V m (r, t) and an error, e(r, t): V r(r, t) V m (r, t) e(r, t), () where, according to Frehlich et al. (1998), the effective wind velocity, V m (r), is a spatial average over the pulse duration and processing range gate of the true line-ofsight (LOS) wind velocity, V t (r). Drobinski et al. (000) used a model that assumed the errors have a zero mean, which is not true in reality due to the asymmetry of the power of the transmitted pulse (Dabas et al. 1998) and low SNR (Frehlich and Yadlowsky 1994). Frehlich (001) corrects the equation to take into account any bias due to the assumed errors having a nonzero mean as follows: V r(r, t) V m (r, t) e(r, t) bias(r,t). (3) A measurement of the bias of a PDL measurement requires an independent in situ measurement. The random error can be determined by 1) comparing with in situ measurements, ) using the statistical properties of the velocity estimator and assuming ideal atmospheric conditions (Drobinski et al. 000), or 3) using actual radial velocity data as outlined in Frehlich (001). Frehlich (001) described three different techniques for deriving the error, e(r, t), from actual PDL data. He estimates the error from the velocity spectrum from the velocity covariance and also from a velocity differencing procedure. In this paper we follow the technique of velocity differencing that Frehlich (001) shows to have the best performance and the smallest estimation error. From the radial velocity measurements from N lidar data points, Frehlich (001) produced two velocity estimates, ˆ e and ˆ o, from the even- and odd-numbered data, respectively. Each estimate will have values from N/ data points. For a fixed beam the estimates should have the same mean, V m (r, t), and the same bias (for well-behaved atmospheric conditions). Hence, N N N Vr r, t, ˆ e r, t, ˆ o r, t,. (4) From Eq. (3), N N N Vr r, t, ee r, t, eo r, t,, (5) and if e e and e o are statistically similar and uncorrelated, the variance is then given by Therefore, N N V r, e r,. (6) e 1 N V e (r, N) r,. (7) 4 Figure 3 shows the calculation for e for the three sets of data taken using the Salford PDL following the methodology of Frehlich and Cornman (00). The fig-

MAY 004 DAVIES ET AL. 757 A simple model for the structure function is s D (s), (9) L o V where is the variance of the radial velocity and L o is the outer scale of the turbulence. For the von Kármán model (Hinze 1959; Frehlich 000), 1/3 (x) 1.0 0.59 548 5 x K 1/3 (x), (10) where K 1/3 (x) is the modified Bessel function of order 1/3, and Li 0.746 834 3Lo and (11) FIG. 3. Estimation error for the Salford Doppler lidar calculated from the velocity differencing procedure for the three datasets detailed in Table. ure shows that the error, e, increases with range, as would be expected, due to decreasing SNR with range. Above 4480 m (range gate 40), the error, e, is affected by bad estimates of radial velocity and data beyond this point are not used in the analysis. Frehlich (001) notes that the velocity differencing algorithm can be modified for data in the weak signal regime, which is not carried out here. A comparison of the curves in Fig. 3 shows that the error, e, is increased when the number of pulse accumulations decreases since May0-0 displays larger errors than May0-01. The May0- data and May0-3 data are on average similar in magnitude. 6. Data analysis The spectral statistics of the velocity field can be described by use of the spatial structure function. The spatial structure function is primarily described by two parameters: the energy dissipation rate,, and the integral length scale, L i. Frehlich and Cornman (00) showed how calculations of the structure function of simulated PDL radial velocity data for different values of the parameter h/p, which had been corrected for the random error due of the lidar data, e, can be fit to the von Kármán turbulence model for isotropic turbulence in the inertial subrange (Hinze 1959; Frehlich 000) to derive values for the energy dissipation rate,, and integral length scale, L i. Following Frehlich and Cornman (00), for stationary conditions the structure function of V(r, t) is independant of t and is defined as D V(s) [ (r 0) (r0 s)], (8) where s is the spatial separation between radial velocity measurements,, and denotes an ensemble mean. 3 L o 0.933 668, (1) where is the energy dissipation rate. The structure function from the PDL data is denoted by Dˆ wgt, and is calculated from ˆD wgt(r 1, r ) [V m(r 1, t) V m(r, t)]. (13) For a stationary and homogeneous wind field, D ˆ (r, r ) D ˆ wgt 1 wgt(r1 r ), (14) and when the estimation error e(r, t) is uncorrelated with the wind velocity V m (r, t), then ˆ ˆ D (r r ) D (r r ) ˆ (r r ), (15) wgt 1 raw 1 e 1 where Dˆ raw is the uncorrected velocity structure function [from Frehlich and Cornman (00), their Eq. (7)], and ˆ e (r 1 r ) is the unbiased correction for the estimation error e(r, t) of the velocity estimate using the velocity differencing error procedure described previously. Frehlich et al. (1998) showed that the simple model for the structure function given in Eq. (9) can be altered to correct for the effect of spatial averaging of the lidar pulse volume. The corrected model is given by s p lnp D wgt(s,, L o) G,,, (16) p L r where G(m,, ) o F(x, )[( m x ) ( x ) dx (17) and F(x, ), given by Frehlich (1998), is the transfer function describing the Gaussian lidar pulse. Using the normalized pulse shape from Fig. 1a, the transfer function describing the spatial averaging produced from the Salford lidar can be calculated. Then, G(m,, ) and D wgt (s,, L o ) can be calculated numerically for the simple von Kármán model. Following the calculation of the correction for the estimation error, e, from Eq. (7) and using the Salford lidar data taken on May 00, the corrected structure function was calculated from Eq. (15). Figure 4 shows

758 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 1 FIG. 5. Virtual potential temperature derived from radiosonde data on May 00 in Salford city center. using Eqs. (11) and (1) we can calculate values for L i and, respectively. To fit the model structure function curve to the lidar data structure function, a weighted least squares fit procedure was followed. The parameters of the wind field, and L o, were determined by minimizing the weighted error J as shown below: J (, L o) N s ˆ I wgt wgt o k1 N (ks)[d (ks) D (ks,, L )], (18) where s is the range separation between adjacent velocity estimates, N s is the maximum number of lags, and N I is the number of independent samples of the velocity difference at separation ks. The function was minimized for and L o separately in an iterative process. FIG. 4. Structure function curves for (a) May0-1, (b) May0-, and (c) May0-3. Curves show calculations of the structure function for the Salford Doppler lidar Dˆ raw (s)(), the corrected structure function Dˆ wgt (s) (diamonds), the von Kármán model D V(s) (dashed line), and the corrected von Kármán model D wgt (s) (solid line). the calculation of the structure function Dˆ raw (s), and the corrected data Dˆ wgt (s). Using Eq. (9) the structure function, D V(s), for the von Kármán model was calculated. Following this the corrected structure function, D wgt (s), from the von Kármán model was calculated from Eq. (16). This curve corrects for the spatial averaging of the Salford lidar pulse. The best-fit curve from the corrected von Kármán model to the corrected structure function from the Salford lidar is also shown in Fig. 4. From the fit to the corrected von Kármán model structure function, D wgt (s), we can derive values for and L o. Then 7. Results The new Salford University pulsed Doppler lidar system was first deployed in an urban field experiment in Salford city center in the United Kingdom from to 7 May 00. The first day of the experiment was intended to establish the system performance capabilities and take measurements to calculate the system errors. This paper is intended to address the question of the size of the system biases and the measurement accuracy. Three sets of data from May 00 were analyzed. The weather on May 00 was convective with cumulus clouds developing through the day. Figure 5 shows the virtual potential temperature from a series of radiosonde ascents showing the boundary layer to be unstable through its depth. The top of the mixed layer at 14 UTC can be seen at a height of 1800 m. The three sets of data discussed in this report were taken for three different system configurations, as detailed in Table. The three datasets were taken in a 45-

MAY 004 DAVIES ET AL. 759 TABLE 4. Mean radial veloctiy, mean radial velocity variance, and estimation error, e, averaged over the line of sight of the lidar beam. Dataset May 0-1 May 0- May 0-3 Mean velocity (m s 1 ) 0.41 0.14 0.94 Velocity variance (m s ) 1.0 1.41 1.66 e (m s ) 0.16 0.3 0.5 min period starting from 1300 UTC. The aim was to calculate the system sensitivity for the three different setups. Following Frehlich (001) a technique of velocity differencing was used to calculate the size of the estimation error. Figure 3 shows the standard deviation of the estimation error as a function of range. Table 4 shows the average estimation error variance, e, for ranges 10 40. It is seen that the error is increased for the May0- and May0-3 datasets where the amount of accumulation of shots is decreased. The difference in e between May0- and May0-3 is not significant. Each dataset comprises approximately 10 min of data, and the mean and trend were removed before the data were analyzed. The lidar beam was pointed to be perpendicular to the mean wind direction at the start of the experiment, as measured using an azimuth scan from the lidar (not shown). The lidar beam was inclined at an angle of 4.7 from the horizontal, which provides data in the height range 110 390 m. Table 4 gives the mean radial velocity and velocity variance through the profile. The increase in the mean radial velocity seen in May0-3 is possibly due to a slight change in the mean wind direction. The increase in velocity variance is due to the increasing turbulence apparent on the day. Following the procedure outlined in section 6, calculations were carried out to measure the kinetic energy dissipation rate from the calculation of the structure function. The calculation of the structure function accounted for the non-gaussian shape of the Salford lidar pulse. This is a modification of the method outlined in Frehlich et al. (1998), which assumed a Gaussian pulse shape. Curves for the structure function for the three sets of data are shown in Fig. 4. The points in Figs. 4a c are from calculations from the lidar data while the lines are from the von Kármán model. It can be seen from the figure that the corrected lidar data and corrected von Kármán model data are a good fit for separations of between 300 and 1500 m. The difference between the uncorrected von Kármán model, D V(s), and the corrected model, D wgt (s), shows the importance of spatially FIG. 6. Correlation spectra for the three datasets detailed in Table, with an example 5/3 line to highlight the inertial subrange in the spectra. correcting the data for both size and shape of the pulse length. Table 5 shows the kinetic energy dissipation rate, integral length scale, and velocity variance from the fitted model curves. The eddy kinetic energy dissipation rate is seen to increase throughout the duration of the experiment. This is typical behavior for a convective boundary layer where normally increases through the morning and early afternoon before slowly decreasing in the late afternoon (Stull 1988). The values are also typical for a convective boundary layer (Stull 1988). Frehlich and Cornman (00) show that s error for this method for h/p in the range 0.1 0. is approximately 1%. Table 5 also shows kinetic energy dissipation rates obtained from the amplitude of the power spectra of the radial velocity correlations. The methodology, outlined by Gal-Chen et al. (199), is used to filter and spectrally analyze the data. Data for range gates 10 40 are used. The spectra for the three datasets are shown in Fig. 6. The values for the kinetic energy dissipation rate from the covariance spectrum correlate well with those from the structure function curves. However, the dissipation rate calculated from the covariance spectrum is slightly larger. This is due to the fact that using the covariance spectral method means no correction is carried out to remove the variance due to the system estimation error,. The deviation of the corrected model fit, D wgt(s), to TABLE 5. Eddy kinetic energy dissipation rates calculated from the covariance spectrum and from the structure function. Also given are the integral length scale and from the fitted von Kármán model curve. Dataset Kinetic energy dissipation rate from covariance spectrum (m s 3 ) Kinetic energy dissipation rate from structure function (m s 3 ) Intergral length scale from structure function (m) Corrected velocity variance, (m s ) May 0-1 May 0- May 0-3 1.1 10 3.3 10 3 4.3 10 3 0.9 10 3. 10 3 3. 10 3 375 90 355 0.63 0.94 1.45

760 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 1 the Salford lidar data, Dˆ raw (s), at very small (s 100 00 m) and very large separations (s 000 3000 m) is possibly due to the pulse-to-pulse variability in the pulse shape, duration, and chirp that is present in CO lidars (Dabas et al. 000). From Fig. 4 it can be seen that at the smallest separations, the corrected model curve is below that for the lidar data, suggesting that there is more coherence to the lidar data at these separations than is calculated from the corrected von Kármán model. Values for the integral length scale can also be obtained from the fit of the von Kármán model. Frehlich and Cornman (00) note however that for h/p in the range 0.1 0. the estimate can be from 5% smaller to 100% larger than the actual value. 8. Conclusions The aim of this initial data analysis was to address the questions regarding instrument bias and estimation error in the Salford lidar system. The Salford pulsed Doppler lidar system was designed and built by QinetiQ [formerly the Defence and Evaluation and Research Agency (DERA), in Malvern, United Kingdom], specifically to be used in urban areas. It is unique in its system setup and is one of only a handful of purposebuilt atmospheric Doppler lidar systems around the world. It was therefore important to investigate and document the specific system biases and errors. The Salford lidar system was successfully deployed in an urban area: city center Salford, as part of the Salford Urban Meteorological Experiment (SALFEX) in May 00. It was used to profile the wind and turbulence above the urban canopy region. This paper outlines the calculation procedure used to estimate the system biases and estimation errors. A bias due to the constant effect of the frequency chirp has been calculated to be 0.5 m s 1. This bias was removed from all data before further analysis was done. Because of the shot accumulation procedure used in the processing algorithm the bias due to the speckle effect of the frequency chirp is only of the order of 0.0 m s 1 under normal atmospheric conditions. This is well within the region of the velocity estimation accuracy, shown in Fig. 4, which is of the order of 0.4 m s 1. The effects of the speckle bias were therefore not considered further. The velocity estimation error is calculated using a method of velocity differencing discussed by Frehlich (001). Frehlich and Cornman (00) used simulated lidar data to calculate the accuracy of fitting a simple von Kármán turbulence model to the structure function derived from the lidar data. They show that values for the kinetic energy dissipation rate, integral length scale, and velocity variance can be successfully obtained using their method. The methodology corrects for the estimation error and the spatial averaging inherent in the lidar data processing procedure and should therefore give more accurate values for the derived quantities. Frehlich and Cornman (00) assumed a Gaussian profile to their lidar pulse, but the Salford lidar pulse is not Gaussian, so their procedure was modified to take into account the actual Salford lidar pulse profile. Following the procedure outlined by Fredlich and Cornman (00), data from the SALFEX experiment was used to calculate the kinetic energy dissipation rate, integral length scale, and velocity variance above the urban canopy. The calculations were done for data taken over 10-min periods and spatially averaged, for heights in the range 110 390 m, and over a horizontal length of 3.5 km above a mixed urban canopy region. For the three datasets analyzed the kinetic energy dissipation rate was found to increase through the day. This is normal behavior for a convective boundary layer where the dissipation rate is normally seen to increase through the morning and early afternoon (Stull 1988). Frehlich and Cornman (00) show that the error on the kinetic energy dissipation rate,, using this methodology is approximately 1%. The dissipation rate was also estimated from the correlation spectrum. These two results are very similar. The structure function estimation is lower, as would be expected due to the correction for lidar velocity estimation error and spatial averaging. The values for the velocity variance reflect those of the dissipation rate in that they increase during the day and are higher in the uncorrected data. Again the lower corrected values are due to the error estimation correction procedure. The values for the integral length scale are of the order of 50 400 m and are therefore only of the order to 4 times the length of the lidar pulse. Frehlich and Cornman (00) note that even with their simulated data the magnitude of the estimated integral length scale can be 100% larger than the input value. However, despite these reservations, lidar systems are unique in that they do actually measure the spatial variability along the line of sight of the beam and therefore no assumptions of frozen turbulence need be applied. Furthermore, values of parameters such as the integral length scale, kinetic energy dissipation rate, as well as integral time scale are important inputs into standard dispersion models. It is therefore important to establish estimation procedures for the parameters described above: the kinetic energy dissipation rate, the velocity variance, and the integral length scale. In principle the Salford PDL can measure these parameters from the top of the urban canopy to the top of the urban boundary layer, but as with any field experiment, local morphology and atmospheric conditions normally dictate the actual range of the system. The vertical resolution of the measured and derived quantities is dependent on the elevation angle of the beam, and the spatial and temporal averaging required. Acknowledgments. The SALFEX experiment was partly funded within the NERC URGENT Urban Meteorology Project. The University of Salford is a member of the U.K. Universities Weather Research Network

MAY 004 DAVIES ET AL. 761 (UWERN), a constituent of the NERC Centres for the Atmospheric Science (NCAS). The authors would like to thank the staff at QinetiQ Malvern for their technical support and the Staff of the U.K. Met Office (Cardington) for providing the radiosonde balloon ascents during the SALFEX experiment. The authors would also like to thank the anonymous referees for their useful and constructive comments. REFERENCES Dabas, A. M., P. Drobinski, and P. H. Flamant, 1998: Chirp-induced bias in velocity measurements by a coherent Doppler CO lidar. J. Atmos. Oceanic Technol., 15, 407 415.,, and, 000: Velocity biases of adaptive filter estimates in heterodyne Doppler lidar measurements. J. Atmos. Oceanic Technol., 17, 1189 10. Drobinski, P. A., A. M. Dabas, and P. H. Flamant, 000: Remote measurement of turbulent wind spectra by heterodyne Doppler lidar techniques. J. Appl. Meteor., 39, 434 451. Frehlich, R., 000: Simulation of coherent Doppler lidar performance for space-based platforms. J. Appl. Meteor., 39, 45 6., 001: Estimation of velocity error for Doppler lidar measurements. J. Atmos. Oceanic Technol., 18, 168 1639., and M. J. Yadlowsky, 1994: Performance of mean frequency estimators for Doppler radar and lidar. J. Atmos. Oceanic Technol., 11, 117 130; Corrigendum, 1, 445 446., and L. Cornman, 00: Estimating spatial velocity statistics with coherent Doppler lidar. J. Atmos. Oceanic Technol., 19, 355 366., S. Hannon, and S. Henderson, 1998: Coherent Doppler lidar measurements of wind field statistics. Bound.-Layer Meteor., 86, 33 56. Gal-Chen, T., M. Xu, and W. L. Eberhard, 199: Estimations of atmospheric boundary layer fluxes and other turbulence parameters from Doppler lidar data. J. Geophys. Res., 97 (D17), 18 409 18 43. Hinze, J. O., 1959: Turbulence: An Introduction to Its Mechanism and Theory. McGraw-Hill, 586 pp. Pearson, G. N., and C. G. Collier, 1999: A compact pulsed coherent CO laser radar for boundary layer meteorology. Quart. J. Roy. Meteor. Soc., 15, 703 71. Roth, M., 000: Review of atmospheric turbulence over cities. Quart. J. Roy. Meteor. Soc., 16, 941 990. Rye, B. J., and R. M. Hardesty, 1993: Discrete spectral peak estimation in incoherent backscatter heterodyne lidar. II. Correlogram accumulation. IEEE Trans. Geosci. Remote Sens., 31, 8 35. Stull, B. R., 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic, 647 pp.